mathematics

mathematics
/math'euh mat"iks/, n.
1. (used with a sing. v.) the systematic treatment of magnitude, relationships between figures and forms, and relations between quantities expressed symbolically.
2. (used with a sing. or pl. v.) mathematical procedures, operations, or properties.
[1350-1400; ME mathematic < L mathematica (ars) < Gk mathematikè (téchne) scientific (craft), equiv. to mathemat- (s. of máthema) science, knowledge + -ike, fem. of -ikos -IC; see -ICS]

* * *

Science of structure, order, and relation that has evolved from counting, measuring, and describing the shapes of objects.

It deals with logical reasoning and quantitative calculation. Since the 17th century it has been an indispensable adjunct to the physical sciences and technology, to the extent that it is considered the underlying language of science. Among the principal branches of mathematics are algebra, analysis, arithmetic, combinatorics, Euclidean and non-Euclidean geometries, game theory, number theory, numerical analysis, optimization, probability, set theory, statistics, topology, and trigonometry.

* * *

▪ 1995

      A year of ups and downs for mathematics, 1994 began with the awareness of a serious gap in Andrew Wiles's proof of Fermat's last theorem. In June 1993 Wiles, a Princeton University mathematician, had claimed a proof, by relating the problem to a deep conjecture in algebraic number theory, of Pierre de Fermat's famous 350-year-old assertion that xn + yn = zn has no solutions for which x, y, and z are whole numbers if n is three or more. The gap emerged in late 1993 in a review of the proof and lingered until October 1994, when Wiles sent colleagues a revised proof that he hoped would finally lay the problem to rest. At year's end the jury remained out on the validity of the new proof.

      More encouraging news was the solution, by Krystyna Kuperberg of Auburn (Ala.) University, of the Seifert conjecture, a problem about the topology of dynamic systems. According to the celebrated hairy ball theorem, it is impossible to comb a hairy ball smoothly; somewhere at least one hair must stand up on end. The theorem is one of dynamics because such an arrangement of "hairs" is a description of the way the states of a system, represented by points on the sphere, change with time, or flow along the directions of the hairs. The hairy ball theorem was proved long ago, but its higher-dimensional cousins have been more elusive. The most notorious is a question asked in 1950 by Herbert Seifert of the University of Heidelberg, Germany, about a three-dimensional analogue of the surface of a sphere—a kind of curved space called a 3-sphere. If one defines a "flow" by filling the 3-sphere with curved lines that fit smoothly together like the flow lines of a fluid, most of the flow lines will wander around in a complicated way. Occasionally, however, one of them may close up into a loop. It is easy to find a flow with only two closed loops but, as Seifert asked, can one be found with no closed loops at all?

      Kuperberg's surprising answer is "yes." All earlier approaches to the problem had used the same basic idea, that of inserting a "plug" into a given flow to change it and eradicate a selected closed loop. The plug is a small region of flow lines in which some of the lines that enter the region never exit but get trapped inside. Specifically, one starts with a known flow that has only two closed loops and then removes them both by inserting two plugs that trap the lines of the loops and render them no longer closed. The central difficulty is to ensure that no new loops are created inside the plugs themselves. Kuperberg succeeded with a seemingly outrageous idea; she made each plug "eat its own tail" like a snake so that closed loops get trapped in a kind of infinite regress. The solution is a geometric gem, and it changes forever mathematicians' most basic ideas about dynamics in three dimensions.

      On the borderline with mathematical physics but clearly on the mathematical side came a fundamental breakthrough in the quantum mechanics of many-particle systems. According to quantum mechanics, the electrons of an atom can occupy only a discrete sequence of energy levels. In particular, there is a minimum energy level, the ground state, below which an atom cannot go. In effect, the ground state is a barrier that prevents atoms from evaporating. In 1981 the U.S. physicist Julian Schwinger (see OBITUARIES (Schwinger, Julian Seymour )) devised an accurate approximation for the way in which the ground-state energy of an atom varies with atomic charge; i.e., with the number of protons in the nucleus. His conjectured formula for the ground-state energy of an atom having charge Z is approximately -aZ7/3+ 1/8 Z2-bZ5/3, in which a and b are particular constants. In the past year Schwinger's conjecture was given a rigorous proof by Charles Fefferman of Princeton University and L.A. Seco of the University of Toronto. Their achievement represented an important step toward a more nearly complete understanding of the way in which chemistry derives from the laws of quantum mechanics. The next step would be to extend the work from atoms to molecules.

      New mathematics does not have to be complicated and technical; it can also be based on very simple ideas. Near the end of the year, Charles Radin of the University of Texas at Austin published a very strange tiling of the plane: a finite set of tiles that can be assembled only in a highly complex way. Most simple tilings are periodic, repeating the same basic unit over again at regular intervals. In 1961, while investigating questions about decidability in mathematical logic, the philosopher Hao Wang introduced the idea of aperiodic tiles, which can cover the plane but not in any periodic way. Radin's tiles, which are based on an idea of John Horton Conway of Princeton, are aperiodic. In fact—and this is the great novelty—the tiles must appear in infinitely many orientations. Lying at the heart of this exotic tiling is a simple right triangle, formed from a domino cut in half along a diagonal. (IAN STEWART)

      This updates the articles analysis; atom; Euclidean geometry; number theory.

▪ 1994

      The outstanding mathematical event of the year, of the decade, and perhaps even of the century was the announcement in June 1993 of a proof for Fermat's last theorem by Andrew Wiles (see BIOGRAPHIES (Wiles, Andrew John ))—a quiet, rather diffident Englishman working at Princeton University. Mathematics has several notorious unsolved problems, and the puzzle posed by the French number theorist Pierre de Fermat over 350 years ago is one of the most notorious of them all.

       Number theory—the study of the deeper properties of whole numbers—goes back to Diophantus of Alexandria, who flourished about AD 250 and wrote a book called the Arithmetica. It included a completely general construction for Pythagorean triples: three whole numbers that can represent the lengths of the sides of a right triangle by satisfying the Pythagorean equation x2 + y2 = z2. Examples are 32 + 42 = 52 and 52 + 122 = 132.

      Some time around 1637 Fermat wondered what would happen if squares are replaced by cubes or higher powers. In other words, are there any solutions in whole numbers of the "Fermat equation" xn + yn = zn if n is three or more? He found none, and in his copy of the Arithmetica he made the most famous marginal note in the history of mathematics: "To resolve a cube into the sum of two cubes, a fourth power into two fourth powers, or in general any power higher than the second into two of the same kind, is impossible; of which fact I have found a remarkable proof. The margin is too small to contain it." This statement became known as his "last" theorem, because for many years it was the only assertion of his that had been neither proved nor disproved by his successors. In 1847 the German mathematician Ernst Kummer invented the theory of "ideal numbers" and proved Fermat's last theorem for all powers up to 100, except for 37, 59, and 67. By early 1993 similar methods had proved all cases up to the four millionth power—but none of these efforts suggested a way to prove the theorem for all powers.

      Meanwhile, in 1922, the English mathematician Leo Mordell had noticed a curious connection between geometry and number theory. If instead of looking at whole-number solutions of an equation, one looks at all solutions, the results can be visualized as a geometric surface, which has a number of holes—like the hole in a doughnut. He observed that if the number of holes in this surface is two or more, then the corresponding equation seems to have only finitely many integer solutions. This idea became known as the Mordell conjecture, and it was proved in 1983 by the young German mathematician Gerd Faltings. It implies that Fermat's last theorem is nearly true: there are only finitely many solutions for any given power n greater than two. This comes close, but Fermat had conjectured that there are no solutions at all.

      The final resolution of the puzzle rests on a beautiful idea that lies at the heart of modern number theory, that of an elliptic curve, an equation of the form y2 = x3 + ax + b, in which a and b are constants. The number theory of elliptic curves is very well understood. Some unsolved problems still exist, however, and the biggest is called the Taniyama-Weil conjecture. It states that every elliptic curve can be described in terms of modular functions—esoteric relatives of trigonometric sines and cosines.

      Early in the 1980s Gerhard Frey of the University of the Saarland, Saarbrücken, Germany, made a crucial connection between Fermat's last theorem and elliptic curves. Suppose, for the sake of argument, that there does exist a solution to the Fermat equation. If some logical contradiction can be deduced from this supposition—any contradiction—then the hypothetical solution cannot exist, and Fermat's last theorem must be true. Frey considered a particular elliptic curve defined in terms of such a hypothetical solution and discovered that it would have an extremely unlikely combination of properties. In 1986 Kenneth Ribet of the University of California at Berkeley proved that if the Taniyama-Weil conjecture is true, then Frey's elliptic curve not only is unlikely but also cannot exist at all.

      Everything thus rested on the Taniyama-Weil conjecture, and Wiles decided to tackle this very difficult key problem in number theory. In a 200-page paper he marshaled enough powerful mathematical machinery to prove one special case of the Taniyama-Weil conjecture, valid for "semistable" elliptic curves, and that was enough to imply Fermat's last theorem. At year's end a prepublication review of the paper uncovered a possible snag, but Wiles expressed confidence that he could clear it up in the near future. A number of his colleagues, too, stated that it would be premature to conclude that Wiles's proof was in trouble. (IAN STEWART)

      This updates the articles Euclidean geometry; number theory.

* * *

Introduction

      the science of structure, order, and relation that has evolved from elemental practices of counting, measuring, and describing the shapes of objects. It deals with logical reasoning and quantitative calculation, and its development has involved an increasing degree of idealization and abstraction of its subject matter. Since the 17th century, mathematics has been an indispensable adjunct to the physical sciences and technology, and in more recent times it has assumed a similar role in the quantitative aspects of the life sciences.

      In many cultures—under the stimulus of the needs of practical pursuits, such as commerce and agriculture—mathematics has developed far beyond basic counting. This growth has been greatest in societies complex enough to sustain these activities and to provide leisure for contemplation and the opportunity to build on the achievements of earlier mathematicians.

      All mathematical systems (for example, Euclidean geometry) are combinations of sets of axioms and of theorems that can be logically deduced from the axioms. Inquiries into the logical and philosophical basis of mathematics reduce to questions of whether the axioms of a given system ensure its completeness and its consistency. For full treatment of this aspect, see mathematics, foundations of.

      This article offers a history of mathematics from ancient times to the present. As a consequence of the exponential growth of science, most mathematics has developed since the 15th century AD, and it is a historical fact that from the 15th century to the late 20th century new developments in mathematics have been largely concentrated in Europe and North America. For these reasons the bulk of this article is devoted to European developments since 1500.

      This does not mean, however, that developments elsewhere have been unimportant. Indeed, to understand the history of mathematics in Europe, it is necessary to know its history at least in Mesopotamia and Egypt, in ancient Greece, and in Islamic civilization from the 9th to the 15th century. The way in which these civilizations influenced one another and the important direct contributions Greece and Islam made to later developments are discussed in the first parts of this article.

      India's contributions to the development of contemporary mathematics were made through the considerable influence of Indian achievements on Islamic mathematics during its formative years. A separate article, South Asian mathematics (mathematics, South Asian), focuses on the early history of mathematics in the Indian subcontinent and the development there of the modern decimal place-value numeral system. The article East Asian mathematics (mathematics, East Asian) covers the mostly independent development of mathematics in China, Japan, Korea, and Vietnam.

      The substantive branches of mathematics are treated in several articles. See algebra; analysis; arithmetic; combinatorics; game theory; geometry; number theory; numerical analysis; optimization; probability theory; set theory; statistics; trigonometry.

Ancient mathematical sources
      It is important to be aware of the character of the sources for the study of the history of mathematics. The history of Mesopotamian and Egyptian mathematics is based on the extant original documents written by scribes. Although in the case of Egypt these documents are few, they are all of a type and leave little doubt that Egyptian mathematics was, on the whole, elementary and profoundly practical in its orientation. For Mesopotamian mathematics, on the other hand, there are a large number of clay tablets, which reveal mathematical achievements of a much higher order than those of the Egyptians. The tablets indicate that the Mesopotamians had a great deal of remarkable mathematical knowledge, although they offer no evidence that this knowledge was organized into a deductive system. Future research may reveal more about the early development of mathematics in Mesopotamia or about its influence on Greek mathematics, but it seems likely that this picture of Mesopotamian mathematics will stand.

      From the period before Alexander the Great, no Greek mathematical documents have been preserved except for fragmentary paraphrases, and, even for the subsequent period, it is well to remember that the oldest copies of Euclid's Elements are in Byzantine manuscripts dating from the 10th century AD. This stands in complete contrast to the situation described above for Egyptian and Babylonian documents. Although in general outline the present account of Greek mathematics is secure, in such important matters as the origin of the axiomatic method, the pre-Euclidean theory of ratios, and the discovery of the conic sections (conic section), historians have given competing accounts based on fragmentary texts, quotations of early writings culled from nonmathematical sources, and a considerable amount of conjecture.

      Many important treatises from the early period of Islamic mathematics have not survived or have survived only in Latin translations, so that there are still many unanswered questions about the relationship between early Islamic mathematics and the mathematics of Greece and India. In addition, the amount of surviving material from later centuries is so large in comparison with that which has been studied that it is not yet possible to offer any sure judgment of what later Islamic mathematics did not contain, and therefore it is not yet possible to evaluate with any assurance what was original in European mathematics from the 11th to the 15th century.

      In modern times the invention of printing has largely solved the problem of obtaining secure texts and has allowed historians of mathematics to concentrate their editorial efforts on the correspondence or the unpublished works of mathematicians. However, the exponential growth of mathematics means that, for the period from the 19th century on, historians are able to treat only the major figures in any detail. In addition, there is, as the period gets nearer the present, the problem of perspective. Mathematics, like any other human activity, has its fashions, and the nearer one is to a given period, the more likely these fashions will look like the wave of the future. For this reason, the present article makes no attempt to assess the most recent developments in the subject.

John L. Berggren

Mathematics in ancient Mesopotamia (Mesopotamia, history of)
      Until the 1920s it was commonly supposed that mathematics had its birth among the ancient Greeks. What was known of earlier traditions, such as the Egyptian as represented by the Rhind papyrus (edited for the first time only in 1877), offered at best a meagre precedent. This impression gave way to a very different view as Orientalists succeeded in deciphering and interpreting the technical materials from ancient Mesopotamia.

      Owing to the durability of the Mesopotamian scribes' clay tablets, the surviving evidence of this culture is substantial. Existing specimens of mathematics represent all the major eras—the Sumerian (Sumer) kingdoms of the 3rd millennium BC, the Akkadian (Akkad) and Babylonian (Babylonia) regimes (2nd millennium), and the empires of the Assyrians (Assyria) (early 1st millennium), Persians (Persia) (6th through 4th centuries BC), and Greeks (3rd century BC to 1st century AD). The level of competence was already high as early as the Old Babylonian dynasty, the time of the lawgiver-king Hammurabi (c. 18th century BC), but after that there were few notable advances. The application of mathematics to astronomy, however, flourished during the Persian and Seleucid (Seleucid kingdom) (Greek) periods.

The numeral system and arithmetic operations
      Unlike the Egyptians, the mathematicians of the Old Babylonian period went far beyond the immediate challenges of their official accounting duties. For example, they introduced a versatile numeral system, which, like the modern system, exploited the notion of place value, and they developed computational methods that took advantage of this means of expressing numbers; they solved linear and quadratic problems by methods much like those now used in algebra; their success with the study of what are now called Pythagorean number triples was a remarkable feat in number theory. The scribes who made such discoveries must have believed mathematics to be worthy of study in its own right, not just as a practical tool.

      The older Sumerian system of numerals followed an additive decimal (base-10) principle similar to that of the Egyptians. But the Old Babylonian system converted this into a place-value system with the base of 60 (sexagesimal). The reasons for the choice of 60 are obscure, but one good mathematical reason might have been the existence of so many divisors (2, 3, 4, and 5, and some multiples) of the base, which would have greatly facilitated the operation of division. For numbers from 1 to 59, the symbols  for 1 and  for 10 were combined in the simple additive manner (e.g.,      represented 32). But, to express larger values, the Babylonians applied the concept of place value: for example, 60 was written as , 70 as  , 80 as   , and so on. In fact,  could represent any power of 60. The context determined which power was intended. The Babylonians appear to have developed a placeholder symbol that functioned as a zero by the 3rd century BC, but its precise meaning and use is still uncertain. Furthermore, they had no mark to separate numbers into integral and fractional parts (as with the modern decimal point). Thus, the three-place numeral 3 7 30 could represent 31/8 (i.e., 3 + 7/60 + 30/602), 1871/2 (i.e., 3 × 60 + 7 + 30/60), 11,250 (i.e., 3 × 602 + 7 × 60 + 30), or a multiple of these numbers by any power of 60.

      The four arithmetic operations were performed in the same way as in the modern decimal system, except that carrying occurred whenever a sum reached 60 rather than 10. Multiplication was facilitated by means of tables; one typical tablet lists the multiples of a number by 1, 2, 3,…, 19, 20, 30, 40, and 50. To multiply two numbers several places long, the scribe first broke the problem down into several multiplications, each by a one-place number, and then looked up the value of each product in the appropriate tables. He found the answer to the problem by adding up these intermediate results. These tables also assisted in division, for the values that head them were all reciprocals of regular numbers.

      Regular numbers are those whose prime factors divide the base; the reciprocals of such numbers thus have only a finite number of places (by contrast, the reciprocals of nonregular numbers produce an infinitely repeating numeral). In base 10, for example, only numbers with factors of 2 and 5 (e.g., 8 or 50) are regular, and the reciprocals (1/8 = 0.125, 1/50 = 0.02) have finite expressions; but the reciprocals of other numbers (such as 3 and 7) repeat infinitely

and
, respectively, where the bar indicates the digits that continually repeat). In base 60, only numbers with factors of 2, 3, and 5 are regular; for example, 6 and 54 are regular, so that their reciprocals (10 and 1 6 40) are finite. The entries in the multiplication table for 1 6 40 are thus simultaneously multiples of its reciprocal 1/54. To divide a number by any regular number, then, one can consult the table of multiples for its reciprocal.

 An interesting tablet in the collection of Yale University (see photograph—>) shows a square with its diagonals; on one side is written “30,” under one diagonal “42 25 35,” and right along the same diagonal “1 24 51 10” (i.e., 1 + 24/60 + 51/602 + 10/603). This third number is the correct value of √2 to four sexagesimal places (equivalent in the decimal system to 1.414213…, which is too low by only 1 in the seventh place), while the second number is the product of the third number and the first and so gives the length of the diagonal when the side is 30. The scribe thus appears to have known an equivalent of the familiar long method of finding square roots. An additional element of sophistication is that, by choosing 30 (that is, 1/2) for the side, the scribe obtained as the diagonal the reciprocal of the value of √2 (since √2/2 = 1/√2), a result useful for purposes of division.

Geometric (geometry) and algebraic problems
      In a Babylonian tablet now in Berlin, the diagonal of a rectangle of sides 40 and 10 is solved as 40 + 102/(2 × 40). Here a very effective approximating rule is being used (that the square root of the sum of a2 + b2 can be estimated as a + b2/2a), the same rule found frequently in later Greek geometric writings. Both these examples for roots illustrate the Babylonians' arithmetic approach in geometry. They also show that the Babylonians were aware of the relation between the hypotenuse and the two legs of a right triangle (now commonly known as the Pythagorean theorem) more than a thousand years before the Greeks used it.

      A type of problem that occurs frequently in the Babylonian tablets seeks the base and height of a rectangle, where their product and sum have specified values. From the given information the scribe worked out the difference, since (b − h)2 = (b + h)2 − 4bh. In the same way, if the product and difference were given, the sum could be found. And, once both the sum and difference were known, each side could be determined, for 2b = (b + h) + (b − h) and 2h = (b + h) − (b − h). This procedure is equivalent to a solution of the general quadratic (quadratic equation) in one unknown. In some places, however, the Babylonian scribes solved quadratic problems in terms of a single unknown, just as would now be done by means of the quadratic formula.

      Although these Babylonian quadratic procedures have often been described as the earliest appearance of algebra, there are important distinctions. The scribes lacked an algebraic symbolism; although they must certainly have understood that their solution procedures were general, they always presented them in terms of particular cases, rather than as the working through of general formulas and identities. They thus lacked the means for presenting general derivations and proofs of their solution procedures. Their use of sequential procedures rather than formulas, however, is less likely to detract from an evaluation of their effort now that algorithmic (algorithm) methods much like theirs have become commonplace through the development of computers.

      As mentioned above, the Babylonian scribes knew that the base (b), height (h), and diagonal (d) of a rectangle satisfy the relation b2 + h2 = d2. If one selects values at random for two of the terms, the third will usually be irrational (irrational number), but it is possible to find cases in which all three terms are integers: for example, 3, 4, 5 and 5, 12, 13. (Such solutions are sometimes called Pythagorean triples.) A tablet in the Columbia University Collection presents a list of 15 such triples (decimal equivalents are shown in parentheses at the right; the gaps in the expressions for h, b, and d separate the place values in the sexagesimal numerals):

      (The entries in the column for h have to be computed from the values for b and d, for they do not appear on the tablet; but they must once have existed on a portion now missing.) The ordering of the lines becomes clear from another column, listing the values of d2/h2 (brackets indicate figures that are lost or illegible), which form a continually decreasing sequence: [1 59 0] 15, [1 56 56] 58 14 50 6 15,…, [1] 23 13 46 40. Accordingly, the angle formed between the diagonal and the base in this sequence increases continually from just over 45° to just under 60°. Other properties of the sequence suggest that the scribe knew the general procedure for finding all such number triples—that for any integers p and q, 2d/h = p/q + q/p and 2b/h = p/q − q/p. (In the table the implied values p and q turn out to be regular numbers falling in the standard set of reciprocals, as mentioned earlier in connection with the multiplication tables.) Scholars are still debating nuances of the construction and the intended use of this table, but no one questions the high level of expertise implied by it.

Mathematical astronomy
      The sexagesimal method developed by the Babylonians has a far greater computational potential than what was actually needed for the older problem texts. With the development of mathematical astronomy in the Seleucid (Seleucid kingdom) period, however, it became indispensable. Astronomers sought to predict future occurrences of important phenomena, such as lunar eclipses and critical points in planetary cycles (conjunctions, oppositions, stationary points, and first and last visibility). They devised a technique for computing these positions (expressed in terms of degrees of latitude and longitude, measured relative to the path of the Sun's apparent annual motion) by successively adding appropriate terms in arithmetic progression. The results were then organized into a table listing positions as far ahead as the scribe chose. (Although the method is purely arithmetic, one can interpret it graphically: the tabulated values form a linear “zigzag” approximation to what is actually a sinusoidal variation.) While observations extending over centuries are required for finding the necessary parameters (e.g., periods, angular range between maximum and minimum values, and the like), only the computational apparatus at their disposal made the astronomers' forecasting effort possible.

      Within a relatively short time (perhaps a century or less), the elements of this system came into the hands of the Greeks (ancient Greek civilization). Although Hipparchus (2nd century BC) favoured the geometric approach of his Greek predecessors, he took over parameters from the Mesopotamians and adopted their sexagesimal style of computation. Through the Greeks it passed to Arab scientists during the Middle Ages and thence to Europe, where it remained prominent in mathematical astronomy during the Renaissance and the early modern period. To this day it persists in the use of minutes and seconds to measure time and angles.

      Aspects of the Old Babylonian mathematics may have come to the Greeks even earlier, perhaps in the 5th century BC, the formative period of Greek geometry. There are a number of parallels that scholars have noted: for example, the Greek technique of “application of area” (see below Greek mathematics (mathematics)) corresponded to the Babylonian quadratic methods (although in a geometric, not arithmetic, form). Further, the Babylonian rule for estimating square roots was widely used in Greek geometric computations, and there may also have been some shared nuances of technical terminology. Although details of the timing and manner of such a transmission are obscure because of the absence of explicit documentation, it seems that Western mathematics, while stemming largely from the Greeks, is considerably indebted to the older Mesopotamians.

Mathematics in ancient Egypt (Egypt, ancient)
      The introduction of writing in Egypt in the predynastic period (c. 3000 BC) brought with it the formation of a special class of literate professionals, the scribes. By virtue of their writing skills, the scribes took on all the duties of a civil service: record keeping, tax accounting, the management of public works (building projects and the like), even the prosecution of war through overseeing military supplies and payrolls. Young men enrolled in scribal schools to learn the essentials of the trade, which included not only reading and writing but also the basics of mathematics.

      One of the texts popular as a copy exercise in the schools of the New Kingdom (Egypt, ancient) (13th century BC) was a satiric letter in which one scribe, Hori, taunts his rival, Amen-em-opet, for his incompetence as an adviser and manager. “You are the clever scribe at the head of the troops,” Hori chides at one point;

a ramp is to be built, 730 cubits long, 55 cubits wide, with 120 compartments—it is 60 cubits high, 30 cubits in the middle…and the generals and the scribes turn to you and say, “You are a clever scribe, your name is famous. Is there anything you don't know? Answer us, how many bricks are needed?” Let each compartment be 30 cubits by 7 cubits.

      This problem, and three others like it in the same letter, cannot be solved without further data. But the point of the humour is clear, as Hori challenges his rival with these hard, but typical, tasks.

      What is known of Egyptian mathematics tallies well with the tests posed by the scribe Hori. The information comes primarily from two long papyrus documents that once served as textbooks within scribal schools. The Rhind papyrus (in the British Museum) is a copy made in the 17th century BC of a text two centuries older still. In it is found a long table of fractional parts to help with division, followed by the solutions of 84 specific problems in arithmetic and geometry. The Golenishchev papyrus (in the Moscow Museum of Fine Arts), dating from the 19th century BC, presents 25 problems of a similar type. These problems reflect well the functions the scribes would perform, for they deal with how to distribute beer and bread as wages, for example, and how to measure the areas of fields as well as the volumes of pyramids and other solids.

The numeral system and arithmetic operations
  The Egyptians, like the Romans after them, expressed numbers according to a decimal scheme, using separate symbols for 1, 10, 100, 1,000, and so on; each symbol appeared in the expression for a number as many times as the value it represented occurred in the number itself. For example, stood for 24. This rather cumbersome notation was used within the hieroglyphic writing (see the figure—>) found in stone inscriptions and other formal texts, but in the papyrus documents the scribes employed a more convenient abbreviated script, called hieratic (hieratic script) writing (see the figure—>), where, for example, 24 was written .

      In such a system, addition and subtraction amount to counting how many symbols of each kind there are in the numerical expressions and then rewriting with the resulting number of symbols. The texts that survive do not reveal what, if any, special procedures the scribes used to assist in this. But for multiplication they introduced a method of successive doubling. For example, to multiply 28 by 11, one constructs a table of multiples of 28 like the following:

      The several entries in the first column that together sum to 11 (i.e., 8, 2, and 1) are checked off. The product is then found by adding up the multiples corresponding to these entries; thus, 224 + 56 + 28 = 308, the desired product.

      To divide 308 by 28, the Egyptians applied the same procedure in reverse. Using the same table as in the multiplication problem, one can see that 8 produces the largest multiple of 28 that is less then 308 (for the entry at 16 is already 448), and 8 is checked off. The process is then repeated, this time for the remainder (84) obtained by subtracting the entry at 8 (224) from the original number (308). This, however, is already smaller than the entry at 4, which consequently is ignored, but it is greater than the entry at 2 (56), which is then checked off. The process is repeated again for the remainder obtained by subtracting 56 from the previous remainder of 84, or 28, which also happens to exactly equal the entry at 1 and which is then checked off. The entries that have been checked off are added up, yielding the quotient: 8 + 2 + 1 = 11. (In most cases, of course, there is a remainder that is less than the divisor.)

      For larger numbers this procedure can be improved by considering multiples of one of the factors by 10, 20,…or even by higher orders of magnitude (100, 1,000,…), as necessary (in the Egyptian decimal notation, these multiples are easy to work out). Thus, one can find the product of 28 by 27 by setting out the multiples of 28 by 1, 2, 4, 8, 10, and 20. Since the entries 1, 2, 4, and 20 add up to 27, one has only to add up the corresponding multiples to find the answer.

      Computations involving fractions are carried out under the restriction to unit parts (that is, fractions that in modern notation are written with 1 as the numerator). To express the result of dividing 4 by 7, for instance, which in modern notation is simply 4/7, the scribe wrote 1/2 + 1/14. The procedure for finding quotients in this form merely extends the usual method for the division of integers, where one now inspects the entries for 2/3, 1/3, 1/6, etc., and 1/2, 1/4, 1/8, etc., until the corresponding multiples of the divisor sum to the dividend. (The scribes included 2/3, one may observe, even though it is not a unit fraction.) In practice the procedure can sometimes become quite complicated (for example, the value for 2/29 is given in the Rhind papyrus as 1/24 + 1/58 + 1/174 + 1/232) and can be worked out in different ways (for example, the same 2/29 might be found as 1/15 + 1/435 or as 1/16 + 1/232 + 1/464, etc.). A considerable portion of the papyrus texts is devoted to tables to facilitate the finding of such unit-fraction values.

      These elementary operations are all that one needs for solving the arithmetic problems in the papyri. For example, “to divide 6 loaves among 10 men” (Rhind papyrus, problem 3), one merely divides to get the answer 1/2 + 1/10. In one group of problems an interesting trick is used: “A quantity (aha) and its 7th together make 19—what is it?” (Rhind papyrus, problem 24). Here one first supposes the quantity to be 7: since 11/7 of it becomes 8, not 19, one takes 19/8 (that is, 2 + 1/4 + 1/8), and its multiple by 7 (16 + 1/2 + 1/8) becomes the required answer. This type of procedure (sometimes called the method of “false position” or “false assumption”) is familiar in many other arithmetic traditions (e.g., the Chinese, Hindu, Muslim, and Renaissance European), although they appear to have no direct link to the Egyptian.

      The geometric problems in the papyri seek measurements of figures, like rectangles and triangles of given base and height, by means of suitable arithmetic operations. In a more complicated problem, a rectangle is sought whose area is 12 and whose height is 1/2 + 1/4 times its base (Golenishchev papyrus, problem 6). To solve the problem, the ratio is inverted and multiplied by the area, yielding 16; the square root of the result (4) is the base of the rectangle, and 1/2 + 1/4 times 4, or 3, is the height. The entire process is analogous to the process of solving the algebraic equation for the problem (x × 3/4x = 12), though without the use of a letter for the unknown. An interesting procedure is used to find the area of the circle (Rhind papyrus, problem 50): 1/9 of the diameter is discarded, and the result is squared. For example, if the diameter is 9, the area is set equal to 64. The scribe recognized that the area of a circle is proportional to the square of the diameter and assumed for the constant of proportionality (that is, π/4) the value 64/81. This is a rather good estimate, being about 0.6 percent too large. (It is not as close, however, as the now common estimate of 31/7, first proposed by Archimedes, which is only about 0.04 percent too large.) But there is nothing in the papyri indicating that the scribes were aware that this rule was only approximate rather than exact.

      A remarkable result is the rule for the volume of the truncated pyramid (Golenishchev papyrus, problem 14). The scribe assumes the height to be 6, the base to be a square of side 4, and the top a square of side 2. He multiplies one-third the height times 28, finding the volume to be 56; here 28 is computed from 2 × 2 + 2 × 4 + 4 × 4. Since this is correct, it can be assumed that the scribe also knew the general rule: A = (h/3)(a2 + ab + b2). How the scribes actually derived the rule is a matter for debate, but it is reasonable to suppose that they were aware of related rules, such as that for the volume of a pyramid: one-third the height times the area of the base.

  The Egyptians employed the equivalent of similar triangles to measure distances. For instance, the seked of a pyramid is stated as the number of palms in the horizontal corresponding to a rise of one cubit (seven palms). (See the figure—>.) Thus, if the seked is 51/4 and the base is 140 cubits, the height becomes 931/3 cubits (Rhind papyrus, problem 57). The Greek sage Thales of Miletus (6th century BC) is said to have measured the height of pyramids by means of their shadows (the report derives from Hieronymus, a disciple of Aristotle in the 4th century BC). (See the figure—>.) In light of the seked computations, however, this report must indicate an aspect of Egyptian surveying that extended back at least 1,000 years before the time of Thales.

Assessment of Egyptian mathematics
      The papyri thus bear witness to a mathematical tradition closely tied to the practical accounting and surveying activities of the scribes. Occasionally, the scribes loosened up a bit: one problem (Rhind papyrus, problem 79), for example, seeks the total from seven houses, seven cats per house, seven mice per cat, seven ears of wheat per mouse, and seven hekat of grain per ear (result: 19,607). Certainly the scribe's interest in progressions (for which he appears to have a rule) goes beyond practical considerations. Other than this, however, Egyptian mathematics falls firmly within the range of practice.

      Even allowing for the scantiness of the documentation that survives, the Egyptian achievement in mathematics must be viewed as modest. Its most striking features are competence and continuity. The scribes managed to work out the basic arithmetic and geometry necessary for their official duties as civil managers, and their methods persisted with little evident change for at least a millennium, perhaps two. Indeed, when Egypt came under Greek domination in the Hellenistic (Hellenistic Age) period (from the 3rd century BC onward), the older school methods continued. Quite remarkably, the older unit-fraction methods are still prominent in Egyptian school papyri written in the demotic (demotic script) (Egyptian) and Greek languages as late as the 7th century AD, for example.

      To the extent that Egyptian mathematics left a legacy at all, it was through its impact on the emerging Greek mathematical tradition between the 6th and 4th centuries BC. Because the documentation from this period is limited, the manner and significance of the influence can only be conjectured. But the report about Thales measuring the height of pyramids is only one of several such accounts of Greek intellectuals learning from Egyptians; Herodotus and Plato describe with approval Egyptian practices in the teaching and application of mathematics. This literary evidence has historical support, since the Greeks maintained continuous trade and military operations in Egypt from the 7th century BC onward. It is thus plausible that basic precedents for the Greeks' earliest mathematical efforts—how they dealt with fractional parts or measured areas and volumes, or their use of ratios in connection with similar figures—came from the learning of the ancient Egyptian scribes.

Greek (ancient Greek civilization) mathematics

The development of pure mathematics
The pre-Euclidean period
 The Greeks divided the field of mathematics into arithmetic (the study of “multitude,” or discrete quantity) and geometry (that of “magnitude,” or continuous quantity) and considered both to have originated in practical activities. Proclus, in his Commentary on Euclid, observes that geometry—literally, “measurement of land”—first arose in surveying practices among the ancient Egyptians, for the flooding of the Nile compelled them each year to redefine the boundaries of properties. Similarly, arithmetic started with the commerce and trade of Phoenician (Phoenicia) merchants. Although Proclus wrote quite late in the ancient period (in the 5th century AD), his account drew upon views proposed much earlier—by Herodotus (mid-5th century BC), for example, and by Eudemus (Eudemus Of Rhodes), a disciple of Aristotle (late 4th century BC).

      However plausible, this view is difficult to check, for there is only meagre evidence of practical mathematics from the early Greek period (roughly, the 8th through the 4th century BC). Inscriptions on stone, for example, reveal use of a numeral system the same in principle as the familiar Roman numerals (Roman numeral). Herodotus seems to have known of the abacus as an aid for computation by both Greeks and Egyptians, and about a dozen stone specimens of Greek abaci survive from the 5th and 4th centuries BC. In the surveying of new cities in the Greek colonies of the 6th and 5th centuries, there was regular use of a standard length of 70 plethra (one plethron equals 100 feet) as the diagonal of a square of side 50 plethra; in fact, the actual diagonal of the square is 50√2 plethra, so this was equivalent to using 7/5 (or 1.4) as an estimate for √2, which is now known to equal 1.414…. In the 6th century BC the engineer Eupalinus of Megara directed an aqueduct through a mountain on the island of Samos, and historians still debate how he did it. In a further indication of the practical aspects of early Greek mathematics, Plato describes in his Laws how the Egyptians drilled their children in practical problems in arithmetic and geometry; he clearly considered this a model for the Greeks to imitate.

      Such hints about the nature of early Greek practical mathematics are confirmed in later sources—for example, in the arithmetic problems in papyrus texts from Ptolemaic Egypt (from the 3rd century BC onward) and the geometric manuals by Heron of Alexandria (1st century AD). In its basic manner this Greek tradition was much like the earlier traditions in Egypt and Mesopotamia. Indeed, it is likely that the Greeks borrowed from such older sources to some extent.

      What was distinctive of the Greeks' contribution to mathematics—and what in effect made them the creators of “mathematics,” as the term is usually understood—was its development as a theoretical discipline. This means two things: mathematical statements are general, and they are confirmed by proof. For example, the Mesopotamians had procedures for finding whole numbers a, b, and c for which a2 + b2 = c2 (e.g., 3, 4, 5; 5, 12, 13; or 119, 120, 169). From the Greeks came a proof of a general rule for finding all such sets of numbers (now called Pythagorean triples): if one takes any whole numbers p and q, both being even or both odd, then a = (p2 − q2)/2, b = pq, and c = (p2 + q2)/2. As Euclid proves in Book X of the Elements, numbers of this form satisfy the relation for Pythagorean triples. Further, the Mesopotamians appear to have understood that sets of such numbers a, b, and c form the sides of right triangles, but the Greeks proved this result (Euclid, in fact, proves it twice: in Elements, Book I, proposition 47, and in a more general form in Elements, Book VI, proposition 31), and these proofs occur in the context of a systematic presentation of the properties of plane geometric figures.

      The Elements, composed by Euclid of Alexandria about 300 BC, was the pivotal contribution to theoretical geometry, but the transition from practical to theoretical mathematics had occurred much earlier, sometime in the 5th century BC. Initiated by men like Pythagoras of Samos (Pythagoras) (late 6th century) and Hippocrates of Chios (late 5th century), the theoretical form of geometry was advanced by others, most prominently the Pythagorean Archytas of Tarentum, Theaetetus of Athens (Theaetetus), and Eudoxus of Cnidus (4th century). Because the actual writings of these men do not survive, knowledge about their work depends on remarks made by later writers. While even this limited evidence reveals how heavily Euclid depended on them, it does not set out clearly the motives behind their studies.

      It is thus a matter of debate how and why this theoretical transition took place. A frequently cited factor is the discovery of irrational numbers (irrational number). The early Pythagoreans (Pythagoreanism) held that “all things are number.” This might be taken to mean that any geometric measure can be associated with some number (that is, some whole number or fraction; in modern terminology, rational number), for in Greek usage the term for number, arithmos, refers exclusively to whole numbers or, in some contexts, to ordinary fractions. This assumption is common enough in practice, as when the length of a given line is said to be so many feet plus a fractional part. However, it breaks down for the lines that form the side and diagonal of the square. (For example, if it is supposed that the ratio between the side and diagonal may be expressed as the ratio of two whole numbers, it can be shown that both of these numbers must be even. This is impossible, since every fraction may be expressed as a ratio of two whole numbers having no common factors.) Geometrically, this means that there is no length that could serve as a unit of measure of both the side and diagonal; that is, the side and diagonal cannot each equal the same length multiplied by (different) whole numbers. Accordingly, the Greeks called such pairs of lengths “incommensurable.” (In modern terminology, unlike that of the Greeks, the term “number” is applied to such quantities as √2, but they are called irrational.)

      This result was already well known at the time of Plato and may well have been discovered within the school of Pythagoras in the 5th century BC, as some late authorities like Pappus of Alexandria (4th century AD) maintain. In any case, by 400 BC it was known that lines corresponding to √3, √5, and other square roots are incommensurable with a fixed unit length. The more general result, the geometric equivalent of the theorem that √p is irrational whenever p is not a rational square number, is associated with Plato's friend Theaetetus. Both Theaetetus and Eudoxus contributed to the further study of irrationals, and their followers collected the results into a substantial theory, as represented by the 115 propositions of Book X of the Elements.

      The discovery of irrationals must have affected the very nature of early mathematical research, for it made clear that arithmetic was insufficient for the purposes of geometry, despite the assumptions made in practical work. Further, once such seemingly obvious assumptions as the commensurability of all lines turned out to be in fact false, then in principle all mathematical assumptions were rendered suspect. At the least it became necessary to justify carefully all claims made about mathematics. Even more basically, it became necessary to establish what a reasoning has to be like to qualify as a proof. Apparently, Hippocrates of Chios, in the 5th century BC, and others soon after him had already begun the work of organizing geometric results into a systematic form in textbooks called “elements” (meaning “fundamental results” of geometry). These were to serve as sources for Euclid in his comprehensive textbook a century later.

      The early mathematicians were not an isolated group but part of a larger, intensely competitive intellectual environment of pre-Socratic thinkers in Ionia and Italy, as well as Sophists (Sophist) at Athens. By insisting that only permanent things could have real existence, the philosopher Parmenides (5th century BC) called into question the most basic claims about knowledge itself. In contrast, Heracleitus (c. 500 BC) maintained that all permanence is an illusion, for the things that are perceived arise through a subtle balance of opposing tensions. What is meant by “knowledge” and “proof” thus came into debate.

      Mathematical issues were often drawn into these debates. For some, like the Pythagoreans (and, later, Plato), the certainty of mathematics was held as a model for reasoning in other areas, like politics and ethics. But for others mathematics seemed prone to contradiction. Zeno Of Elea (5th century BC) posed paradoxes (paradoxes of Zeno) about quantity and motion. In one such paradox it is assumed that a line can be bisected again and again without limit; if the division ultimately results in a set of points of zero length, then even infinitely many of them sum up only to zero, but, if it results in tiny line segments, then their sum will be infinite. In effect, the length of the given line must be both zero and infinite. In the 5th century BC a solution of such paradoxes was attempted by Democritus and the atomists (atomism), philosophers who held that all material bodies are ultimately made up of invisibly small “atoms” (the Greek word atomon means “indivisible”). But in geometry such a view came into conflict with the existence of incommensurable lines, since the atoms would become the measuring units of all lines, even incommensurable ones. Democritus and the Sophist Protagoras puzzled over whether the tangent to a circle meets it at a point or a line. The Sophists Antiphon and Bryson (both 5th century BC) considered how to compare the circle to polygons inscribed in it.

      The pre-Socratics thus revealed difficulties in specific assumptions about the infinitely many and the infinitely small and about the relation of geometry to physical reality, as well as in more general conceptions like “existence” and “proof.” Philosophical questions such as these need not have affected the technical researches of mathematicians, but they did make them aware of difficulties that could bear on fundamental matters and so made them the more cautious in defining their subject matter.

      Any such review of the possible effects of factors such as these is purely conjectural, since the sources are fragmentary and never make explicit how the mathematicians responded to the issues that were raised. But it is the particular concern over fundamental assumptions and proofs that distinguishes Greek mathematics from the earlier traditions. Plausible factors behind this concern can be identified in the special circumstances of the early Greek tradition—its technical discoveries and its cultural environment—even if it is not possible to describe in detail how these changes took place.

The Elements
      The principal source for reconstructing pre-Euclidean mathematics is Euclid's Elements, for the major part of its contents can be traced back to research from the 4th century BC and in some cases even earlier. The first four books present constructions and proofs of plane geometric figures: Book I deals with the congruence of triangles, the properties of parallel lines, and the area relations of triangles and parallelograms; Book II establishes equalities relating to squares, rectangles, and triangles; Book III covers basic properties of circles; and Book IV sets out constructions of polygons in circles. Much of the content of Books I–III was already familiar to Hippocrates, and the material of Book IV can be associated with the Pythagoreans, so that this portion of the Elements has roots in 5th-century research. It is known, however, that questions about parallels were debated in Aristotle's school (c. 350 BC), and so it may be assumed that efforts to prove results—such as the theorem stating that, for any given line and given point, there always exists a unique line through that point and parallel to the line—were tried and failed. Thus, the decision to found the theory of parallels on a postulate, as in Book I of the Elements, must have been a relatively recent development in Euclid's time. (The postulate would later become the subject of much study, and in modern times it led to the discovery of the so-called non-Euclidean geometries (non-Euclidean geometry).)

      Book V sets out a general theory of proportion—that is, a theory that does not require any restriction to commensurable magnitudes. This general theory derives from Eudoxus. On the basis of the theory, Book VI describes the properties of similar plane rectilinear figures and so generalizes the congruence theory of Book I. It appears that the technique of similar figures was already known in the 5th century BC, even though a fully valid justification could not have been given before Eudoxus worked out his theory of proportion.

      Books VII–IX deal with what the Greeks called “ arithmetic,” the theory of whole numbers. It includes the properties of numerical proportions, greatest common divisors, least common multiples, and relative primes (Book VII); propositions on numerical progressions and square and cube numbers (Book VIII); and special results, like unique factorization into primes, the existence of an unlimited number of primes, and the formation of “perfect numbers (perfect number)”—that is, those numbers that equal the sum of their proper divisors (Book IX). In some form Book VII stems from Theaetetus and Book VIII from Archytas.

      Book X presents a theory of irrational lines and derives from the work of Theaetetus and Eudoxus. The remaining books treat the geometry of solids. Book XI sets out results on solid figures analogous to those for planes in Books I and VI; Book XII proves theorems on the ratios of circles, the ratios of spheres, and the volumes of pyramids and cones; Book XIII shows how to inscribe the five regular solids—known as the Platonic solids (Platonic solid)—in a given sphere (compare the constructions of plane figures in Book IV). The measurement of curved figures in Book XII is inferred from that of rectilinear figures; for a particular curved figure, a sequence of rectilinear figures is considered in which succeeding figures in the sequence become continually closer to the curved figure; the particular method used by Euclid derives from Eudoxus. The solid constructions in Book XIII derive from Theaetetus.

      In sum the Elements gathered together the whole field of elementary geometry and arithmetic that had developed in the two centuries before Euclid. Doubtless, Euclid must be credited with particular aspects of this work, certainly with its editing as a comprehensive whole. But it is not possible to identify for certain even a single one of its results as having been his discovery. Other, more advanced fields, though not touched on in the Elements, were already being vigorously studied in Euclid's time, in some cases by Euclid himself. For these fields his textbook, true to its name, provides the appropriate “elementary” introduction.

      One such field is the study of geometric constructions. Euclid, like geometers in the generation before him, divided mathematical propositions into two kinds: “theorems” and “problems.” A theorem makes the claim that all terms of a certain description have a specified property; a problem seeks the construction of a term that is to have a specified property. In the Elements all the problems are constructible on the basis of three stated postulates: that a line can be constructed by joining two given points, that a given line segment can be extended in a line indefinitely, and that a circle can be constructed with a given point as centre and a given line segment as radius. These postulates in effect restricted the constructions to the use of the so-called Euclidean tools—i.e., a compass and a straightedge or unmarked ruler.

The three classical problems
      Although Euclid solves more than 100 construction problems in the Elements, many more were posed whose solutions required more than just compass and straightedge. Three such problems stimulated so much interest among later geometers that they have come to be known as the “classical problems”: doubling the cube (i.e., constructing a cube whose volume is twice that of a given cube), trisecting the angle, and squaring the circle. Even in the pre-Euclidean period the effort to construct a square equal in area to a given circle had begun. Some related results came from Hippocrates (see Sidebar: Quadrature of the Lune); others were reported from Antiphon and Bryson; and Euclid's theorem on the circle in Elements, Book XII, proposition 2, which states that circles are in the ratio of the squares of their diameters, was important for this search. But the first actual constructions (not, it must be noted, by means of the Euclidean tools, for this is impossible) came only in the 3rd century BC. The early history of angle trisection is obscure. Presumably, it was attempted in the pre-Euclidean period, although solutions are known only from the 3rd century or later.

 There are several successful efforts at doubling the cube that date from the pre-Euclidean period, however. Hippocrates (Hippocrates of Chios) showed that the problem could be reduced to that of finding two mean proportionals: if for a given line a it is necessary to find x such that x3 = 2a3, lines x and y may be sought such that a:x = x:y = y:2a; for then a3/x3 = (a/x)3 = (a/x)(x/y)(y/2a) = a/2a = 1/2. (Note that the same argument holds for any multiplier, not just the number 2.) Thus, the cube can be doubled if it is possible to find the two mean proportionals x and y between the two given lines a and 2a. Constructions of the problem of the two means were proposed by Archytas, Eudoxus, and Menaechmus in the 4th century BC. Menaechmus, for example, constructed three curves corresponding to these same proportions: x2 = ay, y2 = 2ax, and xy = 2a2; the intersection of any two of them then produces the line x that solves the problem (see figure—>). Menaechmus's curves are conic sections: the first two are parabolas, the third a hyperbola. Thus, it is often claimed that Menaechmus originated the study of the conic sections (conic section). Indeed, Proclus and his older authority, Geminus (mid-1st century AD), appear to have held this view. The evidence does not indicate how Menaechmus actually conceived of the curves, however, so it is possible that the formal study of the conic sections as such did not begin until later, near the time of Euclid. Both Euclid and an older contemporary, Aristaeus, composed treatments (now lost) of the theory of conic sections.

      In seeking the solutions of problems, geometers developed a special technique, which they called “analysis.” They assumed the problem to have been solved and then, by investigating the properties of this solution, worked back to find an equivalent problem that could be solved on the basis of the givens. To obtain the formally correct solution of the original problem, then, geometers reversed the procedure: first the data were used to solve the equivalent problem derived in the analysis, and, from the solution obtained, the original problem was then solved. In contrast to analysis, this reversed procedure is called “synthesis.”

      Menaechmus's cube duplication is an example of analysis: he assumed the mean proportionals x and y and then discovered them to be equivalent to the result of intersecting the three curves whose construction he could take as known. (The synthesis consists of introducing the curves, finding their intersection, and showing that this solves the problem.) It is clear that geometers of the 4th century BC were well acquainted with this method, but Euclid provides only syntheses, never analyses, of the problems solved in the Elements. Certainly in the cases of the more complicated constructions, however, there can be little doubt that some form of analysis preceded the syntheses presented in the Elements.

Geometry in the 3rd century BC
      The Elements was one of several major efforts by Euclid and others to consolidate the advances made over the 4th century BC. On the basis of these advances, Greek geometry entered its golden age in the 3rd century. This was a period rich with geometric discoveries, particularly in the solution of problems by analysis and other methods, and was dominated by the achievements of two figures: Archimedes of Syracuse (Archimedes) (early 3rd century BC) and Apollonius of Perga (late 3rd century BC).

 Archimedes was most noted for his use of the Eudoxean method of exhaustion (exhaustion, method of) in the measurement of curved surfaces and volumes and for his applications of geometry to mechanics. To him is owed the first appearance and proof of the approximation 31/7 for the ratio of the circumference to the diameter of the circle (what is now designated π). Characteristically, Archimedes went beyond familiar notions, such as that of simple approximation, to more subtle insights, like the notion of bounds. For example, he showed that the perimeters of regular polygons circumscribed about the circle eventually become less than 31/7 the diameter as the number of their sides increases (Archimedes established the result for 96-sided polygons); similarly, the perimeters of the inscribed polygons eventually become greater than 310/71. Thus, these two values are upper and lower bounds, respectively, of π. (See the animation—>.)

 Archimedes' result bears on the problem of circle quadrature in the light of another theorem he proved: that the area of a circle equals the area of a triangle whose height equals the radius of the circle and whose base equals its circumference. He established analogous results for the sphere showing that the volume of a sphere is equal to that of a cone whose height equals the radius of the sphere and whose base equals its surface area; the surface area of the sphere he found to be four times the area of its greatest circle. Equivalently, the volume of a sphere is shown to be two-thirds that of the cylinder which just contains it (that is, having height and diameter equal to the diameter of the sphere), while its surface is also equal to two-thirds that of the same cylinder (that is, if the circles that enclose the cylinder at top and bottom are included). The Greek historian Plutarch (early 2nd century AD) relates that Archimedes requested the figure for this theorem (see the figure—>) to be engraved on his tombstone, which is confirmed by the Roman writer Cicero (Cicero, Marcus Tullius) (1st century BC), who actually located the tomb in 75 BC, when he was quaestor of Sicily.

      The work of Apollonius of Perga extended the field of geometric constructions far beyond the range in the Elements. For example, Euclid in Book III shows how to draw a circle so as to pass through three given points or to be tangent to three given lines; Apollonius (in a work called Tangencies, which no longer survives) found the circle tangent to three given circles, or tangent to any combination of three points, lines, and circles. (The three-circle tangency construction, one of the most extensively studied geometric problems, has attracted more than 100 different solutions in the modern period.)

 Apollonius is best known for his Conics, a treatise in eight books (Books I–IV survive in Greek, V–VII in a medieval Arabic translation; Book VIII is lost). The conic sections (conic section) are the curves formed when a plane intersects the surface of a cone (or double cone), as shown in the figure—>; it is assumed that the surface of the cone is generated by the rotation of a line through a fixed point around the circumference of a circle which is in a plane not containing that point. (The fixed point is the vertex of the cone, and the rotated line its generator.) There are three basic types: if the cutting plane is parallel to one of the positions of the generator, it produces a parabola; if it meets the cone only on one side of the vertex, it produces an ellipse (of which the circle is a special case); but, if it meets both parts of the cone, it produces a hyperbola. Apollonius sets out in detail the properties of these curves. He shows, for example, that for given line segments a and b the parabola corresponds to the relation (in modern notation) y2 = ax, the ellipse to y2 = ax − ax2/b, and the hyperbola to y2 = ax + ax2/b.

      Apollonius's treatise on conics in part consolidated more than a century of work before him and in part presented new findings of his own. As mentioned earlier, Euclid had already issued a textbook on the conics, while even earlier Menaechmus had played a role in their study. The names that Apollonius chose for the curves (the terms may be original with him) indicate yet an earlier connection. In the pre-Euclidean geometry parabolē referred to a specific operation, the “application” of a given area to a given line, in which the line x is sought such that ax = b2 (where a and b are given lines); alternatively, x may be sought such that x(a + x) = b2, or x(a − x) = b2, and in these cases the application is said to be in “excess” (hyperbolē) or “defect” (elleipsis) by the amount of a square figure (namely, x2). These constructions, which amount to a geometric solution of the general quadratic, appear in Books I, II, and VI of the Elements and can be associated in some form with the 5th-century Pythagoreans.

      Apollonius presented a comprehensive survey of the properties of these curves. A sample of the topics he covered includes the following: the relations satisfied by the diameters and tangents of conics (Book I); how hyperbolas are related to their “asymptotes (asymptote),” the lines they approach without ever meeting (Book II); how to draw tangents to given conics (Book II); relations of chords intersecting in conics (Book III); the determination of the number of ways in which conics may intersect (Book IV); how to draw “normal” lines to conics (that is, lines meeting them at right angles; Book V); and the congruence and similarity of conics (Book VI).

      By Apollonius's explicit statement, his results are of principal use as methods for the solution of geometric problems via conics. While he actually solved only a limited set of problems, the solutions of many others can be inferred from his theorems. For instance, the theorems of Book III permit the determination of conics that pass through given points or are tangent to given lines. In another work (now lost) Apollonius solved the problem of cube duplication by conics (a solution related in some way to that given by Menaechmus); further, a solution of the problem of angle trisection given by Pappus may have come from Apollonius or been influenced by his work.

      With the advance of the field of geometric problems by Euclid, Apollonius, and their followers, it became appropriate to introduce a classifying scheme: those problems solvable by means of conics were called solid, while those solvable by means of circles and lines only (as assumed in Euclid's Elements) were called planar. Thus, one can double the square by planar means (as in Elements, Book II, proposition 14), but one cannot double the cube in such a way, although a solid construction is possible (as given above). Similarly, the bisection of any angle is a planar construction (as shown in Elements, Book I, proposition 9), but the general trisection of the angle is of the solid type. It is not known when the classification was first introduced or when the planar methods were assigned canonical status relative to the others, but it seems plausible to date this near Apollonius's time. Indeed, much of his work—books like the Tangencies, the Vergings (or Inclinations), and the Plane Loci, now lost but amply described by Pappus—turns on the project of setting out the domain of planar constructions in relation to solutions by other means. On the basis of the principles of Greek geometry, it cannot be demonstrated, however, that it is impossible to effect by planar means certain solid constructions (like the cube duplication and angle trisection). These results were established only by algebraists in the 19th century (notably by the French mathematician Pierre Laurent Wantzel in 1837).

   A third class of problems, called linear, embraced those solvable by means of curves other than the circle and the conics (in Greek the word for “line,” grammē, refers to all lines, whether curved or straight). For instance, one group of curves, the conchoids (from the Greek word for “shell”), are formed by marking off a certain length on a ruler and then pivoting it about a fixed point in such a way that one of the marked points stays on a given line; the other marked point traces out a conchoid (see the figure—>). These curves can be used wherever a solution involves the positioning of a marked ruler relative to a given line (in Greek such constructions are called neuses, or “vergings” of a line to a given point). For example, any acute angle (figured as the angle between one side and the diagonal of a rectangle) can be trisected by taking a length equal to twice the diagonal and moving it about until it comes to be inserted between two other sides of the rectangle. If instead the appropriate conchoid relative to either of those sides is introduced, the required position of the line can be determined without the trial and error of a moving ruler (see figure—>). Because the same construction can be effected by means of a hyperbola (see figure—>), however, the problem is not linear but solid. Such uses of the conchoids were presented by Nicomedes (middle or late 3rd century BC), and their replacement by equivalent solid constructions appears to have come soon after, perhaps by Apollonius or his associates.

      Some of the curves used for problem solving are not so reducible. For example, the Archimedean spiral couples uniform motion of a point on a half ray with uniform rotation of the ray around a fixed point at its end (see Sidebar: Quadratrix of Hippias). Such curves have their principal interest as means for squaring the circle and trisecting the angle.

Applied geometry
      A major activity among geometers in the 3rd century BC was the development of geometric approaches in the study of the physical sciences—specifically, optics, mechanics, and astronomy. In each case the aim was to formulate the basic concepts and principles in terms of geometric and numerical quantities and then to derive the fundamental phenomena of the field by geometric constructions and proofs.

      In optics, Euclid's (Euclid) textbook (called the Optics) set the precedent. Euclid postulated visual rays to be straight lines, and he defined the apparent size of an object in terms of the angle formed by the rays drawn from the top and the bottom of the object to the observer's eye. He then proved, for example, that nearer objects appear larger and appear to move faster and showed how to measure the height of distant objects from their shadows or reflected images and so on. Other textbooks set out theorems on the phenomena of reflection and refraction (the field called catoptrics). The most extensive survey of optical phenomena is a treatise attributed to the astronomer Ptolemy (2nd century AD), which survives only in the form of an incomplete Latin translation (12th century) based on a lost Arabic translation. It covers the fields of geometric optics and catoptrics, as well as experimental areas, such as binocular vision, and more general philosophical principles (the nature of light, vision, and colour). Of a somewhat different sort are the studies of burning mirrors by Diocles (late 2nd century BC), who proved that the surface that reflects the rays from the Sun to a single point is a paraboloid of revolution. Constructions of such devices remained of interest as late as the 6th century AD, when Anthemius of Tralles (Anthemius), best known for his work as architect of Hagia Sophia at Constantinople, compiled a survey of remarkable mirror configurations.

       mechanics was dominated by the work of Archimedes, who was the first to prove the principle of balance: that two weights are in equilibrium when they are inversely proportional to their distances from the fulcrum. From this principle he developed a theory of the centres of gravity of plane and solid figures. He was also the first to state and prove the principle of buoyancy (Archimedes' principle)—that floating bodies displace their equal in weight—and to use it for proving the conditions of stability of segments of spheres and paraboloids, solids formed by rotating a parabolic segment about its axis. Archimedes proved the conditions under which these solids will return to their initial position if tipped, in particular for the positions now called “stable I” and “stable II,” where the vertex faces up and down, respectively.

   In his work Method Concerning Mechanical Theorems, Archimedes also set out a special “mechanical method” that he used for the discovery of results on volumes and centres of gravity. He employed the bold notion of constituting solids from the plane figures formed as their sections (e.g., the circles that are the plane sections of spheres, cones, cylinders, and other solids of revolution), assigning to such figures a weight proportional to their area. For example, to measure the volume of a sphere, he imagined a balance beam, one of whose arms is a diameter of the sphere with the fulcrum at one endpoint of this diameter and the other arm an extension of the diameter to the other side of the fulcrum by a length equal to the diameter. Archimedes showed that the three circular cross sections made by a plane cutting the sphere and the associated cone and cylinder will be in balance (the circle in the cylinder with the circles in the sphere and cone) if the circle in the cylinder is kept in its original place while the circles in the sphere and cone are placed with their centres of gravity at the opposite end of the balance. Doing this for all the sets of circles formed as cross sections of these solids by planes, he concluded that the solids themselves are in balance—the cylinder with the sphere and the cone together—if the cylinder is left where it is, while the sphere and cone are placed with their centres of gravity at the opposite end of the balance. Since the centre of gravity of the cylinder is the midpoint of its axis, it follows that (sphere + cone):cylinder = 1:2 (by the inverse proportion of weights and distances). Since the volume of the cone is one-third that of the cylinder, however, the volume of the sphere is found to be one-sixth that of the cylinder. In similar manner, Archimedes worked out the volumes and centres of gravity of spherical segments and segments of the solids of revolution of conic sections—paraboloids (see the figures of an elliptic paraboloid—> and a hyperbolic paraboloid—>), ellipsoids (see the figure—>), and hyperboloids. The critical notions—constituting solids out of their plane sections and assigning weights to geometric figures—were not formally valid within the standard conceptions of Greek geometry, and Archimedes admitted this. But he maintained that, although his arguments were not “demonstrations” (i.e., proofs), they had value for the discovery of results about these figures.

 The geometric study of astronomy has pre-Euclidean roots, Eudoxus having developed a model for planetary motions around a stationary Earth. Accepting the principle—which, according to Eudemus, was first proposed by Plato—that only combinations of uniform circular motions are to be used, Eudoxus (Eudoxus of Cnidus) represented the path of a planet as the result of superimposing rotations of three or more concentric spheres whose axes are set at different angles. Although the fit with the phenomena was unsatisfactory, the curves thus generated (the hippopede, or “horse-fetter”) continued to be of interest for their geometric properties, as is known through remarks by Proclus. Later geometers continued the search for geometric patterns satisfying the Platonic conditions. The simplest model, a scheme of circular (heliocentric system) orbits (orbit) centred on the Sun, was introduced by Aristarchus Of Samos (3rd century BC), but this was rejected by others, since a moving Earth was judged to be impossible on physical grounds. But Aristarchus's scheme could have suggested use of an “eccentric” model, in which the planets rotate about the Sun and the Sun in turn rotates about the Earth. Apollonius introduced an alternative “epicyclic” model, in which the planet turns about a point that itself orbits in a circle (the “deferent”) centred at or near the Earth. As Apollonius knew, his epicyclic model is geometrically equivalent to an eccentric. These models were well adapted for explaining other phenomena of planetary motion. For instance, if the Earth is displaced from the centre of a circular orbit (as in the eccentric scheme), the orbiting body will appear to vary in speed (appearing faster when nearer the observer, slower when farther away), as is in fact observed for the Sun, Moon, and planets. By varying the relative sizes and rotation rates of the epicycle and deferent, in combination with the eccentric, a flexible device may be obtained for representing planetary motion. (See the figure—> of Ptolemy's model.)

Later trends in geometry and arithmetic
Greek trigonometry and mensuration
      After the 3rd century BC, mathematical research shifted increasingly away from the pure forms of constructive geometry toward areas related to the applied disciplines, in particular to astronomy. The necessary theorems on the geometry of the sphere (called spherics) were compiled into textbooks, such as the one by Theodosius (3rd or 2nd century BC) that consolidated the earlier work by Euclid and the work of Autolycus of Pitane (flourished c. 300 BC) on spherical astronomy. More significant, in the 2nd century BC the Greeks first came into contact with the fully developed Mesopotamian astronomical systems and took from them many of their observations and parameters (for example, values for the average periods of astronomical phenomena). While retaining their own commitment to geometric models rather than adopting the arithmetic schemes of the Mesopotamians, the Greeks nevertheless followed the Mesopotamians' lead in seeking a predictive astronomy based on a combination of mathematical theory and observational parameters. They thus made it their goal not merely to describe but to calculate the angular positions of the planets on the basis of the numerical and geometric content of the theory. This major restructuring of Greek astronomy, in both its theoretical and practical respects, was primarily due to Hipparchus (2nd century BC), whose work was consolidated and further advanced by Ptolemy.

      To facilitate their astronomical researches, the Greeks developed techniques for the numerical measurement of angles, a precursor of trigonometry, and produced tables suitable for practical computation. Early efforts to measure the numerical ratios in triangles were made by Archimedes and Aristarchus. Their results were soon extended, and comprehensive treatises on the measurement of chords (in effect, a construction of a table of values equivalent to the trigonometric sine) were produced by Hipparchus and by Menelaus of Alexandria (1st century AD). These works are now lost, but the essential theorems and tables are preserved in Ptolemy's Almagest (Book I, chapter 10). For computing with angles, the Greeks adopted the Mesopotamian sexagesimal method in arithmetic, whence it survives in the standard units for angles and time employed to this day.

 Although Euclid handed down a precedent for number theory in Books VII–IX of the Elements, later writers made no further effort to extend the field of theoretical arithmetic in his demonstrative manner. Beginning with Nicomachus of Gerasa (flourished c. AD 100), several writers produced collections expounding a much simpler form of number theory. A favourite result is the representation of arithmetic progressions in the form of “polygonal numbers.” For instance, if the numbers 1, 2, 3, 4,…are added successively, the “triangular” numbers 1, 3, 6, 10,…are obtained; similarly, the odd numbers 1, 3, 5, 7,…sum to the “square” numbers 1, 4, 9, 16,…, while the sequence 1, 4, 7, 10,…, with a constant difference of 3, sums to the “pentagonal” numbers 1, 5, 12, 22,…. In general, these results can be expressed in the form of geometric shapes formed by lining up dots in the appropriate two-dimensional configurations (see the figure—>). In the ancient arithmetics such results are invariably presented as particular cases, without any general notational method or general proof. The writers in this tradition are called neo-Pythagoreans, since they viewed themselves as continuing the Pythagorean school of the 5th century BC, and, in the spirit of ancient Pythagoreanism, they tied their numerical interests to a philosophical theory that was an amalgam of Platonic metaphysical and theological doctrines. With its exponent Iamblichus of Chalcis (Iamblichus) (4th century AD), neo-Pythagoreans became a prominent part of the revival of pagan religion in opposition to Christianity in late antiquity.

      An interesting concept of this school of thought, which Iamblichus attributes to Pythagoras himself, is that of “amicable numbers”: two numbers are amicable if each is equal to the sum of the proper divisors of the other (for example, 220 and 284). Attributing virtues such as friendship and justice to numbers was characteristic of the Pythagoreans at all times.

      Of much greater mathematical significance is the arithmetic work of Diophantus of Alexandria (c. 3rd century AD). His writing, the Arithmetica, originally in 13 books (six survive in Greek, another four in medieval Arabic translation), sets out hundreds of arithmetic problems with their solutions. For example, Book II, problem 8, seeks to express a given square number as the sum of two square numbers (here and throughout, the “numbers” are rational). Like those of the neo-Pythagoreans, his treatments are always of particular cases rather than general solutions; thus, in this problem the given number is taken to be 16, and the solutions worked out are 256/25 and 144/25. In this example, as is often the case, the solutions are not unique; indeed, in the very next problem Diophantus shows how a number given as the sum of two squares (e.g., 13 = 4 + 9) can be expressed differently as the sum of two other squares (for example, 13 = 324/25 + 1/25).

      To find his solutions, Diophantus adopted an arithmetic form of the method of analysis. He first reformulated the problem in terms of one of the unknowns, and he then manipulated it as if it were known until an explicit value for the unknown emerged. He even adopted an abbreviated notational scheme to facilitate such operations, where, for example, the unknown is symbolized by a figure somewhat resembling the Roman letter S. (This is a standard abbreviation for the word number in ancient Greek manuscripts.) Thus, in the first problem discussed above, if S is one of the unknown solutions, then 16 − S2 is a square; supposing the other unknown to be 2S − 4 (where the 2 is arbitrary but the 4 chosen because it is the square root of the given number 16), Diophantus found from summing the two unknowns ([2S − 4]2 and S2) that 4S2 − 16S + 16 + S2 = 16, or 5S2 = 16S; that is, S = 16/5. So one solution is S2 = 256/25, while the other solution is 16 − S2, or 144/25.

Survival and influence of Greek mathematics
      Notable in the closing phase of Greek mathematics were Pappus (Pappus of Alexandria) (early 4th century AD), Theon (late 4th century), and Theon's daughter Hypatia. All were active in Alexandria as professors of mathematics and astronomy, and they produced extensive commentaries on the major authorities—Pappus and Theon on Ptolemy, Hypatia on Diophantus and Apollonius. Later, Eutocius of Ascalon (early 6th century) produced commentaries on Archimedes and Apollonius. While much of their output has since been lost, much survives. They proved themselves reasonably competent in technical matters but little inclined toward significant insights (their aim was usually to fill in minor steps assumed in the proofs, to append alternative proofs, and the like), and their level of originality was very low. But these scholars frequently preserved fragments of older works that are now lost, and their teaching and editorial efforts assured the survival of the works of Euclid, Archimedes, Apollonius, Diophantus, Ptolemy, and others that now do exist, either in Greek manuscripts or in medieval translations (Arabic, Hebrew, and Latin) derived from them.

      The legacy of Greek mathematics, particularly in the fields of geometry and geometric science, was enormous. From an early period the Greeks formulated the objectives of mathematics not in terms of practical procedures but as a theoretical discipline committed to the development of general propositions and formal demonstrations. The range and diversity of their findings, especially those of the masters of the 3rd century BC, supplied geometers with subject matter for centuries thereafter, even though the tradition that was transmitted into the Middle Ages and Renaissance was incomplete and defective.

      The rapid rise of mathematics in the 17th century was based in part on the conscious imitation of the ancient classics and on competition with them. In the geometric mechanics of Galileo and the infinitesimal researches of Johannes Kepler (Kepler, Johannes) and Bonaventura Cavalieri (Cavalieri, Bonaventura), it is possible to perceive a direct inspiration from Archimedes. The study of the advanced geometry of Apollonius and Pappus stimulated new approaches in geometry—for example, the analytic methods of René Descartes (Descartes, René) and the projective theory of Girard Desargues (Desargues, Girard). Purists like Christiaan Huygens (Huygens, Christiaan) and Isaac Newton (Newton, Sir Isaac) insisted on the Greek geometric style as a model of rigour, just as others sought to escape its forbidding demands of completely worked-out proofs. The full impact of Diophantus's work is evident particularly with Pierre de Fermat (Fermat, Pierre de) in his researches in algebra and number theory. Although mathematics has today gone far beyond the ancient achievements, the leading figures of antiquity, like Archimedes, Apollonius, and Ptolemy, can still be rewarding reading for the ingenuity of their insights.

Wilbur R. Knorr

Mathematics in the Islamic world (Islāmic world) (8th–15th century)

Origins
 In Hellenistic times and in late antiquity, scientific learning in the eastern part of the Roman world was spread over a variety of centres, and Justinian (Justinian I)'s closing of the pagan academies in Athens in 529 gave further impetus to this diffusion. An additional factor was the translation and study of Greek scientific and philosophical texts sponsored both by monastic centres of the various Christian churches in the Levant, Egypt, and Mesopotamia and by enlightened rulers of the Sāsānian dynasty in places like the medical school at Gondeshapur.

      Also important were developments in India in the first few centuries AD. Although the decimal system (decimal number system) for whole numbers was apparently not known to the Indian astronomer Aryabhata I (born 476), it was used by his pupil Bhaskara I in 620, and by 670 the system had reached northern Mesopotamia, where the Nestorian bishop Severus Sebokht praised its Hindu inventors as discoverers of things more ingenious than those of the Greeks. Earlier, in the late 4th or early 5th century, the anonymous Hindu author of an astronomical handbook, the Surya Siddhanta, had tabulated the sine function (unknown in Greece) for every 33/4° of arc from 33/4° to 90°. (See South Asian mathematics (mathematics, South Asian).)

      Within this intellectual context the rapid expansion of Islam took place between the time of Muḥammad's return to Mecca in 630 from his exile in Medina and the Muslim conquest of lands extending from Spain to the borders of China by 715. Not long afterward, Muslims began the acquisition of foreign learning, and, by the time of the caliph al-Manṣūr (Manṣūr, al-) (died 775), such Indian and Persian astronomical material as the Brahma-sphuta-siddhanta and the Shah's Tables had been translated into Arabic. The subsequent acquisition of Greek material was greatly advanced when the caliph al-Maʾmūn (Maʾmūn, al-) constructed a translation and research centre, the House of Wisdom, in Baghdad during his reign (813–833). Most of the translations were done from Greek (ancient Greek civilization) and Syriac by Christian scholars, but the impetus and support for this activity came from Muslim patrons. These included not only the caliph but also wealthy individuals such as the three brothers known as the Banū Mūsā, whose treatises on geometry and mechanics formed an important part of the works studied in the Islamic world.

      Of Euclid's works the Elements, the Data, the Optics, the Phaenomena, and On Divisions were translated. Of Archimedes' works only two—Sphere and Cylinder and Measurement of the Circle—are known to have been translated, but these were sufficient to stimulate independent researches from the 9th to the 15th century. On the other hand, virtually all of Apollonius's works were translated, and of Diophantus and Menelaus one book each, the Arithmetica and the Sphaerica, respectively, were translated into Arabic. Finally, the translation of Ptolemy's Almagest furnished important astronomical material.

      Of the minor writings, Diocles' treatise on mirrors, Theodosius's Spherics, Pappus's work on mechanics, Ptolemy's Planisphaerium, and Hypsicles' treatises on regular polyhedra (the so-called Books XIV and XV of Euclid's Elements) were among those translated.

Mathematics in the 9th century
      Thābit ibn Qurrah (Thābit ibn Qurra) (836–901), a Sabian from Ḥarrān in northern Mesopotamia, was an important translator and reviser of these Greek works. In addition to translating works of the major Greek mathematicians (for the Banū Mūsā, among others), he was a court physician. He also translated Nicomachus of Gerasa's Arithmetic and discovered a beautiful rule for finding amicable numbers, a pair of numbers such that each number is the sum of the set of proper divisors of the other number. The investigation of such numbers formed a continuing tradition in Islam. Kamāl al-Dīn al-Fārisī (died c. 1320) gave the pair 17,926 and 18,416 as an example of Thābit's rule, and in the 17th century Muḥammad Bāqir Yazdī gave the pair 9,363,584 and 9,437,056.

      One scientist typical of the 9th century was Muḥammad ibn Mūsā al-Khwārizmī (Khwārizmī, al-). Working in the House of Wisdom, he introduced Indian material in his astronomical works and also wrote an early book explaining Hindu arithmetic, the Book of Addition and Subtraction According to the Hindu Calculation. In another work, the Book of Restoring and Balancing, he provided a systematic introduction to algebra, including a theory of quadratic equations. Both works had important consequences for Islamic mathematics. Hindu Calculation began a tradition of arithmetic books that, by the middle of the next century, led to the invention of decimal fractions (complete with a decimal point), and Restoring and Balancing became the point of departure and model for later writers such as the Egyptian Abū Kāmil. Both books were translated into Latin, and Restoring and Balancing was the origin of the word algebra, from the Arabic word for “restoring” in its title (al-jabr). The Hindu Calculation, from a Latin form of the author's name, algorismi, yielded the word algorithm.

      Al-Khwārizmī's algebra also served as a model for later writers in its application of arithmetic and algebra to the distribution of inheritances (inheritance) according to the complex requirements of Muslim religious law. This tradition of service to the Islamic faith was an enduring feature of mathematical work in Islam and one that, in the eyes of many, justified the study of secular learning. In the same category are al-Khwārizmī's method of calculating the time of visibility of the new moon (which signals the beginning of the Muslim month) and the expositions by astronomers of methods for finding the direction (qiblah) to Mecca for the five daily prayers.

Mathematics in the 10th century
      Islamic scientists in the 10th century were involved in three major mathematical projects: the completion of arithmetic algorithms, the development of algebra, and the extension of geometry.

      The first of these projects led to the appearance of three complete numeration systems, one of which was the finger arithmetic used by the scribes and treasury officials. This ancient arithmetic system, which became known throughout the East and Europe, employed mental arithmetic and a system of storing intermediate results on the fingers as an aid to memory. (Its use of unit fractions recalls the Egyptian system.) During the 10th and 11th centuries capable mathematicians, such as Abūʾl-Wafāʾ (940–997/998), wrote on this system, but it was eventually replaced by the decimal system.

      A second common system was the base-60 numeration inherited from the Babylonians via the Greeks and known as the arithmetic of the astronomers. Although astronomers used this system for their tables, they usually converted numbers to the decimal system for complicated calculations and then converted the answer back to sexagesimals.

      The third system was Indian arithmetic, whose basic numeral forms, complete with the zero, eastern Islam took over from the Hindus. (Different forms of the numerals, whose origins are not entirely clear, were used in western Islam.) The basic algorithms also came from India, but these were adapted by al-Uqlīdisī (c. 950) to pen and paper instead of the traditional dust board, a move that helped to popularize this system. Also, the arithmetic algorithms were completed in two ways: by the extension of root-extraction procedures, known to Hindus and Greeks only for square and cube roots, to roots of higher degree and by the extension of the Hindu decimal system for whole numbers to include decimal fractions. These fractions appear simply as computational devices in the work of both al-Uqlīdisī and al-Baghdādī (c. 1000), but in subsequent centuries they received systematic treatment as a general method. As for extraction of roots, Abūʾl-Wafāʾ wrote a treatise (now lost) on the topic, and Omar Khayyam (1048–1131) solved the general problem of extracting roots of any desired degree. Omar's treatise too is lost, but the method is known from other writers, and it appears that a major step in its development was al-Karajī's (Karajī, al-) 10th-century derivation by means of mathematical induction of the binomial theorem for whole-number exponents—i.e., his discovery that

      During the 10th century Islamic algebraists progressed from al-Khwārizmī's quadratic polynomials to the mastery of the algebra of expressions involving arbitrary positive or negative integral powers of the unknown. Several algebraists explicitly stressed the analogy between the rules for working with powers of the unknown in algebra and those for working with powers of 10 in arithmetic, and there was interaction between the development of arithmetic and algebra from the 10th to the 12th century. A 12th-century student of al-Karajī's works, al-Samawʿal, was able to approximate the quotient (20x2 + 30x)/(6x2 + 12) as

      and also gave a rule for finding the coefficients of the successive powers of 1/x. Although none of this employed symbolic (symbol) algebra, algebraic symbolism was in use by the 14th century in the western part of the Islamic world. The context for this well-developed symbolism was, it seems, commentaries that were destined for teaching purposes, such as that of Ibn Qunfūdh (1330–1407) of Algeria on the algebra of Ibn al-Bannāʿ (1256–1321) of Morocco.

      Other parts of algebra developed as well. Both Greeks and Hindus had studied indeterminate equations, and the translation of this material and the application of the newly developed algebra led to the investigation of Diophantine equations (Diophantine equation) by writers like Abū Kāmil, al-Karajī, and Abū Jaʿfar al-Khāzin (first half of 10th century), as well as to attempts to prove a special case of what is now known as Fermat's last theorem—namely, that there are no rational solutions to x3 + y3 = z3. The great scientist Ibn al-Haytham (965–1040) solved problems involving congruences by what is now called Wilson's theorem, which states that, if p is a prime, then p divides (p − 1) × (p − 2)⋯× 2 × 1 + 1, and al-Baghdādī gave a variant of the idea of amicable numbers by defining two numbers to “balance” if the sums of their divisors are equal.

      However, not only arithmetic and algebra but geometry too underwent extensive development. Thābit ibn Qurrah, his grandson Ibrāhīm ibn Sinān (909–946), Abū Sahl al-Kūhī (died c. 995), and Ibn al-Haytham solved problems involving the pure geometry of conic sections (conic section), including the areas and volumes of plane and solid figures formed from them, and also investigated the optical properties of mirrors made from conic sections. Ibrāhīm ibn Sinān, Abu Sahl al-Kūhī, and Ibn al-Haytham used the ancient technique of analysis to reduce the solution of problems to constructions involving conic sections. (Ibn al-Haytham, for example, used this method to find the point on a convex spherical mirror at which a given object is seen by a given observer.) Thābit and Ibrāhīm showed how to design the curves needed for sundials. Abūʾl-Wafāʾ, whose book on the arithmetic of the scribes is mentioned above, also wrote on geometric methods needed by artisans.

      In addition, in the late 10th century Abūʾl-Wafāʾ and the prince Abū Naṣr Manṣur stated and proved theorems of plane and spherical geometry that could be applied by astronomers and geographers, including the laws of sines and tangents. Abū Naṣr's pupil al-Bīrūnī (Bīrūnī, al-) (973–1048), who produced a vast amount of high-quality work, was one of the masters in applying these theorems to astronomy and to such problems in mathematical geography as the determination of latitudes and longitudes, the distances between cities, and the direction from one city to another.

Omar Khayyam
      The mathematician and poet Omar Khayyam was born in Neyshābūr (in Iran) only a few years before al-Bīrūnī's death. He later lived in Samarkand and Eṣfahān, and his brilliant work there continued many of the main lines of development in 10th-century mathematics. Not only did he discover a general method of extracting roots of arbitrary high degree, but his Algebra contains the first complete treatment of the solution of cubic equations. Omar did this by means of conic sections, but he declared his hope that his successors would succeed where he had failed in finding an algebraic formula for the roots.

 Omar was also a part of an Islamic tradition, which included Thābit and Ibn al-Haytham, of investigating Euclid's parallel postulate. To this tradition Omar contributed the idea of a quadrilateral with two congruent sides perpendicular to the base, as shown in the figure—>. The parallel postulate would be proved, Omar recognized, if he could show that the remaining two angles were right angles. In this he failed, but his question about the quadrilateral became the standard way of discussing the parallel postulate.

      That postulate, however, was only one of the questions on the foundations of mathematics that interested Islamic scientists. Another was the definition of ratios. Omar Khayyam, along with others before him, felt that the theory in Book V of Euclid's Elements was logically satisfactory but intuitively unappealing, so he proved that a definition known to Aristotle was equivalent to that given in Euclid. In fact, Omar argued that ratios should be regarded as “ideal numbers,” and so he conceived of a much broader system of numbers than that used since Greek antiquity, that of the positive real numbers.

Islamic mathematics to the 15th century
      In the 12th century the physician al-Samawʿal continued and completed the work of al-Karajī in algebra and also provided a systematic treatment of decimal fractions as a means of approximating irrational quantities. In his method of finding roots of pure equations, xn = N, he used what is now known as Horner's method to expand the binomial (a + y)n. His contemporary Sharaf al-Dīn al-Ṭūsī late in the 12th century provided a method of approximating the positive roots of arbitrary equations, based on an approach virtually identical to that discovered by François Viète (Viète, François, Seigneur De La Bigotiere) in 16th-century France. The important step here was less the general idea than the development of the numerical algorithms necessary to effect it.

      Sharaf al-Dīn was the discoverer of a device, called the linear astrolabe, that places him in another important Islamic mathematical tradition, one that centred on the design of new forms of the ancient astronomical instrument known as the astrolabe. The astrolabe, whose mathematical theory is based on the stereographic projection of the sphere, was invented in late antiquity, but its extensive development in Islam made it the pocket watch of the medievals. In its original form it required a different plate of horizon coordinates for each latitude, but in the 11th century the Spanish Muslim astronomer al-Zarqallu invented a single plate that worked for all latitudes. Slightly earlier, astronomers in the East had experimented with plane projections of the sphere, and al-Bīrūnī invented such a projection that could be used to produce a map of a hemisphere. The culminating masterpiece was the astrolabe of the Syrian Ibn al-Shāṭir (1305–75), a mathematical tool that could be used to solve all the standard problems of spherical astronomy in five different ways.

      On the other hand, Muslim astronomers had developed other methods for solving these problems using the highly accurate trigonometry tables (trigonometry table) and the new trigonometry theorems they had developed. Out of these developments came the creation of trigonometry as a mathematical discipline, separate from its astronomical applications, by Naṣīr al-Dīn al-Ṭūsī (Ṭūsī, Naṣīr al-Dīn al-) at his observatory in Marāgheh in the 13th century. (It was there too that al-Ṭūsī's pupil Quṭb al-Dīn al-Shīrāzī [1236–1311] and his pupil Kamāl al-Dīn Fārisī, using Ibn al-Haytham's great work, the Optics, were able to give the first mathematically satisfactory explanation of the rainbow.)

      Al-Ṭūsī's observatory was supported by a grandson of Genghis Khan, Hülegü, who sacked Baghdad in 1258. Ulūgh Beg, the grandson of the Mongol conqueror Timur, founded an observatory at Samarkand in the early years of the 15th century. Ulūgh Beg was himself a good astronomer, and his tables of sines and tangents for every minute of arc (accurate to five sexagesimal places) were one of the great achievements in numerical mathematics up to his time. He was also the patron of Jamshīd al-Kāshī (Kāshī, al-) (died 1429), whose work The Reckoners' Key summarizes most of the arithmetic of his time and includes sections on algebra and practical geometry as well. Among al-Kāshī's works is a masterful computation of the value of 2π, which, when expressed in decimal fractions, is accurate to 16 places, as well as the application of a numerical method, now known as fixed-point iteration, for solving the cubic equation with sin 1° as a root. His work was indeed of a quality deserving Ulūgh Beg's description as “known among the famous of the world.”

      Al-Kāshī lived almost five centuries after the first translations of Arabic material into Latin, and by his time the Islamic mathematical tradition had given the West not only its first versions of many of the Greek classics but also a complete set of algorithms for Hindu-Arabic arithmetic, plane and spherical trigonometry, and the powerful tool of algebra. Although mathematical inquiry continued in Islam in the centuries after al-Kāshī's time, the mathematical centre of gravity was shifting to the West. That this was so is, of course, in no small measure due to what the Western mathematicians had learned from their Islamic predecessors during the preceding centuries.

John L. Berggren

European (Europe) mathematics during the Middle Ages and Renaissance
      Until the 11th century only a small part of the Greek mathematical corpus was known in the West. Because almost no one could read Greek, what little was available came from the poor texts written in Latin in the Roman Empire, together with the very few Latin translations of Greek works. Of these the most important were the treatises by Boethius (Boethius, Anicius Manlius Severinus), who about AD 500 made Latin redactions of a number of Greek scientific and logical writings. His Arithmetic, which was based on Nicomachus (Nicomachus of Gerasa), was well known and was the means by which medieval scholars learned of Pythagorean number theory. Boethius and Cassiodorus provided the material for the part of the monastic education called the quadrivium: arithmetic, geometry, astronomy, and music theory. Together with the trivium (grammar, logic, rhetoric), these subjects formed the seven liberal arts, which were taught in the monasteries, cathedral schools, and, from the 12th century on, universities and which constituted the principal university instruction until modern times.

      For monastic life it sufficed to know how to calculate with Roman numerals. The principal application of arithmetic was a method for determining the date of Easter, the computus, that was based on the lunar cycle of 19 solar years (i.e., 235 lunar revolutions) and the 28-year solar cycle. Between the time of Bede (Bede the Venerable, Saint) (died 735), when the system was fully developed, and about 1500, the computus was reduced to a series of verses that were learned by rote. Until the 12th century, geometry was largely concerned with approximate formulas for measuring areas and volumes in the tradition of the Roman surveyors. About AD 1000 the French scholar Gerbert of Aurillac, later Pope Sylvester II, introduced a type of abacus in which numbers were represented by stones bearing Arabic numerals. Such novelties were known to very few.

The transmission of Greek and Arabic learning
      In the 11th century a new phase of mathematics began with the translations from Arabic. Scholars throughout Europe went to Toledo, Córdoba, and elsewhere in Spain to translate into Latin the accumulated learning of the Muslims. Along with philosophy, astronomy, astrology, and medicine, important mathematical achievements of the Greek, Indian, and Islamic civilizations became available in the West. Particularly important were Euclid's Elements, the works of Archimedes, and al-Khwārizmī (Khwārizmī, al-)'s treatises on arithmetic and algebra. Western texts called algorismus (a Latin form of the name al-Khwārizmī) introduced the Hindu-Arabic numerals and applied them in calculations. Thus, modern numerals first came into use in universities and then became common among merchants and other laymen. It should be noted that, up to the 15th century, calculations were often performed with board and counters. Reckoning with Hindu-Arabic numerals was used by merchants at least from the time of Leonardo of Pisa (Leonardo Pisano) (beginning of the 13th century), first in Italy and then in the trading cities of southern Germany and France, where maestri d'abbaco or Rechenmeister taught commercial arithmetic in the various vernaculars. Some schools were private, while others were run by the community.

The universities
      Mathematics was studied from a theoretical standpoint in the universities. The Universities of Paris (Paris I–XIII, Universities of) and Oxford (Oxford, University of), which were founded relatively early (c. 1200), were centres for mathematics and philosophy. Of particular importance in these universities were the Arabic-based versions of Euclid, of which there were at least four by the 12th century. Of the numerous redactions and compendia which were made, that of Johannes Campanus (c. 1250; first printed in 1482) was easily the most popular, serving as a textbook for many generations. Such redactions of the Elements were made to help students not only to understand Euclid's textbook but also to handle other, particularly philosophical, questions suggested by passages in Aristotle. The ratio theory of the Elements provided a means of expressing the various relations of the quantities associated with moving bodies, relations that now would be expressed by formulas. Also in Euclid were to be found methods of analyzing infinity and continuity (paradoxically, because Euclid always avoided infinity).

      Studies of such questions led not only to new results but also to a new approach to what is now called physics. Thomas Bradwardine (Bradwardine, Thomas), who was active in Merton College, Oxford, in the first half of the 14th century, was one of the first medieval scholars to ask whether the continuum can be divided infinitely or whether there are smallest parts (indivisibles). Among other topics, he compared different geometric shapes in terms of the multitude of points that were assumed to compose them, and from such an approach paradoxes were generated that were not to be solved for centuries. Another fertile question stemming from Euclid concerned the angle between a circle and a line tangent to it (called the horn angle): if this angle is not zero, a contradiction quickly ensues, but, if it is zero, then, by definition, there can be no angle. For the relation of force, resistance, and the speed of the body moved by this force, Bradwardine suggested an exponential law. Nicholas Oresme (Oresme, Nicholas) (died 1382) extended Bradwardine's ideas to fractional exponents.

 Another question having to do with the quantification of qualities, the so-called latitude of forms, began to be discussed at about this time in Paris and in Merton College. Various Aristotelian qualities (e.g., heat, density, and velocity) were assigned an intensity and extension, which were sometimes represented by the height and bases (respectively) of a geometric figure. The area of the figure was then considered to represent the quantity of the quality. In the important case in which the quality is the motion of a body, the intensity its speed, and the extension its time, the area of the figure was taken to represent the distance covered by the body. Uniformly accelerated motion starting at zero velocity gives rise to a triangular figure (see the figure—>). It was proved by the Merton school that the quantity of motion in such a case is equal to the quantity of a uniform motion at the speed achieved halfway through the accelerated motion; in modern formulation, s = 1/2at2 (Merton rule). Discussions like this certainly influenced Galileo indirectly and may have influenced the founding of coordinate geometry in the 17th century. Another important development in the scholastic “calculations” was the summation of infinite series.

      Basing his work on translated Greek sources, about 1464 the German mathematician and astronomer Regiomontanus wrote the first book (printed in 1533) in the West on plane and spherical trigonometry independent of astronomy. He also published tables of sines and tangents that were in constant use for more than two centuries.

      Italian (Italy) artists and merchants influenced the mathematics of the late Middle Ages and the Renaissance in several ways. In the 15th century a group of Tuscan artists, including Filippo Brunelleschi (Brunelleschi, Filippo), Leon Battista Alberti (Alberti, Leon Battista), and Leonardo da Vinci, incorporated linear perspective into their practice and teaching, about a century before the subject was formally treated by mathematicians. Italian maestri d'abbaco tried, albeit unsuccessfully, to solve nontrivial cubic equations. In fact, the first general solution was found by Scipione del Ferro (Ferro, Scipione) at the beginning of the 16th century and rediscovered by Niccolò Tartaglia (Tartaglia, Niccolò Fontana) several years later. The solution was published by Gerolamo Cardano (Cardano, Girolamo) in his Ars magna (Ars Magna or the Rules of Algebra) in 1545, together with Lodovico Ferrari (Ferrari, Lodovico)'s solution of the quartic equation.

      By 1380 an algebraic symbolism had been developed in Italy in which letters were used for the unknown, for its square, and for constants. The symbols used today for the unknown (for example, x), the square root sign, and the signs + and − came into general use in southern Germany beginning about 1450. They were used by Regiomontanus and by Fridericus Gerhart and received an impetus about 1486 at the University of Leipzig from Johann Widman. The idea of distinguishing between known and unknown quantities in algebra was first consistently applied by François Viète (Viète, François, Seigneur De La Bigotiere), with vowels for unknown and consonants for known quantities. Viète found some relations between the coefficients of an equation and its roots. This was suggestive of the idea, explicitly stated by Albert Girard in 1629 and proved by Carl Friedrich Gauss (Gauss, Carl Friedrich) in 1799, that an equation of degree n has n roots. Complex numbers (complex number), which are implicit in such ideas, were gradually accepted about the time of Rafael Bombelli (died 1572), who used them in connection with the cubic.

      Apollonius (Apollonius of Perga)'s Conics and the investigations of areas (quadratures) and of volumes (cubatures) by Archimedes formed part of the humanistic learning of the 16th century. These studies strongly influenced the later developments of analytic geometry, the infinitesimal calculus, and the theory of functions (function), subjects that were developed in the 17th century.

Menso Folkerts

Mathematics in the 17th and 18th centuries

The 17th century
      The 17th century, the period of the scientific revolution, witnessed the consolidation of Copernican heliocentric astronomy and the establishment of inertial physics in the work of Johannes Kepler (Kepler, Johannes), Galileo, René Descartes (Descartes, René), and Isaac Newton (Newton, Sir Isaac). This period was also one of intense activity and innovation in mathematics. Advances in numerical calculation, the development of symbolic algebra and analytic geometry, and the invention of the differential and integral calculus resulted in a major expansion of the subject areas of mathematics. By the end of the 17th century, a program of research based in analysis had replaced classical Greek geometry at the centre of advanced mathematics. In the next century this program would continue to develop in close association with physics, more particularly mechanics and theoretical astronomy. The extensive use of analytic methods, the incorporation of applied subjects, and the adoption of a pragmatic attitude to questions of logical rigour distinguished the new mathematics from traditional geometry.

Institutional background
      Until the middle of the 17th century, mathematicians worked alone or in small groups, publishing their work in books or communicating with other researchers by letter. At a time when people were often slow to publish, “invisible colleges,” networks of scientists who corresponded privately, played an important role in coordinating and stimulating mathematical research. Marin Mersenne (Mersenne, Marin) in Paris acted as a clearinghouse for new results, informing his many correspondents—including Pierre de Fermat (Fermat, Pierre de), Descartes, Blaise Pascal (Pascal, Blaise), Gilles Personne de Roberval (Roberval, Gilles Personne de), and Galileo—of challenge problems and novel solutions. Later in the century John Collins, librarian of London's Royal Society, performed a similar function among British mathematicians.

      In 1660 the Royal Society of London was founded, to be followed in 1666 by the French Academy of Sciences (Sciences, Academy of), in 1700 by the Berlin Academy, and in 1724 by the St. Petersburg Academy. The official publications sponsored by the academies, as well as independent journals such as the Acta Eruditorum (founded in 1682), made possible the open and prompt communication of research findings. Although universities in the 17th century provided some support for mathematics, they became increasingly ineffective as state-supported academies assumed direction of advanced research.

Numerical calculation
      The development of new methods of numerical calculation was a response to the increased practical demands of numerical computation, particularly in trigonometry, navigation, and astronomy. New ideas spread quickly across Europe and resulted by 1630 in a major revolution in numerical practice.

      Simon Stevin (Stevin, Simon) of Holland, in his short pamphlet La Disme (1585), introduced decimal fractions to Europe and showed how to extend the principles of Hindu-Arabic arithmetic to calculation with these numbers. Stevin emphasized the utility of decimal arithmetic “for all accounts that are encountered in the affairs of men,” and he explained in an appendix how it could be applied to surveying, stereometry, astronomy, and mensuration. His idea was to extend the base-10 positional principle to numbers with fractional parts, with a corresponding extension of notation to cover these cases. In his system the number 237.578 was denoted

      in which the digits to the left of the zero are the integral part of the number. To the right of the zero are the digits of the fractional part, with each digit succeeded by a circled number that indicates the negative power to which 10 is raised. Stevin showed how the usual arithmetic of whole numbers could be extended to decimal fractions, using rules that determined the positioning of the negative powers of 10.

      In addition to its practical utility, La Disme was significant for the way it undermined the dominant style of classical Greek geometry in theoretical mathematics. Stevin's proposal required a rejection of the distinction in Euclidean geometry between magnitude, which is continuous, and number, which is a multitude of indivisible units. For Euclid, unity, or one, was a special sort of thing, not number but the origin, or principle, of number. The introduction of decimal fractions seemed to imply that the unit could be subdivided and that arbitrary continuous magnitude could be represented numerically; it implicitly supposed the concept of a general positive real number.

      Tables of logarithms (logarithm) were first published in 1614 by the Scottish laird John Napier (Napier, John) in his treatise Description of the Marvelous Canon of Logarithms. This work was followed (posthumously) five years later by another in which Napier set forth the principles used in the construction of his tables. The basic idea behind logarithms is that addition and subtraction are easier to perform than multiplication and division, which, as Napier observed, require a “tedious expenditure of time” and are subject to “slippery errors.” By the law of exponents, anam = an + m; that is, in the multiplication of numbers, the exponents are related additively. By correlating the geometric sequence of numbers a, a2, a3,…(a is called the base) and the arithmetic sequence 1, 2, 3,…and interpolating to fractional values, it is possible to reduce the problem of multiplication and division to one of addition and subtraction. To do this Napier chose a base that was very close to 1, differing from it by only 1/107. The resulting geometric sequence therefore yielded a dense set of values, suitable for constructing a table.

      In his work of 1619 Napier presented an interesting kinematic (kinematics) model to generate the geometric and arithmetic sequences used in the construction of his tables. Assume two particles move along separate lines from given initial points. The particles begin moving at the same instant with the same velocity. The first particle continues to move with a speed that is decreasing, proportional at each instant to the distance remaining between it and some given fixed point on the line. The second particle moves with a constant speed equal to its initial velocity. Given any increment of time, the distances traveled by the first particle in successive increments form a geometrically decreasing sequence. The corresponding distances traveled by the second particle form an arithmetically increasing sequence. Napier was able to use this model to derive theorems yielding precise limits to approximate values in the two sequences.

      Napier's kinematic model indicated how skilled mathematicians had become by the early 17th century in analyzing nonuniform motion. Kinematic ideas, which appeared frequently in mathematics of the period, provided a clear and visualizable means for the generation of geometric magnitude. The conception of a curve traced by a particle moving through space later played a significant role in the development of the calculus.

      Napier's ideas were taken up and revised by the English mathematician Henry Briggs (Briggs, Henry), the first Savilian Professor of Geometry at Oxford. In 1624 Briggs published an extensive table of common logarithms, or logarithms to the base 10. Because the base was no longer close to 1, the table could not be obtained as simply as Napier's, and Briggs therefore devised techniques involving the calculus of finite differences to facilitate calculation of the entries. He also devised interpolation procedures of great computational efficiency to obtain intermediate values.

      In Switzerland the instrument maker Joost Bürgi (Bürgi, Joost) arrived at the idea for logarithms independently of Napier, although he did not publish his results until 1620. Four years later a table of logarithms prepared by Kepler (Kepler, Johannes) appeared in Marburg. Both Bürgi and Kepler were astronomical observers, and Kepler included logarithmic tables in his famous Tabulae Rudolphinae (1627; “ Rudolphine Tables”), astronomical tabulations of planetary motion derived by using the assumption of elliptical orbits about the Sun.

      The invention of analytic geometry was, next to the differential and integral calculus, the most important mathematical development of the 17th century. Originating in the work of the French mathematicians Viète, Fermat, and Descartes, it had by the middle of the century established itself as a major program of mathematical research.

      Two tendencies in contemporary mathematics stimulated the rise of analytic geometry. The first was an increased interest in curves, resulting in part from the recovery and Latin translation of the classical treatises of Apollonius (Apollonius of Perga), Archimedes, and Pappus (Pappus of Alexandria), and in part from the increasing importance of curves in such applied fields as astronomy, mechanics, optics, and stereometry. The second was the emergence a century earlier of an established algebraic practice in the work of the Italian and German algebraists and its subsequent shaping by Viète (Viète, François, Seigneur De La Bigotiere) into a powerful mathematical tool at the end of the century.

      Viète was a prominent representative of the humanist movement in mathematics that set itself the project of restoring and furthering the achievements of the Classical Greek geometers. In his In artem analyticem isagoge (1591; “Introduction to the Analytic Arts”), Viète, as part of his program of rediscovering the method of analysis used by the ancient Greek mathematicians, proposed new algebraic methods that employed variables, constants, and equations, but he saw this as an advancement over the ancient method, a view he arrived at by comparing the geometric analysis contained in Book VII of Pappus's Collection with the arithmetic analysis of Diophantus (Diophantus of Alexandria)'s Arithmetica. Pappus had employed an analytic method for the discovery of theorems and the construction of problems; in analysis, by contrast to synthesis, one proceeds from what is sought until one arrives at something known. In approaching an arithmetic problem by laying down an equation among known and unknown magnitudes and then solving for the unknown, one was, Viète reasoned, following an “analytic” procedure.

      Viète introduced the concept of algebraic variable, which he denoted using a capital vowel (A, E, I, O, U), as well as the concept of parameter (an unspecified constant quantity), denoted by a capital consonant (B, C, D, and so on). In his system the equation 5BA2 − 2CA + A3 = D would appear as B5 in A quad − C plano 2 in A + A cub aequatur D solido.

      Viète retained the classical principle of homogeneity, according to which terms added together must all be of the same dimension. In the above equation, for example, each of the terms has the dimension of a solid or cube; thus, the constant C, which denotes a plane, is combined with A to form a quantity having the dimension of a solid.

      It should be noted that in Viète's scheme the symbol A is part of the expression for the object obtained by operating on the magnitude denoted by A. Thus, operations on the quantities denoted by the variables are reflected in the algebraic notation itself. This innovation, considered by historians of mathematics to be a major conceptual advance in algebra, facilitated the study of the symbolic solution of algebraic equations and led to the creation of the first conscious theory of equations.

      After Viète's death the analytic art was applied to the study of curves by his countrymen Fermat and Descartes. Both men were motivated by the same goal, to apply the new algebraic techniques to Apollonius's theory of loci as preserved in Pappus's Collection. The most celebrated of these problems consisted of finding the curve or locus traced by a point whose distances from several fixed lines satisfied a given relation.

      Fermat adopted Viète's notation in his paper Ad Locos Planos et Solidos Isagoge (1636; “Introduction to Plane and Solid Loci”). The title of the paper refers to the ancient classification of curves as plane (straight lines, circles), solid (ellipses, parabolas, and hyperbolas), or linear (curves defined kinematically or by a locus condition). Fermat considered an equation among two variables. One of the variables represented a line measured horizontally from a given initial point, while the other represented a second line positioned at the end of the first line and inclined at a fixed angle to the horizontal. As the first variable varied in magnitude, the second took on a value determined by the equation, and the endpoint of the second line traced out a curve in space. By means of this construction Fermat was able to formulate the fundamental principle of analytic geometry:

Whenever two unknown quantities are found in final equality, there results a locus fixed in place, and the endpoint of one of these unknown quantities describes a straight line or a curve.

      The principle implied a correspondence between two different classes of mathematical objects: geometric curves and algebraic equations. In the paper of 1636 Fermat showed that, if the equation is a quadratic, then the curve is a conic section—that is, an ellipse, parabola, or hyperbola. He also showed that the determination of the curve given by an equation is simplified by a transformation involving a change of variables to an equation in standard form.

      Descartes's La Géométrie appeared in 1637 as an appendix to his famous Discourse on Method, the treatise that presented the foundation of his philosophical system. Although supposedly an example from mathematics of his rational method, La Géométrie was a technical treatise understandable independently of philosophy. It was destined to become one of the most influential books in the history of mathematics.

      In the opening sections of La Géométrie, Descartes introduced two innovations. In place of Viète's notation he initiated the modern practice of denoting variables by letters at the end of the alphabet (x, y, z) and parameters by letters at the beginning of the alphabet (a, b, c) and of using exponential notation to indicate powers of x (x2, x3,…). More significant conceptually, he set aside Viète's principle of homogeneity, showing by means of a simple construction how to represent multiplication and division of lines by lines; thus, all magnitudes (lines, areas, and volumes) could be represented independently of their dimension in the same way.

      Descartes's goal in La Géométrie was to achieve the construction of solutions to geometric problems by means of instruments that were acceptable generalizations of ruler and compass. Algebra was a tool to be used in this program:

If, then, we wish to solve any problem, we first suppose the solution already effected, and give names to all the lines that seem necessary for its construction—to those that are unknown as well as to those that are known. Then, making no distinction in any way between known and unknown lines, we must unravel the difficulty in any way that shows most naturally the relations between these lines, until we find it possible to express a single quantity in two ways. This will constitute an equation, since the terms of one of these two expressions are together equal to the terms of the other.

      In the problem of Apollonius, for example, one sought to find the locus of points whose distances from a collection of fixed lines satisfied a given relation. One used this relation to derive an equation, and then, using a geometric procedure involving acceptable instruments of construction, one obtained points on the curve given by the roots of the equation.

      Descartes described instruments more general than the compass for drawing “geometric” curves. He stipulated that the parts of the instrument be linked together so that the ratio of the motions of the parts could be knowable. This restriction excluded “mechanical” curves generated by kinematic processes. The Archimedean spiral, for example, was generated by a point moving on a line as the line rotated uniformly about the origin. The ratio of the circumference to the diameter did not permit exact determination:

the ratios between straight and curved lines are not known, and I even believe cannot be discovered by men, and therefore no conclusion based upon such ratios can be accepted as rigorous and exact.

      Descartes concluded that a geometric or nonmechanical curve was one whose equation f(xy) = 0 was a polynomial of finite degree in two variables. He wished to restrict mathematics to the consideration of such curves.

      Descartes's emphasis on construction reflected his classical orientation. His conservatism with respect to what curves were acceptable in mathematics further distinguished him as a traditional thinker. At the time of his death, in 1650, he had been overtaken by events, as research moved away from questions of construction to problems of finding areas (then called problems of quadrature) and tangents. The geometric objects that were then of growing interest were precisely the mechanical curves that Descartes had wished to banish from mathematics.

      Following the important results achieved in the 16th century by Gerolamo Cardano (Cardano, Girolamo) and the Italian algebraists, the theory of algebraic equations (algebraic equation) reached an impasse. The ideas needed to investigate equations of degree higher than four were slow to develop. The immediate historical influence of Viète, Fermat, and Descartes was to furnish algebraic methods for the investigation of curves. A vigorous school of research became established in Leiden around Frans van Schooten, a Dutch mathematician who edited and published in 1649 a Latin translation of La Géométrie. Van Schooten published a second two-volume translation of the same work in 1659–1661 that also contained mathematical appendixes by three of his disciples, Johan de Witt (De Witt, Johan), Johan Hudde (Hudde, Johan van Waveren), and Hendrick van Heuraet. The Leiden group of mathematicians, which also included Christiaan Huygens (Huygens, Christiaan), was in large part responsible for the rapid development of Cartesian geometry in the middle of the century.

      The historian Carl Boyer called the calculus “the most effective instrument for scientific investigation that mathematics has ever produced.” As the mathematics of variability and change, the calculus was the characteristic product of the scientific revolution. The subject was properly the invention of two mathematicians, the German Gottfried Wilhelm Leibniz (Leibniz, Gottfried Wilhelm) and the Englishman Isaac Newton (Newton, Sir Isaac). Both men published their researches in the 1680s, Leibniz in 1684 in the recently founded journal Acta Eruditorum and Newton in 1687 in his great treatise, the Principia. Although a bitter dispute over priority developed later between followers of the two men, it is now clear that they each arrived at the calculus independently.

      The calculus developed from techniques to solve two types of problems, the determination of areas and volumes and the calculation of tangents to curves. In classical geometry Archimedes had advanced farthest in this part of mathematics, having used the method of exhaustion (exhaustion, method of) to establish rigorously various results on areas and volumes and having derived for some curves (e.g., the spiral) significant results concerning tangents. In the early 17th century there was a sharp revival of interest in both classes of problems. The decades between 1610 and 1670, referred to in the history of mathematics as “the precalculus period,” were a time of remarkable activity in which researchers throughout Europe contributed novel solutions and competed with each other to arrive at important new methods.

The precalculus period
      In his treatise Geometria Indivisibilibus Continuorum (1635; “Geometry of Continuous Indivisibles”), Bonaventura Cavalieri (Cavalieri, Bonaventura), a professor of mathematics at the University of Bologna, formulated a systematic method for the determination of areas and volumes. As had Archimedes, Cavalieri regarded a plane figure as being composed of a collection of indivisible lines, “all the lines” of the plane figure. The collection was generated by a fixed line moving through space parallel to itself. Cavalieri showed that these collections could be interpreted as magnitudes obeying the rules of Euclidean ratio theory. In proposition 4 of Book II, he derived the result that is written today as

Let there be given a parallelogram in which a diagonal is drawn; then “all the squares” of the parallelogram will be triple “all the squares” of each of the triangles determined by the diagonal.

 Cavalieri showed that this proposition could be interpreted in different ways—as asserting, for example, that the volume of a cone is one-third the volume of the circumscribed cylinder (see the figure—>) or that the area under a segment of a parabola is one-third the area of the associated rectangle. In a later treatise he generalized the result by proving

      for n = 3 to n = 9. To establish these results, he introduced transformations among the variables of the problem, using a result equivalent to the binomial theorem for integral exponents. The ideas involved went beyond anything that had appeared in the classical Archimedean theory of content.

      Although Cavalieri was successful in formulating a systematic method based on general concepts, his ideas were not easy to apply. The derivation of very simple results required intricate geometric considerations, and the turgid style of the Geometria Indivisibilibus was a barrier to its reception.

      John Wallis (Wallis, John) presented a quite different approach to the theory of quadratures in his Arithmetica Infinitorum (1655; The Arithmetic of Infinitesimals). Wallis, a successor to Henry Briggs as the Savilian Professor of Geometry at Oxford, was a champion of the new methods of arithmetic algebra that he had learned from his teacher William Oughtred (Oughtred, William). Wallis expressed the area under a curve as the sum of an infinite series and used clever and unrigorous inductions to determine its value. To calculate the area under the parabola,

      he considered the successive sums

      and inferred by “induction” the general relation

      By letting the number of terms be infinite, he obtained 1/3 as the limiting value of the expression. With more complicated curves he achieved very impressive results, including the infinite expression now known as Wallis's product:

      Research on the determination of tangents (tangent), the other subject leading to the calculus, proceeded along different lines. In La Géométrie Descartes had presented a method that could in principle be applied to any algebraic or “geometric” curve—i.e., any curve whose equation was a polynomial of finite degree in two variables. The method depended upon finding the normal, the line perpendicular to the tangent, using the algebraic condition that it be the unique radius to intersect the curve in only one point. Descartes's method was simplified by Hudde, a member of the Leiden group of mathematicians, and was published in 1659 in van Schooten's edition of La Géométrie.

 A class of curves of growing interest in the 17th century comprised those generated kinematically by a point moving through space. The famous cycloidal (cycloid) curve, for example, was traced by a point on the perimeter of a wheel that rolled on a line without slipping or sliding (see the figure—>). These curves were nonalgebraic and hence could not be treated by Descartes's method. Gilles Personne de Roberval (Roberval, Gilles Personne de), professor at the Collège Royale in Paris, devised a method borrowed from dynamics to determine their tangents. In his analysis of projectile motion Galileo had shown that the instantaneous velocity of a particle is compounded of two separate motions: a constant horizontal motion and an increasing vertical motion due to gravity. If the motion of the generating point of a kinematic curve is likewise regarded as the sum of two velocities, then the tangent will lie in the direction of their sum. Roberval applied this idea to several different kinematic curves, obtaining results that were often ingenious and elegant.

 In an essay of 1636 circulated among French mathematicians, Fermat presented a method of tangents adapted from a procedure he had devised to determine maxima and minima and used it to find tangents to several algebraic curves of the form y = xn (see the figure—>). His account was short and contained no explanation of the mathematical basis of the new method. It is possible to see in his procedure an argument involving infinitesimals, and Fermat has sometimes been proclaimed the discoverer of the differential calculus. Modern historical study, however, suggests that he was working with concepts introduced by Viète and that his method was based on finite algebraic ideas.

 Isaac Barrow (Barrow, Isaac), the Lucasian Professor of Mathematics at the University of Cambridge, published in 1670 his Geometrical Lectures, a treatise that more than any other anticipated the unifying ideas of the calculus. In it he adopted a purely geometric form of exposition to show how the determinations of areas and tangents are inverse problems. He began with a curve and considered the slope of its tangent corresponding to each value of the abscissa. He then defined an auxiliary curve by the condition that its ordinate be equal to this slope and showed that the area under the auxiliary curve corresponding to a given abscissa is equal to the rectangle whose sides are unity and the ordinate of the original curve. When reformulated analytically, this result expresses the inverse character of differentiation and integration, the fundamental theorem of the calculus (fundamental theorem of calculus) (see the figure—>). Although Barrow's decision to proceed geometrically prevented him from taking the final step to a true calculus, his lectures influenced both Newton (Newton, Sir Isaac) and Leibniz.

Newton and Leibniz
      The essential insight of Newton and Leibniz was to use Cartesian algebra to synthesize the earlier results and to develop algorithms that could be applied uniformly to a wide class of problems. The formative period of Newton's researches was from 1665 to 1670, while Leibniz worked a few years later, in the 1670s. Their contributions differ in origin, development, and influence, and it is necessary to consider each man separately.

      Newton, the son of an English farmer, became in 1669 the Lucasian Professor of Mathematics at the University of Cambridge. Newton's earliest researches in mathematics grew in 1665 from his study of van Schooten's edition of La Géométrie and Wallis's Arithmetica Infinitorum. Using the Cartesian equation of the curve, he reformulated Wallis's results, introducing for this purpose infinite sums in the powers of an unknown x, now known as infinite series. Possibly under the influence of Barrow, he used infinitesimals (infinitesimal) to establish for various curves the inverse relationship of tangents and areas. The operations of differentiation and integration emerged in his work as analytic processes that could be applied generally to investigate curves.

      Unusually sensitive to questions of rigour, Newton at a fairly early stage tried to establish his new method on a sound foundation using ideas from kinematics. A variable was regarded as a “fluent,” a magnitude that flows with time; its derivative or rate of change with respect to time was called a “ fluxion,” denoted by the given variable with a dot above it. The basic problem of the calculus was to investigate relations among fluents and their fluxions. Newton finished a treatise on the method of fluxions as early as 1671, although it was not published until 1736. In the 18th century this method became the preferred approach to the calculus among British mathematicians, especially after the appearance in 1742 of Colin Maclaurin (Maclaurin, Colin)'s influential Treatise of Fluxions.

      Newton first published the calculus in Book I of his great Philosophiae Naturalis Principia Mathematica (1687; Mathematical Principles of Natural Philosophy). Originating as a treatise on the dynamics of particles, the Principia presented an inertial physics that combined Galileo's mechanics and Kepler's planetary astronomy. It was written in the early 1680s at a time when Newton was reacting against Descartes's science and mathematics. Setting aside the analytic method of fluxions, Newton introduced in 11 introductory lemmas his calculus of first and last ratios, a geometric theory of limits that provided the mathematical basis of his dynamics.

      Newton's use of the calculus in the Principia is illustrated by proposition 11 of Book I: if the orbit of a particle moving under a centripetal force is an ellipse with the centre of force at one focus, then the force is inversely proportional to the square of the distance from the centre. Because the planets were known by Kepler's laws to move in ellipses with the Sun at one focus, this result supported his inverse square law of gravitation. To establish the proposition, Newton derived an approximate measure for the force by using small lines defined in terms of the radius (the line from the force centre to the particle) and the tangent to the curve at a point. This result expressed geometrically the proportionality of force to vector acceleration. Using properties of the ellipse known from classical geometry, Newton calculated the limit of this measure and showed that it was equal to a constant times 1 over the square of the radius.

      Newton avoided analytic processes in the Principia by expressing magnitudes and ratios directly in terms of geometric quantities, both finite and infinitesimal. His decision to eschew analysis constituted a striking rejection of the algebraic (algebra) methods that had been important in his own early researches on the calculus. Although the Principia was of inestimable value for later mechanics, it would be reworked by researchers on the Continent and expressed in the mathematical idiom of the Leibnizian (Leibniz, Gottfried Wilhelm) calculus.

      Leibniz's interest in mathematics was aroused in 1672 during a visit to Paris, where the Dutch mathematician Christiaan Huygens (Huygens, Christiaan) introduced him to his work on the theory of curves. Under Huygens's tutelage Leibniz immersed himself for the next several years in the study of mathematics. He investigated relationships between the summing and differencing of finite and infinite sequences of numbers. Having read Barrow's geometric lectures, he devised a transformation rule to calculate quadratures, obtaining the famous infinite series for π/4:

      Leibniz was interested in questions of logic and notation, of how to construct a characteristica universalis for rational investigation. After considerable experimentation he arrived by the late 1670s at an algorithm based on the symbols d and ∫. He first published his research on differential calculus in 1684 in an article in the Acta Eruditorum, Nova Methodus pro Maximis et Minimis, Itemque Tangentibus, qua nec Fractas nec Irrationales Quantitates Moratur, et Singulare pro illi Calculi Genus (“A New Method for Maxima and Minima as Well as Tangents, Which Is Impeded Neither by Fractional nor by Irrational Quantities, and a Remarkable Type of Calculus for This”). In this article he introduced the differential dx satisfying the rules d(x + y) = dx + dy and d(xy) = xdy + ydx and illustrated his calculus with a few examples. Two years later he published a second article, On a Deeply Hidden Geometry, in which he introduced and explained the symbol ∫ for integration. He stressed the power of his calculus to investigate transcendental curves, the very class of “mechanical” objects Descartes had believed lay beyond the power of analysis, and derived a simple analytic formula for the cycloid.

      Leibniz continued to publish results on the new calculus in the Acta Eruditorum and began to explore his ideas in extensive correspondence with other scholars. Within a few years he had attracted a group of researchers to promulgate his methods, including the brothers Johann Bernoulli (Bernoulli, Johann) and Jakob Bernoulli (Bernoulli, Jakob) in Basel and the priest Pierre Varignon and Guillaume-François-Antoine de L'Hospital in Paris. In 1700 he persuaded Frederick William I of Prussia to establish the Brandenburg Society of Sciences (later renamed the Berlin Academy of Sciences), with himself appointed president for life.

      Leibniz's vigorous espousal of the new calculus, the didactic spirit of his writings, and his ability to attract a community of researchers contributed to his enormous influence on subsequent mathematics. In contrast, Newton's slowness to publish and his personal reticence resulted in a reduced presence within European mathematics. Although the British school in the 18th century included capable researchers, Abraham de Moivre (Moivre, Abraham de), James Stirling (Stirling, James), Brook Taylor (Taylor, Brook), and Maclaurin among them, they failed to establish a program of research comparable to that established by Leibniz's followers on the Continent. There is a certain tragedy in Newton's isolation and his reluctance to acknowledge the superiority of continental analysis. As the historian Michael Mahoney observed:

Whatever the revolutionary influence of the Principia, mathematics would have looked much the same if Newton had never existed. In that endeavour he belonged to a community, and he was far from indispensable to it.

The 18th century
Institutional background
      After 1700 a movement to found learned societies on the model of Paris and London spread throughout Europe and the American colonies. The academy was the predominant institution of science until it was displaced by the university in the 19th century. The leading mathematicians of the period, such as Leonhard Euler (Euler, Leonhard), Jean Le Rond d'Alembert (Alembert, Jean Le Rond d'), and Joseph-Louis Lagrange (Lagrange, Joseph-Louis, comte de l'Empire), pursued academic careers at St. Petersburg, Paris, and London.

      The French Academy of Sciences (Sciences, Academy of) (Paris) provides an informative study of the 18th-century learned society. The academy was divided into six sections, three for the mathematical and three for the physical sciences. The mathematical sections were for geometry, astronomy, and mechanics, the physical sections for chemistry, anatomy, and botany. Membership in the academy was divided by section, with each section contributing three pensionnaires, two associates, and two adjuncts. There was also a group of free associates, distinguished men of science from the provinces, and foreign associates, eminent international figures in the field. A larger group of 70 corresponding members had partial privileges, including the right to communicate reports to the academy. The administrative core consisted of a permanent secretary, treasurer, president, and vice president. In a given year the average total membership in the academy was 153.

      Prominent characteristics of the academy included its small and elite membership, made up heavily of men from the middle class, and its emphasis on the mathematical sciences. In addition to holding regular meetings and publishing memoirs, the academy organized scientific expeditions and administered prize competitions on important mathematical and scientific questions.

      The historian Roger Hahn noted that the academy in the 18th century allowed “the coupling of relative doctrinal freedom on scientific questions with rigorous evaluations by peers,” an important characteristic of modern professional science. Academic mathematics and science did, however, foster a stronger individualistic ethos than is usual today. A determined individual such as Euler or Lagrange could emphasize a given program of research through his own work, the publications of the academy, and the setting of the prize competitions. The academy as an institution may have been more conducive to the solitary patterns of research in a theoretical subject like mathematics than it was to the experimental sciences. The separation of research from teaching is perhaps the most striking characteristic that distinguished the academy from the model of university-based science that developed in the 19th century.

Analysis and mechanics
      The scientific revolution had bequeathed to mathematics a major program of research in analysis and mechanics. The period from 1700 to 1800, “the century of analysis,” witnessed the consolidation of the calculus and its extensive application to mechanics. With expansion came specialization as different parts of the subject acquired their own identity: ordinary (ordinary differential equation) and partial differential equations (partial differential equation), calculus of variations, infinite series, and differential geometry. The applications of analysis were also varied, including the theory of the vibrating string, particle dynamics, the theory of rigid bodies, the mechanics of flexible and elastic media, and the theory of compressible and incompressible fluids. Analysis and mechanics developed in close association, with problems in one giving rise to concepts and techniques in the other, and all the leading mathematicians of the period made important contributions to mechanics.

      The close relationship between mathematics and mechanics in the 18th century had roots extending deep into Enlightenment thought. In the organizational chart of knowledge at the beginning of the preliminary discourse to the Encyclopédie, Jean Le Rond d'Alembert (Alembert, Jean Le Rond d') distinguished between “pure” mathematics (geometry, arithmetic, algebra, calculus) and “mixed” mathematics (mechanics, geometric astronomy, optics, art of conjecturing). Mathematics generally was classified as a “science of nature” and separated from logic, a “science of man.” The modern disciplinary division between physics and mathematics and the association of the latter with logic had not yet developed.

      Mathematical mechanics itself as it was practiced in the 18th century differed in important respects from later physics. The goal of modern physics is to explore the ultimate particulate structure of matter and to arrive at fundamental laws of nature to explain physical phenomena. The character of applied investigation in the 18th century was rather different. The material parts of a given system and their interrelationship were idealized for the purposes of analysis. A material object could be treated as a point-mass (a mathematical point at which it is assumed all the mass of the object is concentrated), as a rigid body, as a continuously deformable medium, and so on. The intent was to obtain a mathematical description of the macroscopic behaviour of the system rather than to ascertain the ultimate physical basis of the phenomena. In this respect the 18th-century viewpoint is closer to modern mathematical engineering than it is to physics.

      Mathematical research in the 18th century was coordinated by the Paris, Berlin, and St. Petersburg academies, as well as by several smaller provincial scientific academies and societies. Although England and Scotland were important centres early in the century, with Maclaurin's death in 1746 the British flame was all but extinguished.

History of analysis
      The history of analysis in the 18th century can be followed in the official memoirs of the academies and in independently published expository treatises. In the first decades of the century the calculus was cultivated in an atmosphere of intellectual excitement as mathematicians applied the new methods to a range of problems in the geometry of curves. The brothers Johann (Bernoulli, Johann) and Jakob Bernoulli (Bernoulli, Jakob) showed that the shape of a smooth wire along which a particle descends in the least time is the cycloid, a transcendental curve much studied in the previous century. Working in a spirit of keen rivalry, the two brothers arrived at ideas that would later develop into the calculus of variations. In his study of the rectification of the lemniscate, a ribbon-shaped curve discovered by Jakob Bernoulli in 1694, Giulio Carlo Fagnano (1682–1766) introduced ingenious analytic transformations that laid the foundation for the theory of elliptic integrals. Nikolaus I Bernoulli (1687–1759), the nephew of Johann and Jakob, proved the equality of mixed second-order partial derivatives and made important contributions to differential equations by the construction of orthogonal trajectories to families of curves. Pierre Varignon (1654–1722), Johann Bernoulli, and Jakob Hermann (1678–1733) continued to develop analytic dynamics as they adapted Leibniz's calculus to the inertial mechanics of Newton's Principia.

      Geometric conceptions and problems predominated in the early calculus. This emphasis on the curve as the object of study provided coherence to what was otherwise a disparate collection of analytic techniques. With its continued development, the calculus gradually became removed from its origins in the geometry of curves, and a movement emerged to establish the subject on a purely analytic basis. In a series of textbooks published in the middle of the century, the Swiss mathematician Leonhard Euler (Euler, Leonhard) systematically accomplished the separation of the calculus from geometry. In his Introductio in Analysin Infinitorum (1748; Introduction to the Analysis of the Infinite), he made the notion of function the central organizing concept of analysis:

A function of a variable quantity is an analytical expression composed in any way from the variable and from numbers or constant quantities.

      Euler's analytic approach is illustrated by his introduction of the sine and cosine functions. Trigonometry tables (trigonometry table) had existed since antiquity, and the relations between sines and cosines were commonly used in mathematical astronomy. In the early calculus mathematicians had derived in their study of periodic mechanical phenomena the differential equation

      and they were able to interpret its solution geometrically in terms of lines and angles in the circle. Euler was the first to introduce the sine and cosine functions as quantities whose relation to other quantities could be studied independently of any geometric diagram.

      Euler's analytic approach to the calculus received support from his younger contemporary Joseph-Louis Lagrange (Lagrange, Joseph-Louis, comte de l'Empire), who, following Euler's death in 1783, replaced him as the leader of European mathematics. In 1755 the 19-year-old Lagrange wrote to Euler to announce the discovery of a new algorithm in the calculus of variations, a subject to which Euler had devoted an important treatise 11 years earlier. Euler had used geometric ideas extensively and had emphasized the need for analytic methods. Lagrange's idea was to introduce the new symbol δ into the calculus and to experiment formally until he had devised an algorithm to obtain the variational equations. Mathematically quite distinct from Euler's procedure, his method required no reference to the geometric configuration. Euler immediately adopted Lagrange's idea, and in the next several years the two men systematically revised the subject using the new techniques.

      In 1766 Lagrange was invited by the Prussian king, Frederick the Great (Frederick II), to become mathematics director of the Berlin Academy. During the next two decades he wrote important memoirs on nearly all of the major areas of mathematics. In 1788 he published his famous Mécanique analytique, a treatise that used variational ideas to present mechanics from a unified analytic viewpoint. In the preface Lagrange wrote:

One will find no Figures in this Work. The methods that I present require neither constructions nor geometrical or mechanical reasonings, but only algebraic operations, subject to a regular and uniform course. Those who admire Analysis, will with pleasure see Mechanics become a new branch of it, and will be grateful to me for having extended its domain.

      Following the death of Frederick the Great, Lagrange traveled to Paris to become a pensionnaire of the Academy of Sciences. With the establishment of the École Polytechnique (French: “Polytechnic School”) in 1794, he was asked to deliver the lectures on mathematics. There was a concern in European mathematics at the time to place the calculus on a sound basis, and Lagrange used the occasion to develop his ideas for an algebraic (algebra) foundation of the subject. The lectures were published in 1797 under the title Théorie des fonctions analytiques (“Theory of Analytical Functions”), a treatise whose contents were summarized in its longer title, “Containing the Principles of the Differential Calculus Disengaged from All Consideration of Infinitesimals, Vanishing Limits, or Fluxions and Reduced to the Algebraic Analysis of Finite Quantities.” Lagrange published a second treatise on the subject in 1801, a work that appeared in a revised and expanded form in 1806.

      The range of subjects presented and the consistency of style distinguished Lagrange's didactic writings from other contemporary expositions of the calculus. He began with Euler's notion of a function as an analytic expression composed of variables and constants. He defined the “derived function,” or derivative f′(x) of f(x), to be the coefficient of i in the Taylor (Taylor series) expansion of f(x + i). Assuming the general possibility of such expansions, he attempted a rather complete theory of the differential and integral calculus, including extensive applications to geometry and mechanics. Lagrange's lectures represented the most advanced development of the 18th-century analytic conception of the calculus.

      Beginning with Baron Cauchy (Cauchy, Augustin-Louis, Baron) in the 1820s, later mathematicians used the concept of limit to establish the calculus on an arithmetic basis. The algebraic viewpoint of Euler and Lagrange was rejected. To arrive at a proper historical appreciation of their work, it is necessary to reflect on the meaning of analysis in the 18th century. Since Viète, analysis had referred generally to mathematical methods that employed equations, variables, and constants. With the extensive development of the calculus by Leibniz and his school, analysis became identified with all calculus-related subjects. In addition to this historical association, there was a deeper sense in which analytic methods were fundamental to the new mathematics. An analytic equation implied the existence of a relation that remained valid as the variables changed continuously in magnitude. Analytic algorithms and transformations presupposed a correspondence between local and global change, the basic concern of the calculus. It is this aspect of analysis that fascinated Euler and Lagrange and caused them to see in it the “true metaphysics” of the calculus.

Other developments
      During the period 1600–1800 significant advances occurred in the theory of equations, foundations of Euclidean geometry, number theory, projective geometry, and probability theory (probability and statistics). These subjects, which became mature branches of mathematics only in the 19th century, never rivaled analysis and mechanics as programs of research.

      After the dramatic successes of Niccolò Fontana Tartaglia (Tartaglia, Niccolò Fontana) and Lodovico Ferrari in the 16th century, the theory of equations developed slowly, as problems resisted solution by known techniques. In the later 18th century the subject experienced an infusion of new ideas. Interest was concentrated on two problems. The first was to establish the existence of a root of the general polynomial equation of degree n. The second was to express the roots as algebraic functions of the coefficients or to show why it was not, in general, possible to do so.

      The proposition that the general polynomial with real coefficients has a root of the form a + b √(−1) became known later as the fundamental theorem of algebra. By 1742 Euler had recognized that roots appear in conjugate pairs; if a + b √(−1) is a root, then so is a − b √(−1) . Thus, if a + b √(−1) is a root of f(x) = 0, then f(x) = (x2 − 2ax − a2 − b2)g(x). The fundamental theorem (fundamental theorem of algebra) was therefore equivalent to asserting that a polynomial may be decomposed into linear and quadratic factors. This result was of considerable importance for the theory of integration, since by the method of partial fractions it ensured that a rational function, the quotient of two polynomials, could always be integrated in terms of algebraic and elementary transcendental functions.

      Although d'Alembert, Euler, and Lagrange worked on the fundamental theorem, the first successful proof was developed by Carl Friedrich Gauss (Gauss, Carl Friedrich) in his doctoral dissertation of 1799. Earlier researchers had investigated special cases or had concentrated on showing that all possible roots were of the form a ± b √(−1) . Gauss tackled the problem of existence directly. Expressing the unknown in terms of the polar coordinate (polar coordinates) variables r and θ, he showed that a solution of the polynomial would lie at the intersection of two curves of the form T(r, θ) = 0 and U(r, θ) = 0. By a careful and rigorous investigation he proved that the two curves intersect.

      Gauss's demonstration of the fundamental theorem initiated a new approach to the question of mathematical existence. In the 18th century mathematicians were interested in the nature of particular analytic processes or the form that given solutions should take. Mathematical entities were regarded as things that were given, not as things whose existence needed to be established. Because analysis was applied in geometry and mechanics, the formalism seemed to possess a clear interpretation that obviated any need to consider questions of existence. Gauss's demonstration was the beginning of a change of attitude in mathematics, of a shift to the rigorous, internal development of the subject.

      The problem of expressing the roots of a polynomial as functions of the coefficients was addressed by several mathematicians independently about 1770. The Cambridge mathematician Edward Waring (Waring, Edward) published treatises in 1762 and 1770 on the theory of equations. In 1770 Lagrange presented a long expository memoir on the subject to the Berlin Academy, and in 1771 Alexandre Vandermonde submitted a paper to the French Academy of Sciences. Although the ideas of the three men were related, Lagrange's memoir was the most extensive and most influential historically.

      Lagrange presented a detailed analysis of the solution by radicals of second-, third-, and fourth-degree equations and investigated why these solutions failed when the degree was greater than or equal to five. He introduced the novel idea of considering functions of the roots and examining the values they assumed as the roots were permuted. He was able to show that the solution of an equation depends on the construction of a second resolvent equation, but he was unable to provide a general procedure for solving the resolvent when the degree of the original equation was greater than four. Although his theory left the subject in an unfinished condition, it provided a solid basis for future work. The search for a general solution to the polynomial equation would provide the greatest single impetus for the transformation of algebra in the 19th century.

      The efforts of Lagrange, Vandermonde, and Waring illustrate how difficult it was to develop new concepts in algebra. The history of the theory of equations belies the view that mathematics is subject to almost automatic technical development. Much of the later algebraic work would be devoted to devising terminology, concepts, and methods necessary to advance the subject.

Foundations of (mathematics, foundations of) geometry
      Although the emphasis of mathematics after 1650 was increasingly on analysis, foundational questions in classical geometry continued to arouse interest. Attention centred on the fifth postulate of Book I of the Elements, which Euclid had used to prove the existence of a unique parallel through a point to a given line. Since antiquity, Greek, Islamic, and European geometers had attempted unsuccessfully to show that the parallel postulate need not be a postulate but could instead be deduced from the other postulates of Euclidean geometry. During the period 1600–1800 mathematicians continued these efforts by trying to show that the postulate was equivalent to some result that was considered self-evident. Although the decisive breakthrough to non-Euclidean geometry would not occur until the 19th century, researchers did achieve a deeper and more systematic understanding of the classical properties of space.

      Interest in the parallel postulate developed in the 16th century after the recovery and Latin translation of Proclus's commentary on Euclid's Elements. The Italian researchers Christopher Clavius in 1574 and Giordano Vitale in 1680 showed that the postulate is equivalent to asserting that the line equidistant from a straight line is a straight line. In 1693 John Wallis (Wallis, John), Savilian Professor of Geometry at Oxford, attempted a different demonstration, proving that the axiom follows from the assumption that to every figure there exists a similar figure of arbitrary magnitude.

 In 1733 the Italian Girolamo Saccheri published his Euclides ab Omni Naevo Vindicatus (“Euclid Cleared of Every Flaw”). This was an important work of synthesis in which he provided a complete analysis of the problem of parallels in terms of Omar Khayyam's quadrilateral (see the figure—>). Using the Euclidean assumption that straight lines do not enclose an area, he was able to exclude geometries that contain no parallels. It remained to prove the existence of a unique parallel through a point to a given line. To do this, Saccheri adopted the procedure of reductio ad absurdum; he assumed the existence of more than one parallel and attempted to derive a contradiction. After a long and detailed investigation, he was able to convince himself (mistakenly) that he had found the desired contradiction.

      In 1766 Johann Heinrich Lambert (Lambert, Johann Heinrich) of the Berlin Academy composed Die Theorie der Parallellinien (“The Theory of Parallel Lines”; published 1786), a penetrating study of the fifth postulate in Euclidean geometry. Among other theorems Lambert proved is that the parallel axiom is equivalent to the assertion that the sum of the angles of a triangle is equal to two right angles. He combined this fact with Wallis's result to arrive at an unexpected characterization of classical space. According to Lambert, if the parallel postulate is rejected, it follows that for every angle θ less than 2R/3 (R is a right angle) an equilateral triangle can be constructed with corner angle θ. By Wallis's result any triangle similar to this triangle must be congruent to it. It is therefore possible to associate with every angle a definite length, the side of the corresponding equilateral triangle. Since the measurement of angles is absolute, independent of any convention concerning the selection of units, it follows that an absolute unit of length exists. Hence, to accept the parallel postulate is to deny the possibility of an absolute concept of length.

      The final 18th-century contribution to the theory of parallels was Adrien-Marie Legendre (Legendre, Adrien-Marie)'s textbook Éléments de géométrie (Elements of Geometry and Trigonometry), the first edition of which appeared in 1794. Legendre presented an elegant demonstration that purported to show that the sum of the angles of a triangle is equal to two right angles. He believed that he had conclusively established the validity of the parallel postulate. His work attracted a large audience and was influential in informing readers of the new ideas in geometry.

      The 18th-century failure to develop a non-Euclidean geometry was rooted in deeply held philosophical beliefs. In his Critique of Pure Reason (1781), Immanuel Kant (Kant, Immanuel) had emphasized the synthetic a priori character of mathematical judgments. From this standpoint, statements of geometry and arithmetic were necessarily true propositions with definite empirical content. The existence of similar figures of different size, or the conventional character of units of length, appeared self-evident to mathematicians of the period. As late as 1824 Pierre-Simon, marquis de Laplace (Laplace, Pierre-Simon, marquis de), wrote:

Thus the notion of space includes a special property, self-evident, without which the properties of parallels cannot be rigorously established. The idea of a bounded region, e.g., the circle, contains nothing which depends on its absolute magnitude. But if we imagine its radius to diminish, we are brought without fail to the diminution in the same ratio of its circumference and the sides of all the inscribed figures. This proportionality appears to me a more natural postulate than that of Euclid, and it is worthy of note that it is discovered afresh in the results of the theory of universal gravitation.

Craig G. Fraser

Mathematics in the 19th and 20th centuries
      Most of the powerful abstract mathematical theories in use today originated in the 19th century, so any historical account of the period should be supplemented by reference to detailed treatments of these topics. Yet mathematics grew so much during this period that any account must necessarily be selective. Nonetheless, some broad features stand out. The growth of mathematics as a profession was accompanied by a sharpening division between mathematics and the physical sciences, and contact between the two subjects takes place today across a clear professional boundary. One result of this separation has been that mathematics, no longer able to rely on its scientific import for its validity, developed markedly higher standards of rigour. It was also freed to develop in directions that had little to do with applicability. Some of these pure creations have turned out to be surprisingly applicable, while the attention to rigour has led to a wholly novel conception of the nature of mathematics and logic. Moreover, many outstanding questions in mathematics yielded to the more conceptual approaches that came into vogue.

      The French Revolution provoked a radical rethinking of education in France, and mathematics was given a prominent role. The École Polytechnique was established in 1794 with the ambitious task of preparing all candidates for the specialist civil and military engineering schools of the republic. Mathematicians of the highest calibre were involved; the result was a rapid and sustained development of the subject. The inspiration for the École was that of Gaspard Monge (Monge, Gaspard, comte de Péluse), who believed strongly that mathematics should serve the scientific and technical needs of the state. To that end he devised a syllabus that promoted his own descriptive geometry, which was useful in the design of forts, gun emplacements, and machines and which was employed to great effect in the Napoleonic survey of Egyptian historical sites.

      In Monge's descriptive geometry, three-dimensional objects are described by their orthogonal projections onto a horizontal and a vertical plane, the plan and elevation of the object. A pupil of Monge, Jean-Victor Poncelet (Poncelet, Jean-Victor), was taken prisoner during Napoleon's retreat from Moscow and sought to keep up his spirits while in jail in Saratov by thinking over the geometry he had learned. He dispensed with the restriction to orthogonal projections and decided to investigate what properties figures have in common with their shadows. There are several of these properties: a straight line casts a straight shadow, and a tangent to a curve casts a shadow that is tangent to the shadow of the curve. But some properties are lost: the lengths and angles of a figure bear no relation to the lengths and angles of its shadow. Poncelet felt that the properties that survive are worthy of study, and, by considering only those properties that a figure shares with all its shadows, Poncelet hoped to put truly geometric reasoning on a par with algebraic geometry.

      In 1822 Poncelet published the Traité des propriétés projectives des figures (“Treatise on the Projective Properties of Figures”). From his standpoint every conic section is equivalent to a circle, so his treatise contained a unified treatment of the theory of conic sections. It also established several new results. Geometers who took up his work divided into two groups: those who accepted his terms and those who, finding them obscure, reformulated his ideas in the spirit of algebraic geometry. On the algebraic side it was taken up in Germany by August Ferdinand Möbius (Möbius, August Ferdinand), who seems to have come to his ideas independently of Poncelet, and then by Julius Plücker (Plücker, Julius). They showed how rich was the projective geometry of curves defined by algebraic equations and thereby gave an enormous boost to the algebraic study of curves, comparable to the original impetus provided by Descartes. Germany also produced synthetic projective geometers, notably Jakob Steiner (Steiner, Jakob) (born in Switzerland but educated in Germany) and Karl Georg Christian von Staudt (Staudt, Karl Georg Christian von), who emphasized what can be understood about a figure from a careful consideration of all its transformations.

 Within the debates about projective geometry emerged one of the few synthetic ideas to be discovered since the days of Euclid, that of duality. This associates with each point a line and with each line a point, in such a way that (1) three points lying in a line give rise to three lines meeting in a point and, conversely, three lines meeting in a point give rise to three points lying on a line and (2) if one starts with a point (or a line) and passes to the associated line (point) and then repeats the process, one returns to the original point (line). One way of using duality (presented by Poncelet) is to pick an arbitrary conic and then to associate with a point P lying outside the conic the line that joins the points R and S at which the tangents through P to the conic touch the conic (see the figure—>). A second method is needed for points on or inside the conic. The feature of duality that makes it so exciting is that one can apply it mechanically to every proof in geometry, interchanging “point” and line” and “collinear” and “concurrent” throughout, and so obtain a new result. Sometimes a result turns out to be equivalent to the original, sometimes to its converse, but at a single stroke the number of theorems was more or less doubled.

Making the calculus rigorous
      Monge's educational ideas were opposed by Lagrange, who favoured a more traditional and theoretical diet of advanced calculus and rational mechanics (the application of the calculus to the study of the motion of solids and liquids). Eventually Lagrange won, and the vision of mathematics that was presented to the world was that of an autonomous subject that was also applicable to a broad range of phenomena by virtue of its great generality, a view that has persisted to the present day.

      During the 1820s Augustin-Louis, Baron Cauchy (Cauchy, Augustin-Louis, Baron), lectured at the École Polytechnique on the foundations of the calculus. Since its invention it had been generally agreed that the calculus gave correct answers, but no one had been able to give a satisfactory explanation of why this was so. Cauchy rejected Lagrange's algebraic approach and proved that Lagrange's basic assumption that every function has a power series expansion is in fact false. Newton had suggested a geometric or dynamic basis for calculus, but this ran the risk of introducing a vicious circle when the calculus was applied to mechanical or geometric problems. Cauchy proposed basing the calculus on a sophisticated and difficult interpretation of the idea of two points or numbers being arbitrarily close together. Although his students disliked the new approach, and Cauchy was ordered to teach material that the students could actually understand and use, his methods gradually became established and refined to form the core of the modern rigorous calculus, a subject now called mathematical analysis.

      Traditionally, the calculus had been concerned with the two processes of differentiation and integration and the reciprocal relation that exists between them. Cauchy provided a novel underpinning by stressing the importance of the concept of continuity, which is more basic than either. He showed that, once the concepts of a continuous function and limit are defined, the concepts of a differentiable function and an integrable function can be defined in terms of them. Unfortunately, neither of these concepts is easy to grasp, and the much-needed degree of precision they bring to mathematics has proved difficult to appreciate. Roughly speaking, a function is continuous at a point in its domain if small changes in the input around the specified value produce only small changes in the output.

 Thus, the familiar graph—> of a parabola y = x2 is continuous around the point x = 0; as x varies by small amounts, so necessarily does y. On the other hand, the graph of the function that takes the value 0 when x is negative or zero, and the value 1 when x is positive, plainly has a discontinuous graph at the point x = 0, and it is indeed discontinuous there according to the definition. If x varies from 0 by any small positive amount, the value of the function jumps by the fixed amount 1, which is not an arbitrarily small amount.

      Cauchy said that a function f(x) tends to a limiting value 1 as x tends to the value a whenever the value of the difference f(x) − f(a) becomes arbitrarily small as the difference x − a itself becomes arbitrarily small. He then showed that, if f(x) is continuous at a, the limiting value of the function as x tended to a was indeed f(a). The crucial feature of this definition is that it defines what it means for a variable quantity to tend to something entirely without reference to ideas of motion.

  Cauchy then said a function f(x) is differentiable at the point a if, as x tends to a (which it is never allowed to reach), the value of the quotient [f(x) − f(a)]/(x − a) (see the figure—>, left) tends to a limiting value, called the derivative of the function f(x) at a. To define the integral of a function f(x) between the values a and b, Cauchy went back to the primitive idea of the integral as the measure of the area under the graph of the function. He approximated this area by rectangles and said that, if the sum of the areas of the rectangles tends to a limit as their number increases indefinitely (see the figure—>, right) and if this limiting value is the same however the rectangles are obtained, then the function is integrable. Its integral is the common limiting value. After he had defined the integral independently of the differential calculus, Cauchy had to prove that the processes of integrating and differentiating are mutually inverse. This he did, giving for the first time a rigorous foundation to all the elementary calculus of his day.

Fourier series
      The other crucial figure of the time in France was Joseph, Baron Fourier (Fourier, Joseph, Baron). His major contribution, presented in The Analytical Theory of Heat (1822), was to the theory of heat diffusion in solid bodies. He proposed that any function could be written as an infinite sum of the trigonometric functions cosine and sine; for example,

      Expressions of this kind had been written down earlier, but Fourier's treatment was new in the degree of attention given to their convergence. He investigated the question “Given the function f(x), for what range of values of x does the expression above sum to a finite number?” It turned out that the answer depends on the coefficients an, and Fourier gave rules for obtaining them of the form

      Had Fourier's work been entirely correct, it would have brought all functions into the calculus, making possible the solution of many kinds of differential equations and greatly extending the theory of mathematical physics. But his arguments were unduly naive: after Cauchy it was not clear that the function f(x) sin (nx) was necessarily integrable. When Fourier's ideas were finally published, they were eagerly taken up, but the more cautious mathematicians, notably the influential German Peter Gustav Lejeune Dirichlet (Dirichlet, Peter Gustav Lejeune), wanted to rederive Fourier's conclusions in a more rigorous way. Fourier's methodology was widely accepted, but questions about its validity in detail were to occupy mathematicians for the rest of the century.

Elliptic functions
      The theory of functions of a complex variable was also being decisively reformulated. At the start of the 19th century, complex numbers (complex number) were discussed from a quasi-philosophical standpoint by several French writers, notably Jean-Robert Argand. A consensus emerged that complex numbers should be thought of as pairs of real numbers, with suitable rules for their addition and multiplication so that the pair (0, 1) was a square root of −1. The underlying meaning of such a number pair was given by its geometric interpretation either as a point in a plane or as a directed segment joining the coordinate origin to the point in question. (This representation is sometimes called the Argand diagram.) In 1827, while revising an earlier manuscript for publication, Cauchy showed how the problem of integrating functions of two variables can be illuminated by a theory of functions of a single complex variable, which he was then developing. But the decisive influence on the growth of the subject came from the theory of elliptic functions (elliptic equation).

      The study of elliptic functions originated in the 18th century, when many authors studied integrals of the form

      where p(t) and q(t) are polynomials in t and q(t) is of degree 3 or 4 in t. Such integrals arise naturally, for example, as an expression for the length of an arc of an ellipse (whence the name). These integrals cannot be evaluated explicitly; they do not define a function that can be obtained from the rational and trigonometric functions, a difficulty that added to their interest. Elliptic integrals were intensively studied for many years by the French mathematician Legendre, who was able to calculate tables of values for such expressions as functions of their upper endpoint, x. But the topic was completely transformed in the late 1820s by the independent but closely overlapping discoveries of two young mathematicians, the Norwegian Niels Henrik Abel (Abel, Niels Henrik) and the German Carl Jacobi (Jacobi, Carl). These men showed that, if one allowed the variable x to be complex and the problem was inverted, so that the object of study became

      considered as defining a function x of a variable u, then a remarkable new theory became apparent. The new function, for example, possessed a property that generalized the basic property of periodicity of the trigonometric functions sine and cosine: sin (x) = sin (x + 2π). Any function of the kind just described has two distinct periods, ω1 and ω2:

      These new functions, the elliptic functions, aroused a considerable degree of interest. The analogy with trigonometric functions ran very deep (indeed the trigonometric functions turned out to be special cases of elliptic functions), but their greatest influence was on the burgeoning general study of functions of a complex variable. The theory of elliptic functions became the paradigm of what could be discovered by allowing variables to be complex instead of real. But their natural generalization to functions defined by more complicated integrands, although it yielded partial results, resisted analysis until the second half of the 19th century.

The theory of numbers (number theory)
      While the theory of elliptic functions typifies the 19th century's enthusiasm for pure mathematics, some contemporary mathematicians said that the simultaneous developments in number theory carried that enthusiasm to excess. Nonetheless, during the 19th century the algebraic theory of numbers grew from being a minority interest to its present central importance in pure mathematics. The earlier investigations of Fermat had eventually drawn the attention of Euler and Lagrange. Euler proved some of Fermat's unproven claims and discovered many new and surprising facts; Lagrange not only supplied proofs of many remarks that Euler had merely conjectured but also worked them into something like a coherent theory. For example, it was known to Fermat that the numbers that can be written as the sum of two squares are the number 2, squares themselves, primes of the form 4n + 1, and products of these numbers. Thus, 29, which is 4 × 7 + 1, is 52 + 22, but 35, which is not of this form, cannot be written as the sum of two squares. Euler had proved this result and had gone on to consider similar cases, such as primes of the form x2 + 2y2 or x2 + 3y2. But it was left to Lagrange to provide a general theory covering all expressions of the form ax2 + bxy + cy2, quadratic forms, as they are called.

      Lagrange's theory of quadratic forms had made considerable use of the idea that a given quadratic form could often be simplified to another with the same properties but with smaller coefficients. To do this in practice, it was often necessary to consider whether a given integer left a remainder that was a square when it was divided by another given integer. (For example, 48 leaves a remainder of 4 upon division by 11, and 4 is a square.) Legendre discovered a remarkable connection between the question “Does the integer p leave a square remainder on division by q?” and the seemingly unrelated question “Does the integer q leave a square remainder upon division by p?” He saw, in fact, that, when p and q are primes, both questions have the same answer unless both primes are of the form 4n − 1. Because this observation connects two questions in which the integers p and q play mutually opposite roles, it became known as the law of quadratic reciprocity. Legendre also gave an effective way of extending his law to cases when p and q are not prime.

      All this work set the scene for the emergence of Carl Friedrich Gauss (Gauss, Carl Friedrich), whose Disquisitiones Arithmeticae (1801) not only consummated what had gone before but also directed number theorists in new and deeper directions. He rightly showed that Legendre's proof of the law of quadratic reciprocity was fundamentally flawed and gave the first rigorous proof. His work suggested that there were profound connections between the original question and other branches of number theory, a fact that he perceived to be of signal importance for the subject. He extended Lagrange's theory of quadratic forms by showing how two quadratic forms can be “multiplied” to obtain a third. Later mathematicians were to rework this into an important example of the theory of finite commutative groups (group). And in the long final section of his book, Gauss gave the theory that lay behind his first discovery as a mathematician: that a regular 17-sided figure can be constructed by circle and straightedge alone.

      The discovery that the regular “17-gon” is so constructible was the first such discovery since the Greeks, who had known only of the equilateral triangle, the square, the regular pentagon, the regular 15-sided figure, and the figures that can be obtained from these by successively bisecting all the sides. But what was of much greater significance than the discovery was the theory that underpinned it, the theory of what are now called algebraic numbers (algebraic number). It may be thought of as an analysis of how complicated a number may be while yet being amenable to an exact treatment.

      The simplest numbers to understand and use are the integers and the rational numbers. The irrational numbers (irrational number) seem to pose problems. Famous among these is √2. It cannot be written as a finite or repeating decimal (because it is not rational), but it can be manipulated algebraically very easily. It is necessary only to replace every occurrence of (√2)2 by 2. In this way expressions of the form m + n√2, where m and n are integers, can be handled arithmetically. These expressions have many properties akin to those of whole numbers, and mathematicians have even defined prime numbers of this form; therefore, they are called algebraic integers. In this case they are obtained by grafting onto the rational numbers a solution of the polynomial equation x2 − 2 = 0. In general an algebraic integer is any solution, real or complex, of a polynomial equation with integer coefficients in which the coefficient of the highest power of the unknown is 1.

      Gauss's theory of algebraic integers led to the question of determining when a polynomial of degree n with integer coefficients can be solved given the solvability of polynomial equations of lower degree but with coefficients that are algebraic integers. For example, Gauss regarded the coordinates of the 17 vertices of a regular 17-sided figure as complex numbers satisfying the equation x17 − 1 = 0 and thus as algebraic integers. One such integer is 1. He showed that the rest are obtained by solving a succession of four quadratic equations. Because solving a quadratic equation is equivalent to performing a construction with a ruler and compass, as Descartes had shown long before, Gauss had shown how to construct the regular 17-gon.

      Inspired by Gauss's works on the theory of numbers, a growing school of mathematicians was drawn to the subject. Like Gauss, the German mathematician Ernst Eduard Kummer (Kummer, Ernst Eduard) sought to generalize the law of quadratic reciprocity to deal with questions about third, fourth, and higher powers of numbers. He found that his work led him in an unexpected direction, toward a partial resolution of Fermat's last theorem. In 1637 Fermat wrote in the margin of his copy of Diophantus's Arithmetica the claim to have a proof that there are no solutions in positive integers to the equation xn + yn = zn if n > 2. However, no proof was ever discovered among his notebooks.

      Kummer's approach was to develop the theory of algebraic integers. If it could be shown that the equation had no solution in suitable algebraic integers, then a fortiori there could be no solution in ordinary integers. He was eventually able to establish the truth of Fermat's last theorem for a large class of prime exponents n (those satisfying some technical conditions needed to make the proof work). This was the first significant breakthrough in the study of the theorem. Together with the earlier work of the French mathematician Sophie Germain (Germain, Sophie), it enabled mathematicians to establish Fermat's last theorem for every value of n from 3 to 4,000,000. However, Kummer's way around the difficulties he encountered further propelled the theory of algebraic integers into the realm of abstraction. It amounted to the suggestion that there should be yet other types of integers, but many found these ideas obscure.

      In Germany Richard Dedekind (Dedekind, Richard) patiently created a new approach, in which each new number (called an ideal) was defined by means of a suitable set of algebraic integers in such a way that it was the common divisor of the set of algebraic integers used to define it. Dedekind's work was slow to gain approval, yet it illustrates several of the most profound features of modern mathematics. It was clear to Dedekind that the ideal algebraic integers were the work of the human mind. Their existence can be neither based on nor deduced from the existence of physical objects, analogies with natural processes, or some process of abstraction from more familiar things. A second feature of Dedekind's work was its reliance on the idea of sets of objects, such as sets of numbers, even sets of sets. Dedekind's work showed how basic the naive conception of a set could be. The third crucial feature of his work was its emphasis on the structural aspects of algebra. The presentation of number theory as a theory about objects that can be manipulated (in this case, added and multiplied) according to certain rules akin to those governing ordinary numbers was to be a paradigm of the more formal theories of the 20th century.

The theory of equations
      Another subject that was transformed in the 19th century was the theory of equations. Ever since Tartaglia and Ferrari in the 16th century had found rules giving the solutions of cubic and quartic equations in terms of the coefficients of the equations, formulas had unsuccessfully been sought for equations of the fifth and higher degrees. At stake was the existence of a formula that expresses the roots of a quintic equation in terms of the coefficients. This formula, moreover, must involve only the operations of addition, subtraction, multiplication, and division, together with the extraction of roots, since that was all that had been required for the solution of quadratic, cubic, and quartic equations. If such a formula were to exist, the quintic would accordingly be said to be solvable by radicals.

      In 1770 Lagrange had analyzed all the successful methods he knew for second-, third-, and fourth-degree equations in an attempt to see why they worked and how they could be generalized. His analysis of the problem in terms of permutations of the roots was promising, but he became more and more doubtful as the years went by that his complicated line of attack could be carried through. The first valid proof that the general quintic is not solvable by radicals was offered only after his death, in a startlingly short paper by Niels Henrik Abel (Abel, Niels Henrik), written in 1824.

      Abel also showed by example that some quintic equations were solvable by radicals and that some equations could be solved unexpectedly easily. For example, the equation x5 − 1 = 0 has one root x = 1, but the remaining four roots can be found just by extracting square roots, not fourth roots as might be expected. He therefore raised the question “What equations of degree higher than four are solvable by radicals?”

      Abel died in 1829 at the age of 26 and did not resolve the problem he had posed. Almost at once, however, the astonishing prodigy Évariste Galois (Galois, Évariste) burst upon the Parisian mathematical scene. He submitted an account of his novel theory of equations to the Academy of Sciences in 1829, but the manuscript was lost. A second version was also lost and was not found among Fourier's papers when Fourier, the secretary of the academy, died in 1830. Galois was killed in a duel in 1832, at the age of 20, and it was not until his papers were published in Joseph Liouville (Liouville, Joseph)'s Journal de mathématiques in 1846 that his work began to receive the attention it deserved. His theory eventually made the theory of equations into a mere part of the theory of groups (group). Galois emphasized the group (as he called it) of permutations (permutations and combinations) of the roots of an equation. This move took him away from the equations themselves and turned him instead toward the markedly more tractable study of permutations. To any given equation there corresponds a definite group, with a definite collection of subgroups. To explain which equations were solvable by radicals and which were not, Galois analyzed the ways in which these subgroups were related to one another: solvable equations gave rise to what are now called a chain of normal subgroups with cyclic quotients. This technical condition makes it clear how far mathematicians had gone from the familiar questions of 18th-century mathematics, and it marks a transition characteristic of modern mathematics: the replacement of formal calculation by conceptual analysis. This is a luxury available to the pure mathematician that the applied mathematician faced with a concrete problem cannot always afford.

      According to this theory, a group is a set of objects that one can combine in pairs in such a way that the resulting object is also in the set. Moreover, this way of combination has to obey the following rules (here objects in the group are denoted a, b, etc., and the combination of a and b is written a * b):
● There is an element e such that a * e = a = e * a for every element a in the group. This element is called the identity element of the group.
● For every element a there is an element, written a−1, with the property that a * a−1 = e = a−1 * a. The element a−1 is called the inverse of a.
● For every a, b, and c in the group the associative law holds: (a * b) * c = a * (b * c).

      Examples of groups include the integers with * interpreted as addition and the positive rational numbers with * interpreted as multiplication. An important property shared by some groups but not all is commutativity: for every element a and b, a * b = b * a. The rotations of an object in the plane around a fixed point form a commutative group, but the rotations of a three-dimensional object around a fixed point form a noncommutative group.

      A convenient way to assess the situation in mathematics in the mid-19th century is to look at the career of its greatest exponent, Carl Friedrich Gauss (Gauss, Carl Friedrich), the last man to be called the “Prince of Mathematics.” In 1801, the same year in which he published his Disquisitiones Arithmeticae, he rediscovered the asteroid Ceres (which had disappeared behind the Sun not long after it was first discovered and before its orbit was precisely known). He was the first to give a sound analysis of the method of least squares (least squares approximation) in the analysis of statistical data. Gauss did important work in potential theory and, with the German physicist Wilhelm Weber (Weber, Wilhelm Eduard), built the first electric telegraph. He helped conduct the first survey of the Earth's magnetic field and did both theoretical and field work in cartography and surveying. He was a polymath who almost single-handedly embraced what elsewhere was being put asunder: the world of science and the world of mathematics. It is his purely mathematical work, however, that in its day was—and ever since has been—regarded as the best evidence of his genius.

 Gauss's writings transformed the theory of numbers. His theory of algebraic integers lay close to the theory of equations as Galois was to redefine it. More remarkable are his extensive writings, dating from 1797 to the 1820s but unpublished at his death, on the theory of elliptic functions. In 1827 he published his crucial discovery that the curvature of a surface can be defined intrinsically—that is, solely in terms of properties defined within the surface and without reference to the surrounding Euclidean space (see figure—>). This result was to be decisive in the acceptance of non-Euclidean geometry. All of Gauss's work displays a sharp concern for rigour and a refusal to rely on intuition or physical analogy, which was to serve as an inspiration to his successors. His emphasis on achieving full conceptual understanding, which may have led to his dislike of publication, was by no means the least influential of his achievements.

Non-Euclidean geometry
      Perhaps it was this desire for conceptual understanding that made Gauss reluctant to publish the fact that he was led more and more “to doubt the truth of geometry,” as he put it. For if there was a logically consistent geometry differing from Euclid's only because it made a different assumption about the behaviour of parallel lines, it too could apply to physical space, and so the truth of (Euclidean) geometry could no longer be assured a priori, as Kant had thought.

      Gauss's investigations into the new geometry went further than any one else's before him, but he did not publish them. The honour of being the first to proclaim the existence of a new geometry belongs to two others, who did so in the late 1820s: Nicolay Ivanovich Lobachevsky (Lobachevsky, Nikolay Ivanovich) in Russia and János Bolyai (Bolyai, János) in Hungary. Because the similarities in the work of these two men far exceed the differences, it is convenient to describe their work together.

 Both men made an assumption about parallel lines that differed from Euclid's and proceeded to draw out its consequences. This way of working cannot guarantee the consistency of one's findings, so, strictly speaking, they could not prove the existence of a new geometry in this way. Both men described a three-dimensional space different from Euclidean space by couching their findings in the language of trigonometry. The formulas they obtained were exact analogs of the formulas that describe triangles drawn on the surface of a sphere, with the usual trigonometric functions replaced by those of hyperbolic trigonometry (hyperbolic functions). The functions hyperbolic cosine, written cosh, and hyperbolic (hyperbolic geometry) sine, written sinh (see the figure—>), are defined as follows: cosh x = (ex + ex)/2, and sinh x = (ex − ex)/2. They are called hyperbolic because of their use in describing the hyperbola. Their names derive from the evident analogy with the trigonometric functions, which Euler showed satisfy these equations: cos x = (eix + eix)/2, and sin x = (eix − eix)/2i. The formulas were what gave the work of Lobachevsky and of Bolyai the precision needed to give conviction in the absence of a sound logical structure. Both men observed that it had become an empirical matter to determine the nature of space, Lobachevsky even going so far as to conduct astronomical observations, although these proved inconclusive.

      The work of Bolyai and of Lobachevsky was poorly received. Gauss endorsed what they had done, but so discreetly that most mathematicians did not find out his true opinion on the subject until he was dead. The main obstacle each man faced was surely the shocking nature of their discovery. It was easier, and in keeping with 2,000 years of tradition, to continue to believe that Euclidean geometry was correct and that Bolyai and Lobachevsky had somewhere gone astray, like many an investigator before them.

 The turn toward acceptance came in the 1860s, after Bolyai and Lobachevsky had died. The Italian mathematician Eugenio Beltrami (Beltrami, Eugenio) decided to investigate Lobachevsky's work and to place it, if possible, within the context of differential geometry as redefined by Gauss. He therefore moved independently in the direction already taken by Bernhard Riemann (Riemann, Bernhard). Beltrami investigated the surface of constant negative curvature (see the figure—>) and found that on such a surface triangles obeyed the formulas of hyperbolic trigonometry that Lobachevsky had discovered were appropriate to his form of non-Euclidean geometry. Thus, Beltrami gave the first rigorous description of a geometry other than Euclid's. Beltrami's account of the surface of constant negative curvature was ingenious. He said it was an abstract surface that he could describe by drawing maps of it, much as one might describe a sphere by means of the pages of a geographic atlas. He did not claim to have constructed the surface embedded in Euclidean two-dimensional space; David Hilbert (Hilbert, David) later showed that it cannot be done.

      When Gauss died in 1855, his post at Göttingen was taken by Peter Gustav Lejeune Dirichlet (Dirichlet, Peter Gustav Lejeune). One mathematician who found the presence of Dirichlet a stimulus to research was Bernhard Riemann (Riemann, Bernhard), and his few short contributions to mathematics were among the most influential of the century. Riemann's first paper, his doctoral thesis (1851) on the theory of complex functions, provided the foundations for a geometric treatment of functions of a complex variable. His main result guaranteed the existence of a wide class of complex functions satisfying only modest general requirements and so made it clear that complex functions could be expected to occur widely in mathematics. More important, Riemann achieved this result by yoking together the theory of complex functions with the theory of harmonic functions (harmonic function) and with potential theory. The theories of complex and harmonic functions were henceforth inseparable.

      Riemann then wrote on the theory of Fourier series and their integrability (integration). His paper was directly in the tradition that ran from Cauchy and Fourier to Dirichlet, and it marked a considerable step forward in the precision with which the concept of integral can be defined. In 1854 he took up a subject that much interested Gauss, the hypotheses lying at the basis of geometry.

      The study of geometry has always been one of the central concerns of mathematicians. It was the language, and the principal subject matter, of Greek mathematics, was the mainstay of elementary education in the subject, and has an obvious visual appeal. It seems easy to apply, for one can proceed from a base of naively intelligible concepts. In keeping with the general trends of the century, however, it was just the naive concepts that Riemann chose to refine. What he proposed as the basis of geometry was far more radical and fundamental than anything that had gone before.

      Riemann took his inspiration from Gauss's discovery that the curvature of a surface is intrinsic, and he argued that one should therefore ignore Euclidean space and treat each surface by itself. A geometric property, he argued, was one that was intrinsic to the surface. To do geometry, it was enough to be given a set of points and a way of measuring lengths along curves in the surface. For this, traditional ways of applying the calculus to the study of curves could be made to suffice. But Riemann (Riemannian geometry) did not stop with surfaces. He proposed that geometers study spaces of any dimension in this spirit—even, he said, spaces of infinite dimension.

      Several profound consequences followed from this view. It dethroned Euclidean geometry, which now became just one of many geometries. It allowed the geometry of Bolyai and Lobachevsky to be recognized as the geometry of a surface of constant negative curvature, thus resolving doubts about the logical consistency of their work. It highlighted the importance of intrinsic concepts in geometry. It helped open the way to the study of spaces of many dimensions. Last but not least, Riemann's work ensured that any investigation of the geometric nature of physical space would thereafter have to be partly empirical. One could no longer say that physical space is Euclidean because there is no geometry but Euclid's. This realization finally destroyed any hope that questions about the world could be answered by a priori reasoning.

      In 1857 Riemann published several papers applying his very general methods for the study of complex functions to various parts of mathematics. One of these papers solved the outstanding problem of extending the theory of elliptic functions to the integration of any algebraic function. It opened up the theory of complex functions of several variables and showed how Riemann's novel topological ideas were essential in the study of complex functions. (In subsequent lectures Riemann showed how the special case of the theory of elliptic functions could be regarded as the study of complex functions on a torus.)

      In another paper Riemann dealt with the question of how many prime numbers are less than any given number x. The answer is a function of x, and Gauss had conjectured on the basis of extensive numerical evidence that this function was approximately x/ln(x). This turned out to be true, but it was not proved until 1896, when both Charles-Jean de la Vallée Poussin of Belgium and Jacques-Salomon Hadamard (Hadamard, Jacques-Salomon) of France independently proved it. It is remarkable that a question about integers led to a discussion of functions of a complex variable, but similar connections had previously been made by Dirichlet. Riemann took the expression Π(1 − ps)−1 = Σns, introduced by Euler the century before, where the infinite product is taken over all prime numbers p and the sum over all whole numbers n, and treated it as a function of s. The infinite sum makes sense whenever s is real and greater than 1. Riemann proceeded to study this function when s is complex (now called the Riemann zeta function), and he thereby not only helped clarify the question of the distribution of primes but also was led to several other remarks that later mathematicians were to find of exceptional interest. One remark has continued to elude proof and remains one of the greatest conjectures in mathematics: the claim that the nonreal zeros of the zeta function are complex numbers whose real part is always equal to 1/2.

Riemann's influence
      In 1859 Dirichlet died and Riemann became a full professor, but he was already ill with tuberculosis, and in 1862 his health broke. He died in 1866. His work, however, exercised a growing influence on his successors. His work on trigonometric series, for example, led to a deepening investigation of the question of when a function is integrable. Attention was concentrated on the nature of the sets of points at which functions and their integrals (when these existed) had unexpected properties. The conclusions that emerged were at first obscure, but it became clear that some properties of point sets were important in the theory of integration, while others were not. (These other properties proved to be a vital part of the emerging subject of topology.) The properties of point sets that matter in integration have to do with the size of the set. If one can change the values of a function on a set of points without changing its integral, it is said that the set is of negligible size. The naive idea is that integrating is a generalization of counting: negligible sets do not need to be counted. About the turn of the century the French mathematician Henri-Léon Lebesgue (Lebesgue, Henri-Léon) managed to systematize this naive idea into a new theory about the size of sets, which included integration as a special case. In this theory, called measure theory, there are sets that can be measured, and they either have positive measure or are negligible (they have zero measure), and there are sets that cannot be measured at all.

 The first success for Lebesgue's (Lebesgue integral) theory was that, unlike the Cauchy-Riemann integral, it obeyed the rule that, if a sequence of functions fn(x) tends suitably to a function f(x), then the sequence of integrals ∫fn(x)dx tends to the integral ∫f(x)dx. This has made it the natural theory of the integral when dealing with questions about trigonometric series. (See the figure—>.) Another advantage is that it is very general. For example, in probability theory it is desirable to estimate the likelihood of certain outcomes of an experiment. By imposing a measure on the space of all possible outcomes, the Russian mathematician Andrey Kolmogorov (Kolmogorov, Andrey Nikolayevich) was the first to put probability theory on a rigorous mathematical footing.

      Another example is provided by a remarkable result discovered by the 20th-century American mathematician Norbert Wiener (Wiener, Norbert): within the set of all continuous functions on an interval, the set of differentiable functions has measure zero. In probabilistic terms, therefore, the chance that a function taken at random is differentiable has probability zero. In physical terms, this means that, for example, a particle moving under Brownian motion almost certainly is moving on a nondifferentiable path. This discovery clarified Albert Einstein (Einstein, Albert)'s fundamental ideas about Brownian motion (displayed by the continual motion of specks of dust in a fluid under the constant bombardment of surrounding molecules). The hope of physicists is that Richard Feynman (Feynman, Richard P.)'s theory of quantum electrodynamics will yield to a similar measure-theoretic treatment, for it has the disturbing aspect of a theory that has not been made rigorous mathematically but that accords excellently with observation.

      Yet another setting for Lebesgue's ideas was to be the theory of Lie groups. The Hungarian mathematician Alfréd Haar showed how to define the concept of measure so that functions defined on Lie groups could be integrated. This became a crucial part of Hermann Weyl (Weyl, Hermann)'s way of representing a Lie group as acting linearly on the space of all (suitable) functions on the group (for technical reasons, suitable means that the square of the function is integrable with respect to a Haar measure on the group).

Differential equations (differential equation)
      Another field that developed considerably in the 19th century was the theory of differential equations. The pioneer in this direction once again was Cauchy. Above all, he insisted that one should prove that solutions do indeed exist; it is not a priori obvious that every ordinary differential equation has solutions. The methods that Cauchy (Cauchy, Augustin-Louis, Baron) proposed for these problems fitted naturally into his program of providing rigorous foundations for all the calculus. The solution method he preferred, although the less-general of his two approaches, worked equally well in the real and complex cases. It established the existence of a solution equal to the one obtainable by traditional power series methods using newly developed techniques in his theory of functions of a complex variable.

      The harder part of the theory of differential equations concerns partial differential equations (partial differential equation), those for which the unknown function is a function of several variables. In the early 19th century there was no known method of proving that a given second- or higher-order partial differential equation had a solution, and there was not even a method of writing down a plausible candidate. In this case progress was to be much less marked. Cauchy found new and more rigorous methods for first-order partial differential equations, but the general case eluded treatment.

      An important special case was successfully prosecuted, that of dynamics. Dynamics is the study of the motion of a physical system under the action of forces. Working independently of each other, William Rowan Hamilton (Hamilton, Sir William Rowan) in Ireland and Carl Jacobi (Jacobi, Carl) in Germany showed how problems in dynamics could be reduced to systems of first-order partial differential equations. From this base grew an extensive study of certain partial differential operators. These are straightforward generalizations of a single partial differentiation (∂/∂x) to a sum of the form

      where the a's are functions of the x's. The effect of performing several of these in succession can be complicated, but Jacobi and the other pioneers in this field found that there are formal rules which such operators tend to satisfy. This enabled them to shift attention to these formal rules, and gradually an algebraic analysis of this branch of mathematics began to emerge.

      The most influential worker in this direction was the Norwegian Sophus Lie (Lie, Sophus). Lie, and independently Wilhelm Killing in Germany, came to suspect that the systems of partial differential operators they were studying came in a limited variety of types. Once the number of independent variables was specified (which fixed the dimension of the system), a large class of examples, including many of considerable geometric significance, seemed to fall into a small number of patterns. This suggested that the systems could be classified, and such a prospect naturally excited mathematicians. After much work by Lie and by Killing and later by the French mathematician Élie-Joseph Cartan (Cartan, Élie-Joseph), they were classified. Initially, this discovery aroused interest because it produced order where previously the complexity had threatened chaos and because it could be made to make sense geometrically. The realization that there were to be major implications of this work for the study of physics lay well in the future.

Linear algebra (algebra, linear)
      Differential equations, whether ordinary or partial, may profitably be classified as linear or nonlinear; linear differential equations are those for which the sum of two solutions is again a solution. The equation giving the shape of a vibrating string is linear, which provides the mathematical reason why a string may simultaneously emit more than one frequency. The linearity of an equation makes it easy to find all its solutions, so in general linear problems have been tackled successfully, while nonlinear equations continue to be difficult. Indeed, in many linear problems there can be found a finite family of solutions with the property that any solution is a sum of them (suitably multiplied by arbitrary constants). Obtaining such a family, called a basis, and putting them into their simplest and most useful form, was an important source of many techniques in the field of linear algebra.

      Consider, for example, the system of linear differential equations

 It is evidently much more difficult to study than the system dy1/dx = αy1, dy2/dx = βy2, whose solutions are (constant multiples of) y1 = exp (αx) and y2 = exp (βx). But if a suitable linear combination of y1 and y2 can be found so that the first system reduces to the second, then it is enough to solve the second system. The existence of such a reduction is determined by an array (called a matrix) of the four numbers (see the table—> of matrix operation rules). In 1858 the English mathematician Arthur Cayley (Cayley, Arthur) began the study of matrices in their own right when he noticed that they satisfy polynomial equations. The matrix
for example, satisfies the equation A2 − (a + d)A + (ad − bc) = 0. Moreover, if this equation has two distinct roots (root)—say, α and β—then the sought-for reduction will exist, and the coefficients of the simpler system will indeed be those roots α and β. If the equation has a repeated root, then the reduction usually cannot be carried out. In either case the difficult part of solving the original differential equation has been reduced to elementary algebra.

      The study of linear algebra begun by Cayley and continued by Leopold Kronecker (Kronecker, Leopold) includes a powerful theory of vector spaces (vector space). These are sets whose elements can be added together and multiplied by arbitrary numbers, such as the family of solutions of a linear differential equation. A more familiar example is that of three-dimensional space. If one picks an origin, then every point in space can be labeled by the line segment (called a vector) joining it to the origin. Matrices appear as ways of representing linear transformations of a vector space—i.e., transformations that preserve sums and multiplication by numbers: the transformation T is linear if, for any vectors u, v, T(u + v) = T(u) + T(v) and, for any scalar λ, T(λv) = λT(v). When the vector space is finite-dimensional, linear algebra and geometry form a potent combination. Vector spaces of infinite dimensions also are studied.

      The theory of vector spaces is useful in other ways. Vectors in three-dimensional space represent such physically important concepts as velocities (velocity) and forces (force). Such an assignment of vector to point is called a vector field; examples include electric and magnetic fields. Scientists such as James Clerk Maxwell (Maxwell, James Clerk) and J. Willard Gibbs (Gibbs, J Willard) took up vector analysis and were able to extend vector methods to the calculus. They introduced in this way measures of how a vector field varies infinitesimally, which, under the names div, grad, and curl, have become the standard tools in the study of electromagnetism and potential theory. To the modern mathematician, div, grad, and curl form part of a theory to which Stokes's law (a special case of which is Green's theorem) is central. The Gauss-Green-Stokes theorem, named after Gauss and two leading English applied mathematicians of the 19th century (George Stokes (Stokes, Sir George Gabriel, 1st Baronet) and George Green (Green, George)), generalizes the fundamental theorem of the calculus to functions of several variables. The fundamental theorem of calculus asserts that

      which can be read as saying that the integral of the derivative of some function in an interval is equal to the difference in the values of the function at the endpoints of the interval. Generalized to a part of a surface or space, this asserts that the integral of the derivative of some function over a region is equal to the integral of the function over the boundary of the region. In symbols this says that ∫dω = ∫ω, where the first integral is taken over the region in question and the second integral over its boundary, while dω is the derivative of ω.

The foundations of geometry
      By the late 19th century the hegemony of Euclidean geometry had been challenged by non-Euclidean geometry and projective geometry. The first notable attempt to reorganize the study of geometry was made by the German mathematician Felix Klein (Klein, Felix) and published at Erlangen in 1872. In his Erlanger Programm Klein proposed that Euclidean and non-Euclidean geometry be regarded as special cases of projective geometry. In each case the common features that, in Klein's opinion, made them geometries were that there were a set of points, called a “space,” and a group of transformations by means of which figures could be moved around in the space without altering their essential properties. For example, in Euclidean plane geometry the space is the familiar plane, and the transformations are rotations, reflections, translations, and their composites, none of which change either length or angle, the basic properties of figures in Euclidean geometry. Different geometries would have different spaces and different groups, and the figures would have different basic properties.

      Klein produced an account that unified a large class of geometries—roughly speaking, all those that were homogeneous in the sense that every piece of the space looked like every other piece of the space. This excluded, for example, geometries on surfaces of variable curvature, but it produced an attractive package for the rest and gratified the intuition of those who felt that somehow projective geometry was basic. It continued to look like the right approach when Lie's ideas appeared, and there seemed to be a good connection between Lie's classification and the types of geometry organized by Klein.

      Mathematicians could now ask why they had believed Euclidean geometry to be the only one when, in fact, many different geometries existed. The first to take up this question successfully was the German mathematician Moritz Pasch, who argued in 1882 that the mistake had been to rely too heavily on physical intuition. In his view an argument in mathematics should depend for its validity not on the physical interpretation of the terms involved but upon purely formal (formalism) criteria. Indeed, the principle of duality did violence to the sense of geometry as a formalization of what one believed about (physical) points and lines; one did not believe that these terms were interchangeable.

      The ideas of Pasch caught the attention of the German mathematician David Hilbert (Hilbert, David), who, with the French mathematician Henri Poincaré (Poincaré, Henri), came to dominate mathematics at the beginning of the 20th century. In wondering why it was that mathematics—and in particular geometry—produced correct results, he came to feel increasingly that it was not because of the lucidity of its definitions. Rather, mathematics worked because its (elementary) terms were meaningless. What kept it heading in the right direction was its rules of inference. Proofs (proof) were valid because they were constructed through the application of the rules of inference, according to which new assertions could be declared to be true simply because they could be derived, by means of these rules, from the axioms (axiom) or previously proven theorems. The theorems and axioms were viewed as formal statements that expressed the relationships between these terms.

      The rules governing the use of mathematical terms were arbitrary, Hilbert argued, and each mathematician could choose them at will, provided only that the choices made were self-consistent. A mathematician produced abstract systems unconstrained by the needs of science, and, if scientists found an abstract system that fit one of their concerns, they could apply the system secure in the knowledge that it was logically consistent.

      Hilbert first became excited about this point of view (presented in his Grundlagen der Geometrie [1899; Foundations of Geometry) when he saw that it led not merely to a clear way of sorting out the geometries in Klein's hierarchy according to the different axiom systems they obeyed but to new geometries as well. For the first time there was a way of discussing geometry that lay beyond even the very general terms proposed by Riemann. Not all of these geometries have continued to be of interest, but the general moral that Hilbert first drew for geometry he was shortly to draw for the whole of mathematics.

The foundations of mathematics
      By the late 19th century the debates about the foundations of geometry had become the focus for a running debate about the nature of the branches of mathematics. Cauchy's work on the foundations of the calculus, completed by the German mathematician Karl Weierstrass (Weierstrass, Karl) in the late 1870s, left an edifice that rested on concepts such as that of the natural numbers (the integers 1, 2, 3, and so on) and on certain constructions involving them. The algebraic theory of numbers and the transformed theory of equations had focused attention on abstract structures in mathematics. Questions that had been raised about numbers since Babylonian times turned out to be best cast theoretically in terms of entirely modern creations whose independence from the physical world was beyond dispute. Finally, geometry, far from being a kind of abstract physics, was now seen as dealing with meaningless terms obeying arbitrary systems of rules. Although there had been no conscious plan leading in that direction, the stage was set for a consideration of questions about the fundamental nature of mathematics.

      Similar currents were at work in the study of logic, which had also enjoyed a revival during the 19th century. The work of the English mathematician George Boole (Boole, George) and the American Charles Sanders Peirce (Peirce, Charles Sanders) had contributed to the development of a symbolism adequate to explore all elementary logical deductions. Significantly, Boole's book on the subject was called An Investigation of the Laws of Thought, on Which Are Founded the Mathematical Theories of Logic and Probabilities (1854). In Germany the logician Gottlob Frege (Frege, Gottlob) had directed keen attention to such fundamental questions as what it means to define something and what sorts of purported definitions actually do define.

Cantor
      All of these debates came together through the pioneering work of the German mathematician Georg Cantor (Cantor, Georg) on the concept of a set. Cantor had begun work in this area because of his interest in Riemann's theory of trigonometric series, but the problem of what characterized the set of all real numbers came to occupy him more and more. He began to discover unexpected properties of sets. For example, he could show that the set of all algebraic numbers (algebraic number), and a fortiori the set of all rational numbers, is countable in the sense that there is a one-to-one correspondence between the integers and the members of each of these sets by means of which for any member of the set of algebraic numbers (or rationals), no matter how large, there is always a unique integer it may be placed in correspondence with. But, more surprisingly, he could also show that the set of all real numbers (real number) is not countable. So, although the set of all integers and the set of all real numbers are both infinite, the set of all real numbers is a strictly larger infinity. This was in complete contrast to the prevailing orthodoxy, which proclaimed that infinite could mean only “larger than any finite amount.”

      Here the concept of number was being extended and undermined at the same time. The concept was extended because it was now possible to count and order sets that the set of integers was too small to measure, and it was undermined because even the integers ceased to be basic undefined objects. Cantor himself had given a way of defining real numbers as certain infinite sets of rational numbers. Rational numbers were easy to define in terms of the integers, but now integers could be defined by means of sets. One way was given by Frege in Die Grundlagen der Arithmetik (1884; The Foundations of Arithmetic). He regarded two sets as the same if they contained the same elements. So in his opinion there was only one empty set (today symbolized by Ø), the set with no members. A second set could be defined as having only one element by letting that element be the empty set itself (symbolized by {Ø}), a set with two elements by letting them be the two sets just defined (i.e., {Ø, {Ø}}), and so on. Having thus defined the integers in terms of the primitive concepts “set” and “element of,” Frege agreed with Cantor that there was no logical reason to stop, and he went on to define infinite sets in the same way Cantor had. Indeed, Frege was clearer than Cantor about what sets and their elements actually were.

      Frege's proposals went in the direction of a reduction of all mathematics to logic. He hoped that every mathematical term could be defined precisely and manipulated according to agreed, logical rules of inference. This, the “logicist (logicism)” program, was dealt an unexpected blow in 1902 by the English mathematician and philosopher Bertrand Russell (Russell, Bertrand), who pointed out unexpected complications with the naive concept of a set. Nothing seemed to preclude the possibility that some sets were elements of themselves while others were not, but, asked Russell, “What then of the set of all sets that were not elements of themselves?” If it is an element of itself, then it is not (an element of itself), but, if it is not, then it is. Russell had identified a fundamental problem in set theory with his paradox (Russell's paradox). Either the idea of a set as an arbitrary collection of already defined objects was flawed, or else the idea that one could legitimately form the set of all sets of a given kind was incorrect. Frege's program never recovered from this blow, and Russell's similar approach of defining mathematics in terms of logic, which he developed together with Alfred North Whitehead (Whitehead, Alfred North) in their Principia Mathematica (1910–13), never found lasting appeal with mathematicians.

      Greater interest attached to the ideas that Hilbert and his school began to advance. It seemed to them that what had worked once for geometry could work again for all of mathematics. Rather than attempt to define things so that problems could not arise, they suggested that it was possible to dispense with definitions and cast all of mathematics in an axiomatic (axiomatic method) structure using the ideas of set theory. Indeed, the hope was that the study of logic could be embraced in this spirit, thus making logic a branch of mathematics, the opposite of Frege's intention. There was considerable progress in this direction, and there emerged both a powerful school of mathematical logicians (notably in Poland) and an axiomatic theory of sets that avoided Russell's paradoxes and the others that had sprung up.

      In the 1920s Hilbert put forward his most detailed proposal for establishing the validity of mathematics. According to his theory of proofs, everything was to be put into an axiomatic form, allowing the rules of inference to be only those of elementary logic, and only those conclusions that could be reached from this finite set of axioms and rules of inference were to be admitted. He proposed that a satisfactory system would be one that was consistent, complete, and decidable. By “consistent” Hilbert meant that it should be impossible to derive both a statement and its negation; by “complete,” that every properly written statement should be such that either it or its negation was derivable from the axioms; by “decidable,” that one should have an algorithm that determines of any given statement whether it or its negation is provable. Such systems did exist—for example, the first-order predicate calculus—but none had been found capable of allowing mathematicians to do interesting mathematics.

      Hilbert's program, however, did not last long. In 1931 the Austrian-born American mathematician and logician Kurt Gödel (Gödel, Kurt) showed that there was no system of Hilbert's type within which the integers could be defined and that was both consistent and complete. Gödel and, independently, the English mathematician Alan Turing (Turing, Alan M.) later showed that decidability was also unattainable. Perhaps paradoxically, the effect of this dramatic discovery was to alienate mathematicians from the whole debate. Instead, mathematicians, who may not have been too unhappy with the idea that there is no way of deciding the truth of a proposition automatically, learned to live with the idea that not even mathematics rests on rigorous foundations. Progress since has been in other directions. An alternative axiom system for set theory was later put forward by the Hungarian-born American mathematician John von Neumann (von Neumann, John), which he hoped would help resolve contemporary problems in quantum mechanics. There was also a renewal of interest in statements that are both interesting mathematically and independent of the axiom system in use. The first of these was the American mathematician Paul Cohen (Cohen, Paul Joseph)'s surprising resolution in 1963 of the continuum hypothesis, which was Cantor's conjecture that the set of all subsets of the rational numbers was of the same size as the set of all real numbers. This turns out to be independent of the usual axioms for set theory, so there are set theories (and therefore types of mathematics) in which it is true and others in which it is false.

Mathematical physics
      At the same time that mathematicians were attempting to put their own house in order, they were also looking with renewed interest at contemporary work in physics. The man who did the most to rekindle their interest was Poincaré. Poincaré showed that dynamic systems described by quite simple differential equations, such as the solar system, can nonetheless yield the most random-looking, chaotic behaviour. He went on to explore ways in which mathematicians can nonetheless say things about this chaotic behaviour and so pioneered the way in which probabilistic statements about dynamic systems can be found to describe what otherwise defies intelligence.

      Poincaré later turned to problems of electrodynamics. After many years' work, the Dutch physicist Hendrik Antoon Lorentz (Lorentz, Hendrik Antoon) had been led to an apparent dependence of length and time on motion, and Poincaré was pleased to notice that the transformations (Lorentz transformations) that Lorentz proposed as a way of converting one observer's data into another's formed a group. This appealed to Poincaré and strengthened his belief that there was no sense in a concept of absolute motion; all motion was relative. Poincaré thereupon gave an elegant mathematical formulation of Lorentz's ideas, which fitted them into a theory in which the motion of the electron is governed by Maxwell's equations. Poincaré, however, stopped short of denying the reality of the ether or of proclaiming that the velocity of light (speed of light) is the same for all observers, so credit for the first truly relativistic theory of the motion of the electron rests with Einstein and his special theory of relativity (1905).

      Einstein's special theory is so called because it treats only the special case of uniform relative motion. The much more important case of accelerated motion and motion in a gravitational field was to take a further decade and to require a far more substantial dose of mathematics. Einstein changed his estimate of the value of pure mathematics, which he had hitherto disdained, only when he discovered that many of the questions he was led to had already been formulated mathematically and had been solved. He was most struck by theories derived from the study of geometry in the sense in which Riemann had formulated it.

      By 1915 a number of mathematicians were interested in reapplying their discoveries to physics. The leading institution in this respect was the University of Göttingen (Göttingen, University of), where Hilbert had unsuccessfully attempted to produce a general theory of relativity before Einstein, and it was there that many of the leaders of the coming revolution in quantum mechanics were to study. There too went many of the leading mathematicians of their generation, notably John von Neumann (von Neumann, John) and Hermann Weyl (Weyl, Hermann), to study with Hilbert. In 1904 Hilbert had turned to the study of integral equations (integral equation). These arise in many problems where the unknown is itself a function of some variable, and especially in those parts of physics that are expressed in terms of extremal principles (such as the principle of least action). The extremal principle usually yields information about an integral involving the sought-for function, hence the name integral equation. Hilbert's contribution was to bring together many different strands of contemporary work and to show how they could be elucidated if cast in the form of arguments about objects in certain infinite-dimensional vector spaces.

      The extension to infinite dimensions was not a trivial task, but it brought with it the opportunity to use geometric intuition and geometric concepts to analyze problems about integral equations. Hilbert left it to his students to provide the best abstract setting for his work, and thus was born the concept of a Hilbert space. Roughly, this is an infinite-dimensional vector space in which it makes sense to speak of the lengths of vectors and the angles between them; useful examples include certain spaces of sequences and certain spaces of functions. Operators defined on these spaces are also of great interest; their study forms part of the field of functional analysis (analysis).

      When in the 1920s mathematicians and physicists were seeking ways to formulate the new quantum mechanics, von Neumann proposed that the subject be written in the language of functional analysis. The quantum mechanical world of states and observables, with its mysterious wave packets that were sometimes like particles and sometimes like waves depending on how they were observed, went very neatly into the theory of Hilbert spaces. Functional analysis has ever since grown with the fortunes of particle physics.

      The early 20th century saw the emergence of a number of theories whose power and utility reside in large part in their generality. Typically, they are marked by an attention to the set or space of all examples of a particular kind. (Functional analysis is such an endeavour.) One of the most energetic of these general theories was that of algebraic topology. In this subject a variety of ways are developed for replacing a space by a group and a map between spaces by a map between groups. It is like using X-rays: information is lost, but the shadowy image of the original space may turn out to contain, in an accessible form, enough information to solve the question at hand.

 Interest in this kind of research came from various directions. Galois's theory of equations was an example of what could be achieved by transforming a problem in one branch of mathematics into a problem in another, more abstract branch. Another impetus came from Riemann's theory of complex functions. He had studied algebraic functions—that is, loci defined by equations of the form f(xy) = 0, where f is a polynomial in x whose coefficients are polynomials in y. When x and y are complex variables, the locus can be thought of as a real surface spread out over the x plane of complex numbers (today called a Riemann surface). To each value of x there correspond a finite number of values of y. Such surfaces are not easy to comprehend, and Riemann had proposed to draw curves along them in such a way that, if the surface was cut open along them, it could be opened out into a polygonal disk (see the figure—>). He was able to establish a profound connection between the minimum number of curves needed to do this for a given surface and the number of functions (becoming infinite at specified points) that the surface could then support.

 The natural problem was to see how far Riemann's ideas could be applied to the study of spaces of higher dimension. Here two lines of inquiry developed. One emphasized what could be obtained from looking at the projective geometry involved. This point of view was fruitfully applied by the Italian school of algebraic geometers. It ran into problems, which it was not wholly able to solve, having to do with the singularities a surface can possess. Whereas a locus given by f(xy) = 0 may intersect itself only at isolated points, a locus given by an equation of the form f(xyz) = 0 may intersect itself along curves (see figure—>), a problem that caused considerable difficulties. The second approach emphasized what can be learned from the study of integrals along paths on the surface. This approach, pursued by Charles-Émile Picard (Picard, Charles-Émile) and by Poincaré, provided a rich generalization of Riemann's original ideas.

      On this base, conjectures were made and a general theory produced, first by Poincaré and then by the American engineer-turned-mathematician Solomon Lefschetz, concerning the nature of manifolds (manifold) of arbitrary dimension. Roughly speaking, a manifold is the n-dimensional generalization of the idea of a surface; it is a space any small piece of which looks like a piece of n-dimensional space. Such an object is often given by a single algebraic equation in n + 1 variables. At first the work of Poincaré and of Lefschetz was concerned with how these manifolds may be decomposed into pieces, counting the number of pieces and decomposing them in their turn. The result was a list of numbers, called Betti numbers in honour of the Italian mathematician Enrico Betti (Betti, Enrico), who had taken the first steps of this kind to extend Riemann's work. It was only in the late 1920s that the German mathematician Emmy Noether (Noether, Emmy) suggested how the Betti numbers might be thought of as measuring the size of certain groups. At her instigation a number of people then produced a theory of these groups, the so-called homology and cohomology groups of a space.

      Two objects that can be deformed into one another will have the same homology and cohomology groups. To assess how much information is lost when a space is replaced by its algebraic topological picture, Poincaré asked the crucial converse question “According to what algebraic conditions is it possible to say that a space is topologically equivalent to a sphere?” He showed by an ingenious example that having the same homology is not enough and proposed a more delicate index, which has since grown into the branch of topology called homotopy theory. Being more delicate, it is both more basic and more difficult. There are usually standard methods for computing homology and cohomology groups, and they are completely known for many spaces. In contrast, there is scarcely an interesting class of spaces for which all the homotopy groups are known. Poincaré's conjecture that a space with the homotopy of a sphere actually is a sphere was shown to be true in the 1960s in dimensions five and above, and in the 1980s it was shown to be true for four-dimensional spaces. In 2006 Grigori Perelman (Perelman, Grigori) was awarded a Fields Medal for proving Poincaré's conjecture true in three dimensions, the only dimension in which Poincaré had studied it.

Developments in pure mathematics
      The interest in axiomatic systems at the turn of the century led to axiom systems for the known algebraic structures, that for the theory of fields, for example, being developed by the German mathematician Ernst Steinitz in 1910. The theory of rings (ring) (structures in which it is possible to add, subtract, and multiply but not necessarily divide) was much harder to formalize. It is important for two reasons: the theory of algebraic integers forms part of it, because algebraic integers naturally form into rings; and (as Kronecker and Hilbert had argued) algebraic geometry forms another part. The rings that arise there are rings of functions definable on the curve, surface, or manifold or are definable on specific pieces of it.

      Problems in number theory and algebraic geometry are often very difficult, and it was the hope of mathematicians such as Noether, who laboured to produce a formal, axiomatic theory of rings, that, by working at a more rarefied level, the essence of the concrete problems would remain while the distracting special features of any given case would fall away. This would make the formal theory both more general and easier, and to a surprising extent these mathematicians were successful.

      A further twist to the development came with the work of the American mathematician Oscar Zariski, who had studied with the Italian school of algebraic geometers but came to feel that their method of working was imprecise. He worked out a detailed program whereby every kind of geometric configuration could be redescribed in algebraic terms. His work succeeded in producing a rigorous theory, although some, notably Lefschetz, felt that the geometry had been lost sight of in the process.

      The study of algebraic geometry was amenable to the topological methods of Poincaré and Lefschetz so long as the manifolds were defined by equations whose coefficients were complex numbers (complex number). But, with the creation of an abstract theory of fields, it was natural to want a theory of varieties defined by equations with coefficients in an arbitrary field. This was provided for the first time by the French mathematician André Weil (Weil, André), in his Foundations of Algebraic Geometry (1946), in a way that drew on Zariski's work without suppressing the intuitive appeal of geometric concepts. Weil's theory of polynomial equations is the proper setting for any investigation that seeks to determine what properties of a geometric object can be derived solely by algebraic means. But it falls tantalizingly short of one topic of importance: the solution of polynomial equations in integers. This was the topic that Weil took up next.

      The central difficulty is that in a field it is possible to divide but in a ring it is not. The integers form a ring but not a field (dividing 1 by 2 does not yield an integer). But Weil showed that simplified versions (posed over a field) of any question about integer solutions to polynomials could be profitably asked. This transferred the questions to the domain of algebraic geometry. To count the number of solutions, Weil proposed that, since the questions were now geometric, they should be amenable to the techniques of algebraic topology. This was an audacious move, since there was no suitable theory of algebraic topology available, but Weil conjectured what results it should yield. The difficulty of Weil's conjectures may be judged by the fact that the last of them was a generalization to this setting of the famous Riemann hypothesis about the zeta function (Riemann zeta function), and they rapidly became the focus of international attention.

      Weil, along with Claude Chevalley, Henri Cartan (Cartan, Henri), Jean Dieudonné (Dieudonné, Jean), and others, created a group of young French mathematicians who began to publish virtually an encyclopaedia of mathematics under the name Nicolas Bourbaki (Bourbaki, Nicolas), taken by Weil from an obscure general of the Franco-German War. Bourbaki became a self-selecting group of young mathematicians who were strong on algebra, and the individual Bourbaki members were interested in the Weil conjectures. In the end they succeeded completely. A new kind of algebraic topology was developed, and the Weil conjectures were proved. The generalized Riemann hypothesis was the last to surrender, being established by the Belgian Pierre Deligne (Deligne, Pierre René) in the early 1970s. Strangely, its resolution still leaves the original Riemann hypothesis unsolved.

      Bourbaki was a key figure in the rethinking of structural mathematics. Algebraic topology was axiomatized by Samuel Eilenberg, a Polish-born American mathematician and Bourbaki member, and the American mathematician Norman Steenrod. Saunders Mac Lane (Mac Lane, Saunders), also of the United States, and Eilenberg extended this axiomatic approach until many types of mathematical structures were presented in families, called categories. Hence there was a category consisting of all groups and all maps between them that preserve multiplication, and there was another category of all topological spaces and all continuous maps between them. To do algebraic topology was to transfer a problem posed in one category (that of topological spaces) to another (usually that of commutative groups or rings). When he created the right algebraic topology for the Weil conjectures, the German-born French mathematician Alexandre Grothendieck (Grothendieck, Alexandre), a Bourbaki of enormous energy, produced a new description of algebraic geometry. In his hands it became infused with the language of category theory. The route to algebraic geometry became the steepest ever, but the views from the summit have a naturalness and a profundity that have brought many experts to prefer it to the earlier formulations, including Weil's.

      Grothendieck's formulation makes algebraic geometry the study of equations defined over rings rather than fields. Accordingly, it raises the possibility that questions about the integers can be answered directly. Building on the work of like-minded mathematicians in the United States, France, and Russia, the German Gerd Faltings (Faltings, Gerd) triumphantly vindicated this approach when he solved the Englishman Louis Mordell's conjecture in 1983. This conjecture states that almost all polynomial equations that define curves have at most finitely many rational solutions; the cases excluded from the conjecture are the simple ones that are much better understood.

      Meanwhile, Gerhard Frey of Germany had pointed out that, if Fermat's last theorem is false, so that there are integers u, v, w such that up + vp = wp (p greater than 5), then for these values of u, v, and p the curve y2 = x(x − up)(x + vp) has properties that contradict major conjectures of the Japanese mathematicians Taniyama Yutaka and Shimura Goro about elliptic curves. Frey's observation, refined by Jean-Pierre Serre (Serre, Jean-Pierre) of France and proved by the American Ken Ribet, meant that by 1990 Taniyama's unproven conjectures were known to imply Fermat's last theorem.

      In 1993 the English mathematician Andrew Wiles (Wiles, Andrew John) established the Shimura-Taniyama conjectures in a large range of cases that included Frey's curve and therefore Fermat's last theorem—a major feat even without the connection to Fermat. It soon became clear that the argument had a serious flaw; but in May 1995 Wiles, assisted by another English mathematician, Richard Taylor, published a different and valid approach. In so doing, Wiles not only solved the most famous outstanding conjecture in mathematics but also triumphantly vindicated the sophisticated and difficult methods of modern number theory.

Mathematical physics and the theory of groups
      In the 1910s the ideas of Lie and Killing were taken up by the French mathematician Élie-Joseph Cartan (Cartan, Élie-Joseph), who simplified their theory and rederived the classification of what came to be called the classical complex Lie algebras. The simple Lie algebras, out of which all the others in the classification are made, were all representable as algebras of matrices, and, in a sense, Lie algebra is the abstract setting for matrix algebra. Connected to each Lie algebra there were a small number of Lie groups, and there was a canonical simplest one to choose in each case. The groups had an even simpler geometric interpretation than the corresponding algebras, for they turned out to describe motions that leave certain properties of figures unaltered. For example, in Euclidean three-dimensional space, rotations leave unaltered the distances between points; the set of all rotations about a fixed point turns out to form a Lie group, and it is one of the Lie groups in the classification. The theory of Lie algebras and Lie groups shows that there are only a few sensible ways to measure properties of figures in a linear space and that these methods yield groups of motions leaving the figures, which are (more or less) groups of matrices, unaltered. The result is a powerful theory that could be expected to apply to a wide range of problems in geometry and physics.

      The leader in the endeavours to make Cartan's theory, which was confined to Lie algebras, yield results for a corresponding class of Lie groups was the German American Hermann Weyl (Weyl, Hermann). He produced a rich and satisfying theory for the pure mathematician and wrote extensively on differential geometry and group theory and its applications to physics. Weyl attempted to produce a theory that would unify gravitation and electromagnetism. His theory met with criticism from Einstein and was generally regarded as unsuccessful; only in the last quarter of the 20th century did similar unified field theories (unified field theory) meet with any acceptance. Nonetheless, Weyl's approach demonstrates how the theory of Lie groups can enter into physics in a substantial way.

      In any physical theory the endeavour is to make sense of observations. Different observers make different observations. If they differ in choice and direction of their coordinate axes, they give different coordinates to the same points, and so on. Yet the observers agree on certain consequences of their observations: in Newtonian physics and Euclidean geometry they agree on the distance between points. Special relativity explains how observers in a state of uniform relative motion differ about lengths and times but agree on a quantity called the interval. In each case they are able to do so because the relevant theory presents them with a group of transformations that converts one observer's measurements into another's and leaves the appropriate basic quantities invariant. What Weyl proposed was a group that would permit observers in nonuniform relative motion, and whose measurements of the same moving electron would differ, to convert their measurements and thus permit the (general) relativistic study of moving electric charges (electric charge).

      In the 1950s the American physicists Chen Ning Yang (Yang, Chen Ning) and Robert L. Mills (Mills, Robert) gave a successful treatment of the so-called strong interaction (strong force) in particle physics from the Lie group point of view. Twenty years later mathematicians took up their work, and a dramatic resurgence of interest in Weyl's theory began. These new developments, which had the incidental effect of enabling mathematicians to escape the problems in Weyl's original approach, were the outcome of lines of research that had originally been conducted with little regard for physical questions. Not for the first time, mathematics was to prove surprisingly effective—or, as the Hungarian-born American physicist Eugene Wigner (Wigner, Eugene Paul) said, “unreasonably effective”—in science.

 Cartan had investigated how much may be accomplished in differential geometry by using the idea of moving frames of reference (reference frame). This work, which was partly inspired by Einstein's theory of general relativity, was also a development of the ideas of Riemannian geometry that had originally so excited Einstein. In the modern theory one imagines a space (usually a manifold) made up of overlapping coordinatized pieces. On each piece one supposes some functions to be defined, which might in applications be the values of certain physical quantities. Rules are given for interpreting these quantities where the pieces overlap. The data are thought of as a bundle of information provided at each point. For each function defined on each patch, it is supposed that at each point a vector space is available as mathematical storage space for all its possible values. Because a vector space is attached at each point, the theory is called the theory of vector bundles. Other kinds of space may be attached, thus entering the more general theory of fibre bundles. The subtle and vital point is that it is possible to create quite different bundles which nonetheless look similar in small patches. (An example of this is illustrated in the figure—>.) The cylinder and the Möbius band (Möbius strip) look alike in small pieces but are topologically distinct, since it is possible to give a standard sense of direction to all the lines in the cylinder but not to those in the Möbius band. Both spaces can be thought of as one-dimensional vector bundles over the circle, but they are very different. The cylinder is regarded as a “trivial” bundle, the Möbius band as a twisted one.

      In the 1940s and '50s a vigorous branch of algebraic topology established the main features of the theory of bundles. Then, in the 1960s, work chiefly by Grothendieck and the English mathematician Michael Atiyah (Atiyah, Sir Michael Francis) showed how the study of vector bundles on spaces could be regarded as the study of cohomology theory (called K theory). More significantly still, in the 1960s Atiyah, the American Isadore Singer (Singer, Isadore Manuel), and others found ways of connecting this work to the study of a wide variety of questions involving partial differentiation, culminating in the celebrated Atiyah-Singer theorem for elliptic operators. (Elliptic is a technical term for the type of operator studied in potential theory.) There are remarkable implications for the study of pure geometry, and much attention has been directed to the problem of how the theory of bundles embraces the theory of Yang and Mills, which it does precisely because there are nontrivial bundles, and to the question of how it can be made to pay off in large areas of theoretical physics. These include the theories of superspace and supergravity and the string theory of fundamental particles, which involves the theory of Riemann surfaces in novel and unexpected ways.

Jeremy John Gray

Additional Reading

General sources
Two standard texts are Carl B. Boyer, A History of Mathematics, rev. by Uta C. Merzbach, 2nd ed. rev. (1989, reissued 1991); and, on a more elementary level, Howard Eves, An Introduction to the History of Mathematics, 6th ed. (1990). Discussions of the mathematics of various periods may be found in O. Neugebauer, The Exact Sciences in Antiquity, 2nd ed. (1957, reissued 1993); Morris Kline, Mathematical Thought from Ancient to Modern Times, 3 vol. (1972, reissued 1990); and B.L. van der Waerden, Science Awakening, trans. by Arnold Dresden, 4th ed. (1975, reissued 1988; originally published in Dutch, 1950). See also Kenneth O. May, Bibliography and Research Manual of the History of Mathematics (1973); and Joseph W. Dauben, The History of Mathematics from Antiquity to the Present: A Selective Bibliography (1985). A good source for biographies of mathematicians is Charles Coulston Gillispie (ed.), Dictionary of Scientific Biography, 16 vol. (1970–80, reissued 16 vol. in 8, 1981). Those wanting to study the writings of the mathematicians themselves will find the following sourcebooks useful: Henrietta O. Midonick (ed.), The Treasury of Mathematics: A Collection of Source Material in Mathematics, new ed. (1968); John Fauvel and Jeremy Gray (eds.), The History of Mathematics: A Reader (1987, reissued 1990); D.J. Struik (ed.), A Source Book in Mathematics, 1200–1800 (1969, reprinted 1986); and David Eugene Smith, A Source Book in Mathematics (1929; reissued in 2 vol., 1959). A study of the development of numeric notation can be found in Georges Ifrah, From One to Zero, trans. by Lowell Bair (1985; originally published in French, 1981).

Mathematics in ancient Mesopotamia
O. Neugebauer and A. Sachs, Mathematical Cuneiform Texts (1945, reissued 1986), is the principal English edition of mathematical tablets. A brief look at Babylonian mathematics is contained in the first chapter of Asger Aaboe, Episodes from the Early History of Mathematics (1964, reissued 1998), pp. 5–31.

Mathematics in ancient Egypt
Editions of the basic texts are T. Eric Peet (ed. and trans.), The Rhind Mathematical Papyrus: British Museum 10057 and 10058 (1923, reprinted 1970); and Arnold Buffam Chace and Henry Parker Manning (trans.), The Rhind Mathematical Papyrus, 2 vol. (1927–29, reprinted 2 vol. in 1, 1979). A brief but useful summary appears in G.J. Toomer, “Mathematics and Astronomy,” chapter 2 in J.R. Harris (ed.), The Legacy of Egypt, 2nd ed. (1971), pp. 27–54. For an extended account of Egyptian mathematics, see Richard J. Gillings, Mathematics in the Time of the Pharaohs (1972, reprinted 1982).

Greek mathematics
Critical editions of Greek mathematical texts include Dana Densmore (ed.), Euclid's Elements, trans. by Thomas L. Heath (2002; also published as The Thirteen Books of Euclid's Elements, 1926, reprinted 1956); Thomas L. Heath (ed. and trans.), The Works of Archimedes (1897, reissued 2002); E.J. Dijksterhuis, Archimedes, trans. by C. Dikshoorn (1956, reprinted 1987; originally published in Dutch, 1938); Thomas L. Heath, Apollonius of Perga: Treatise on Conic Sections (1896, reissued 1961), and Diophantus of Alexandria: A Study in the History of Greek Algebra, 2nd ed. (1910, reprinted 1964); and Jacques Sesiano, Books IV to VII of Diophantus' “Arithmetica” in the Arabic Translation Attributed to Qusṭā ibn Lū (1982). General surveys are Thomas L. Heath, A History of Greek Mathematics, 2 vol. (1921, reprinted 1993); Jacob Klein, Greek Mathematical Thought and the Origin of Algebra, trans. by Eva Brann (1968, reissued 1992; originally published in German, 1934); and Wilbur Richard Knorr, The Ancient Tradition of Geometric Problems (1986, reissued 1993). Special topics are examined in O.A.W. Dilke, Mathematics and Measurement (1987); Árpád Szabó, The Beginnings of Greek Mathematics, trans. by A.M. Ungar (1978; originally published in German, 1969); and Wilbur Richard Knorr, The Evolution of the Euclidean Elements: A Study of the Theory of Incommensurable Magnitudes and Its Significance for Early Greek Geometry (1975).

Mathematics in the Islamic world
Sources for Arabic mathematics include J.P. Hogendijk (ed. and trans.), Ibn Al-Haytham's Completion of the Conics, trans. from Arabic (1985); Martin Levey and Marvin Petruck (eds. and trans.), Principles of Hindu Reckoning, trans. from Arabic (1965), the only extant text of Kūshyār ibn Labbān's work; Martin Levey (ed. and trans.), The Algebra of Abū Kāmil, trans. from Arabic and Hebrew (1966), with a 13th-century Hebrew commentary by Mordecai Finzi; Daoud S. Kasir (ed. and trans.), The Algebra of Omar Khayyam, trans. from Arabic (1931, reprinted 1972); Frederic Rosen (ed. and trans.), The Algebra of Mohammed ben Musa, trans. from Arabic (1831, reprinted 1986); and A.S. Saidan (ed. and trans.), The Arithmetic of al-Uqlīdisī, trans. from Arabic (1978). Islamic mathematics is examined in J.L. Berggren, Episodes in the Mathematics of Medieval Islam (1986); E.S. Kennedy, Studies in the Islamic Exact Sciences (1983); and Rushdi Rashid (Roshdi Rashed), The Development of Arabic Mathematics: Between Arithmetic and Algebra, trans. by A.F.W. Armstrong (1994; originally published in French, 1984).

European mathematics during the Middle Ages and Renaissance
An overview is provided by Michael S. Mahoney, “Mathematics,” in David C. Lindberg (ed.), Science in the Middle Ages (1978), pp. 145–178. Other sources include Alexander Murray, Reason and Society in the Middle Ages (1978, reissued 1990), chapters 6–8; George Sarton, Introduction to the History of Science (1927–48, reissued 1975), part 2, “From Rabbi Ben Ezra to Roger Bacon,” and part 3, “Science and Learning in the Fourteenth Century”; and, on a more advanced level, Edward Grant and John E. Murdoch (eds.), Mathematics and Its Applications to Science and Natural Philosophy in the Middle Ages (1987). For the Renaissance, see Paul Lawrence Rose, The Italian Renaissance of Mathematics: Studies on Humanists and Mathematicians from Petrarch to Galileo (1975).

Mathematics in the 17th and 18th centuries
An overview of this period is contained in Derek Thomas Whiteside, “Patterns of Mathematical Thought in the Later Seventeenth Century,” Archive for History of Exact Sciences, 1(3):179–388 (1961). Specific topics are examined in Margaret E. Baron, The Origins of the Infinitesimal Calculus (1969, reprinted 1987); Roberto Bonola, Non-Euclidean Geometry: A Critical and Historical Study of Its Development, trans. by H.S. Carslaw (1955; originally published in Italian, 1912); Carl B. Boyer, The Concepts of the Calculus: A Critical and Historical Discussion of the Derivative and the Integral (1939; also published as The History of the Calculus and Its Conceptual Development, 1949, reprinted 1959); Herman H. Goldstine, A History of Numerical Analysis from the 16th Through the 19th Century (1977); Judith V. Grabiner, The Origins of Cauchy's Rigorous Calculus (1981); I. Grattan-Guinness, The Development of the Foundations of Mathematical Analysis from Euler to Riemann (1970); Roger Hahn, The Anatomy of a Scientific Institution: The Paris Academy of Sciences, 1666–1803 (1971); and Luboš Nový, Origins of Modern Algebra, trans. from the Czech by Jaroslav Tauer (1973).

Mathematics in the 19th and 20th centuries
Surveys include Herbert Mehrtens, Henk Bos, and Ivo Schneider (eds.), Social History of Nineteenth Century Mathematics (1981); William Aspray and Philip Kitcher (eds.), History and Philosophy of Modern Mathematics (1988); and Keith Devlin, Mathematics: The New Golden Age, new and rev. ed. (1999). Special topics are examined in Umberto Bottazzini, The Higher Calculus: A History of Real and Complex Analysis from Euler to Weierstrass, trans. by Warren Van Egmond (1986; originally published in Italian, 1981); Julian Lowell Coolidge, A History of Geometrical Methods (1940, reissued 2003); Joseph Warren Dauben, Georg Cantor: His Mathematics and Philosophy of the Infinite (1979, reprinted 1990); Harold M. Edwards, Fermat's Last Theorem: A Genetic Introduction to Algebraic Number Theory (1977, reissued 2000); I. Grattan-Guinness (ed.), From the Calculus to Set Theory, 1630–1910: An Introductory History (1980, reissued 2000); Jeremy Gray, Ideas of Space: Euclidian, Non-Euclidean, and Relativistic, 2nd ed. (1989); Thomas Hawkins, Lebesgue's Theory of Integration: Its Origins and Development, 3rd ed. (1979, reissued 2001); Jesper Lützen, The Prehistory of the Theory of Distributions (1982); and Michael Monastyrsky, Riemann, Topology, and Physics, trans. from Russian by Roger Cooke, James King, and Victoria King, 2nd ed. (1987).John L. Berggren Wilbur R. Knorr Menso Folkerts Craig G. Fraser Jeremy John Gray

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Mathematics — Mathematics …   Википедия

  • Mathematics — oder Allah Mathematics, eigentlich Ronald M. Bean, ist ein US amerikanischer Produzent und DJ der Musikrichtung Hip Hop unter anderem für den Wu Tang Clan. Mathematics tritt neben weiteren Projekten auch als Solokünstler auf und hat als solcher… …   Deutsch Wikipedia

  • mathematics — is treated as a singular noun when it is the name of a subject (Mathematics is not a requirement) and as a plural noun when it means ‘the process of calculating’ (The mathematics of the problem are complex) …   Modern English usage

  • mathematics — ► PLURAL NOUN (usu. treated as sing. ) ▪ the branch of science concerned with number, quantity, and space, either as abstract ideas (pure mathematics) or as applied to physics, engineering, and other subjects (applied mathematics). DERIVATIVES… …   English terms dictionary

  • Mathematics — Math e*mat ics, n. [F. math[ e]matiques, pl., L. mathematica, sing., Gr. ? (sc. ?) science. See {Mathematic}, and { ics}.] That science, or class of sciences, which treats of the exact relations existing between quantities or magnitudes, and of… …   The Collaborative International Dictionary of English

  • mathematics — mathematics, philosophy of …   Philosophy dictionary

  • mathematics — 1580s, plural of MATHEMATIC (Cf. mathematic) (see ICS (Cf. ics)). Originally denoting the mathematical sciences collectively, including geometry, astronomy, optics …   Etymology dictionary

  • mathematics — [n] arithmetic addition, algebra, calculation, calculus, division, figures, geometry, math, multiplication, numbers, subtraction, trigonometry; concepts 349,764 …   New thesaurus

  • mathematics — [math΄ə mat′iks] n. [see MATHEMATICAL & ICS] 1. the group of sciences (including arithmetic, geometry, algebra, calculus, etc.) dealing with quantities, magnitudes, and forms, and their relationships, attributes, etc., by the use of numbers and… …   English World dictionary

  • Mathematics — Maths and Math redirect here. For other uses see Mathematics (disambiguation) and Math (disambiguation). Euclid, Greek mathematician, 3r …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”