Number theory

From Knowino
Jump to: navigation, search

Contents

Number theory is a branch of mathematics devoted primarily to the study of the integers. Any attempt to conduct such a study naturally leads to an examination of the properties of that which integers are made out of (namely, prime numbers) as well as the properties of objects made out of integers (such as rational numbers) or defined as generalisations of the integers (such as, for example, algebraic integers).

Integers can be considered either as such or as solutions to equations (diophantine geometry). Some of the main questions are those of distribution: questions, say, on patterns or their absence (in the primes or other sequences) or, more generally, questions on size, number and growth. Such matters are often best understood through the study of analytical objects (e.g., the Riemann zeta function) that encode them in some fashion (analytical number theory). One may also study real numbers in relation to rational numbers, e.g., as approximated by the latter (diophantine approximation).

The older term for number theory is arithmetic; it was superseded by "number theory" in the nineteenth century, though the adjective arithmetical is still fully current. By 1921, T. L. Heath had to explain: "By arithmetic Plato meant, not arithmetic in our sense, but the science which considers numbers in themselves, in other words, what we mean by the Theory of Numbers."[1] The general public now uses arithmetic to mean elementary calculations, whereas mathematicians use arithmetic as this article shall, viz., as an older synonym for number theory. (The use of the term arithmetic for number theory has regained some ground since Heath's time, arguably in part due to French influence.[2] In particular, arithmetic is preferred as an adjective to number-theoretic. Moreover, "the arithmetic of" is used, whereas "the number theory of" is not; thus, for example, the arithmetic of elliptic curves.)

[edit] Origins

[edit] The dawn of arithmetic

The first historical find of an arithmetical nature is a fragment of a table: the broken clay tablet Plimpton 322 (Larsa, Mesopotamia, ca. 1800 BCE) contains a list of "Pythagorean triples", i.e., integers \scriptstyle (a,b,c) such that \scriptstyle a^2+b^2=c^2. The triples are too many and too large to have been obtained by brute force. The heading over the first column reads: "The takiltum of the diagonal which has been substracted such that the width..."[3]

The table's outlay suggests[4] that it was constructed by means of what amounts, in modern language, to the identity

\left(\frac{1}{2} \left(x - \frac{1}{x}\right)\right)^2 + 1 =
\left(\frac{1}{2} \left(x + \frac{1}{x}\right)\right)^2,

which is implicit in routine Old Babylonian exercises[5]. If some other method was used,[6] the triples were first constructed and then reordered by c / a, presumably for actual use as a "table", i.e., with a view to applications.

We do not know what these applications may have been, or whether there could have been any; Babylonian astronomy, for example, truly flowered only later. It has been suggested instead that the table was a source of numerical examples for school problems.[7][8]

While Babylonian number theory - or what survives of Babylonian mathematics that can be called thus - consists of this single, striking fragment, Babylonian algebra (in the secondary-school sense of "algebra") was exceptionally well developed.[9] Late Neoplatonic sources[10] state that Pythagoras learned mathematics from the Babylonians. (Much earlier sources[11] state that Thales and Pythagoras travelled and studied in Egypt.)

Euclid IX 21--34 is very probably Pythagorean;[12] it is very simple material ("odd times even is odd", "if an odd number measures [= divides] an even number, then it also measures [= divides] half of it"), but it is all that is needed to prove that \scriptstyle \sqrt{2} is irrational[13]. (Pythagoraean mystics gave great importance to the odd and the even[14].) The discovery that \scriptstyle \sqrt{2} is irrational is credited to the early Pythagoreans (pre-Theodorus).[15] By revealing (in modern terms) that numbers could be irrational, this discovery seems to have provoked the first foundational crisis in mathematical history; its proof or its divulgation are sometimes credited to Hippasus, who was expelled or split from the Pythagorean sect.[16] It is only here that we can start to speak of a clear, conscious division between numbers (integers and the rationals - the subjects of arithmetic) and lengths (real numbers, whether rational or not).

The Pythagorean tradition spoke also of so-called polygonal or figured numbers.[17] While square numbers, cubic numbers, etc., are seen now as more natural than triangular numbers, square numbers, pentagonal numbers, etc., the study of the sums of triangular and pentagonal numbers would prove fruitful in the early modern period (17th to early 19th century).

We know of no clearly arithmetical material in ancient Egyptian or Vedic sources, though there is some algebra in both. The Chinese remainder theorem appears as an exercise [18] in Sun Zi's Suan Ching (also known as Sun Tzu's Mathematical Classic; 3rd, 4th or 5th century CE.[19]). (There is one important step glossed over in Sun Zi's solution:[20] it is the problem that was later solved by Āryabhaṭa's kuṭṭaka - see below.)

There is also some numerical mysticism in Chinese mathematics,[21] but, unlike that of the Pythagoreans, it seems to have led nowhere. Like the Pythagoreans' perfect numbers, magic squares have passed from superstition into recreation.

[edit] Classical Greece and the early Hellenistic period

Aside from a few fragments, the mathematics of Classical Greece is known to us either through the reports of contemporary non-mathematicians or through mathematical works from the early Hellenistic period.[22] In the case of number theory, this means, large and by, Plato and Euclid, respectively.

Plato had a keen interest in mathematics, and distinguished clearly between arithmetic and calculation. (By arithmetic he meant, in part, theorising on number, rather than what arithmetic or number theory have come to mean.) It is through one of Plato's dialogues -- namely, Theaetetus -- that we know that Theodorus had proven that \scriptstyle \sqrt{3}, \sqrt{5}, \dots, \sqrt{17} are irrational. Theaetetus was, like Plato, a disciple of Theodorus's; he worked on distinguishing different kind of inconmensurables, and was thus arguably a pioneer in the study of number systems. (Book X of Euclid is described by Pappus as being largely based on Theaetetus's work.)

... Euclid devoted part of his Elements to prime numbers and divisibility, topics that belong unambiguously to number theory and are basic thereto; in particular, he gave the first known proof of the infinitude of primes ...

Archimedes and the cattle problem...

[edit] Diophantus

Very little is known about Diophantus of Alexandria; he probably lived in the third century CE, that is, about five hundred years after Euclid. Six out of the thirteen books of Diophantus's Arithmetica survive in the original Greek; four more books survive in an Arabic translation. The Arithmetica is a collection of worked-out problems where the task is invariably to find rational solutions to a system of polynomial equations, usually of the form \scriptstyle f(x,y)=z^2 or \scriptstyle f(x,y,z)=w^2. Thus, nowadays, we speak of Diophantine equations when we speak of polynomial equations to which rational or integer solutions must be found.

One may say that Diophantus was studying rational points -- i.e., points whose coordinates are rational -- on curves and varieties; however, unlike the Greeks of the Classical period, who did what we would now call basic algebra in geometrical terms, Diophantus did what we would now call basic algebraic geometry in purely algebraic terms. In modern language, what Diophantus does is to find rational parametrisations of many varieties; in other words, he shows how to obtain infinitely many rational numbers satisfying a system of equations by giving a procedure that can be made into an algebraic expression (say, \scriptstyle x=g_1(r,s), \scriptstyle y=g_2(r,s), \scriptstyle z=g_3(r,s), where \scriptstyle g_1, \scriptstyle g_2 and \scriptstyle g_3 are polynomials or quotients of polynomials; this would be what is sought for if such \scriptstyle x, y, z satisfied a given equation \scriptstyle f(x,y)=z^2 (say) for all values of r and s).

Diophantus also studies the equations of some non-rational curves, for which no rational parametrisation is possible. He manages to find some rational points on these curves -- elliptic curves, as it happens, in what seems to be their first known occurrence -- by means of what amounts to a tangent construction: translated into coordinate geometry (which did not exist in Diophantus's time), his method would be visualised as drawing a tangent to a curve at a known rational point, and then finding the other point of intersection of the tangent with the curve; that other point is a new rational point. (Diophantus also resorts to what could be called a special case of a secant construction.)

While Diophantus is concerned largely with rational solutions, he assumes some results on integer numbers; in particular, he seems to assume that every integer is the sum of four squares, though he never states as much explicitly.

[edit] The Indian school: Āryabhaṭa, Brahmagupta, Bhāskara

Āryabhaṭa (476–550 CE) showed that pairs of simultaneous congruences \scriptstyle n\equiv a_1 \text{ mod } m_1, \scriptstyle n\equiv a_2 \text{ mod } m_2 could be solved by a method he called kuṭṭaka, or pulveriser;[23] this is a procedure close to (a generalisation of) the Euclidean algorithm. He seems to have had in mind applications to astronomical calculations.[24]

Brahmagupta (628 CE) started the systematic study of indefinite quadratic equations -- in particular, the misnamed Pell equation, in which Archimedes may have first been interested. Later Sanskrit authors would follow, using Brahmagupta's technical terminology. A general method (the chakravala) for solving Pell's equation was finally found by Jayadeva (cited in the eleventh century; his work is otherwise lost) and Bhāskara II (twelfth century).

[edit] Arithmetic in the Islamic golden age

In the early ninth century, the caliph Al-Ma'mun ordered translations of many Greek mathematical works and at least one Sanskrit work (the Sindhind, generally presumed to be Brahmagupta's Brāhmasphuţasiddhānta[25]), thus giving rise to the rich tradition of Islamic mathematics. Diophantus's main work, the Arithmetica, was translated into Arabic by Qusta ibn Luqa (820-912). Part of the treatise al-Fakhri (by al-Karajī, 953 - ca. 1029) builds on it to some extent. Al-Karajī's contemporary Ibn al-Haytham knew[26] what would later be called Wilson's theorem, which, arguably, was thus the first clearly non-trivial result on congruences to prime moduli ever known.

Other than a treatise on squares in arithmetic progression by Fibonacci - who lived and studied in north Africa and Constantinople during his formative years, ca. 1175-1200 - no number theory to speak of was done in western Europe while it went through the Middle Ages. Matters started to change in Europe in the late Rennaissance, thanks to a renewed study of the works of Greek antiquity. A key catalyst was the textual emendation and translation into Latin of Diophantus's Arithmetica (Bachet, 1621).

[edit] Early modern number theory

[edit] Fermat

Pierre de Fermat (1601 - 1665) never published his writings; in particular, his work on number theory is contained entirely in letters to mathematicians and in private marginal notes[27]. He wrote down nearly no proofs in number theory; he had no models in the area.[28] He did make repeated use of mathematical induction, introducing the method of infinite descent.

One of Fermat's first interests was perfect numbers (which appear in Euclid, Elements IX) and amicable numbers[29]; this led him to work on integer divisors, which were from the beginning among the subjects of the correspondence (1636-onwards) that put him in touch with the mathematical community of the day.[30] He had already studied Bachet's edition of Diophantus carefully;[31] by 1643, his interests had shifted largely to diophantine problems and sums of squares[32] (also treated by Diophantus).

Fermat's achievements in arithmetic include:

This is a consequence of the fact that the order of an element of a group divides the order of the group. The modern proof would have been within Fermat's means (and was indeed given later by Euler), even though the modern concept of a group came long after Fermat or Euler. (It helps to know that inverses exist modulo p (i.e., given a not divisible by a prime p, there is an integer x such that \scriptstyle x a \equiv 1 \text{ mod } p); this fact (which, in modern language, makes the residues mod p into a group, and which was already known to Āryabhaṭa; see above) was familiar to Fermat thanks to its rediscovery by Bachet.[35]

(to be continued)

[edit] Euler

[edit] Lagrange and his contemporaries

[edit] Gauss

[edit] Number theory as it is now

[edit] Introductory texts and elementary tools

[edit] Analytic number theory

Analytic number theory is generally held to denote the study of problems in number theory by analytic means, i.e., by the tools of calculus. Some would emphasize the use of complex analysis: the study of the Riemann zeta function and other L-functions can be seen as the epitome of analytic number theory. At the same time, the subfield is often held to cover studies of elementary problems by elementary means, e.g., the study of the divisors of a number without the use of analysis, or the application of sieve methods. A problem in number theory can be said to be analytic simply if it involves statements on quantity or distribution, or if the ordering of the objects studied (e.g., the primes) is crucial. Several different senses of the word analytic are thus conflated in the designation analytic number theory as it is commonly used.

The following are examples of problems in analytic number theory: the prime number theorem, the Goldbach conjecture (or the twin prime conjecture, or the Hardy-Littlewood conjectures), the Waring problem and the Riemann Hypothesis. Some of the most important tools of analytic number theory are the circle method, sieve methods and L-functions (or, rather, the study of their properties).

One may ask analytic questions about algebraic numbers, and use analytic means to answer such questions; it is thus that algebraic and analytic number theory intersect. For example, one may define prime ideals (generalisations of prime numbers living in the field of algebraic numbers) and ask how many prime ideals there are up to a certain size. This question can be answered by means of an examination of Dedekind zeta functions, which are generalisations of the Riemann zeta function, an all-important analytic object that controls the distribution of prime numbers.

[edit] Algebraic number theory

Algebraic number theory studies number theory using algebraic techniques drawn from group theory and field theory. A principal topic of study is that of the algebraic numbers, which are generalisations of the rational numbers. (Briefly, an algebraic number is any complex number that is a solution to some polynomial equation \scriptstyle f(x)=0 with rational coefficients; for example, every solution x of \scriptstyle x^5 + (11/2) x^3 - 7 x^2 + 9 = 0
(say) is an algebraic number.) Fields of algebraic numbers are also called algebraic number fields.

It could be argued that the simplest kind of number fields (viz., quadratic fields) were already studied by Gauss, as the discussion of quadratic forms in Disquisitiones arithmeticae can be restated in terms of ideals and norms in quadratic fields. (A quadratic field consists of all numbers of the form \scriptstyle a + b \sqrt{d}, where a and b are rational numbers and d is a fixed rational number whose square root is not rational.) For that matter, the 11th-century chakravala method amounts - in modern terms - to an algorithm for finding the units of a real quadratic number field. However, neither Bhāskara nor Gauss knew of number fields as such.

The grounds of the subject as we know it were set in the late nineteenth century, when ideal numbers, the theory of ideals and valuation theory were developed; these are three complementary ways of dealing with the lack of unique factorisation in algebraic number fields. (For example, in the field generated by the rationals and \scriptstyle \sqrt{-5}, the number 6 can be factorised both as \scriptstyle 6 = 2 \cdot 3 and \scriptstyle 6 = (1 + \sqrt{-5}) ( 1 - \sqrt{-5}); all of 2, 3, \scriptstyle 1 + \sqrt{-5} and \scriptstyle 1 - \sqrt{-5} are irreducible, and thus, in a naïve sense, analogous to primes among the integers.) A failure of awareness of this lack had led to an early erroneous "proof" of Fermat's Last Theorem by G. Lamé; the realisation that this proof was erroneous made others study the consequences of this lack, and ways in which it could be alleviated.

Number fields are often studied as extensions of smaller number fields: a field L is said to be an extension of a field K if L contains K. (For example, the complex numbers C are an extension of the reals R, and the reals R are an extension of the rationals Q.) Classifying the possible extensions of a given number field is a difficult and partially open problem. Abelian extensions -- that is, extensions L of K such that the Galois group[36] Gal(L/K) of L over K is an abelian group -- are relatively well understood. Their classification was the object of the programme of class field theory, which was initiated in the late 19th century (partly by Kronecker and Eisenstein) and carried out largely in 1900--1950.

The Langlands program, one of the main current large-scale research plans in mathematics, is sometimes described as an attempt to generalise class field theory to non-abelian extensions of number fields.

[edit] Diophantine geometry

Consider an equation or system of equations. Does it have rational or integer solutions, and if so, how many? This is the central question of Diophantine geometry.

We may think of this question in the following graphic way. An equation in two variables defines a curve in the plane; more generally, an equation, or system of equations, in two or more variables defines a curve, a surface or some other such object in n-dimensional space. We are asking whether there are any rational points (points all of whose coordinates are rationals) or integer points (points all of whose coordinates are integers) on the curve or surface. If there are any such points on the curve or surface, we may ask how many there are and how they are distributed. Most importantly: are there finitely or infinitely many rational points on a given curve (or surface)? What about integer points?

An example here may be helpful. Consider the equation x2 + y2 = 1; we would like to study its rational solutions, i.e., its solutions (x,y) such that x and y are both rational. This is the same as asking for all integer solutions to a2 + b2 = c2; any solution to the latter equation gives us a solution x = a / c, y = b / c to the former. It is also the same as asking for all points with rational coordinates on the curve described by x2 + y2 = 1. (This curve happens to be a circle of radius 1 around the origin.)

The rephrasing of questions on equations in terms of points on curves turns out to be felicitous. The finiteness or not of the number of rational or integer points on an algebraic curve - that is, rational or integer solutions to an equation f(x,y) = 0, where f is a polynomial in two variables - turns out to depend crucially on the genus of the curve. The genus can be defined as follows: allow the variables in f(x,y) = 0 to be complex numbers; then f(x,y) = 0 defines a 2-dimensional surface in 4-dimensional surface; count[37] the number of (doughnut) holes in the surface; call this number the genus of f(x,y) = 0. Other geometrical notions turn out to be just as crucial.

There is also the closely linked area of diophantine approximations: given a number x, how well can it be approximated by rationals? (We are looking for approximations that are good relative to the amount of space that it takes to write the rational: call a / q (with gcd(a,q) = 1) a good approximation to x if \scriptstyle |x-a/q|<\frac{1}{q^c}, where c is large.) This question is of special interest if x is an algebraic number. If x cannot be well approximated, then some equations do not have integer or rational solutions. Moreover, several concepts (especially that of height) turn out to be crucial both in diophantine geometry and in the study of diophantine approximations.

Diophantine geometry should not be confused with the geometry of numbers, which is a collection of graphical methods for answering certain questions in algebraic number theory.

[edit] Arithmetic combinatorics

Let A be a set of integers. Consider the set A + A consisting of all sums of two elements of A. Is A + A much larger than A? Barely larger? If A + A is barely larger than A, must A have plenty of arithmetic structure - e.g., does it look like an arithmetic progression?

If we begin from a fairly "thick" infinite set A (say, the primes), does it contain many elements in arithmetic progression: a, a + b, a + 2b, a + 3b, ... , a + 10b, say? Should it be possible to write large integers as sums of elements of A?

These questions are characteristic of arithmetic combinatorics. This is a presently coalescing field; it subsumes additive number theory (which concerns itself with certain very specific sets A of arithmetic significance, such as the primes or the squares) and, arguably, some of the geometry of numbers, together with some rapidly developing new material. Its focus on issues of growth and distribution make the strengthening of links with ergodic theory likely. The term additive combinatorics is also used; however, the sets A being studied need not be sets of integers, but rather subsets of non-commutative groups, for which the multiplication symbol, not the addition symbol, is traditionally used; they can also be subsets of rings, in which case the growth of A + A and A·A may be compared.

[edit] Probabilistic number theory

Take a number at random between one and a million. How likely is it to be prime? This is just another way of asking how many primes there are between one and a million. Very well; ask further: how many prime divisors will it have, on average? How many divisors will it have altogether, and with what likelihood? What is the probability that it have many more or many fewer divisors or prime divisors than the average?

Much of probabilistic number theory can be seen as an important special case of the study of variables that are almost, but not quite, mutually independent. For example, the event that a random integer between one and a million be divisible by two and the event that it be divisible by three are almost independent, but not quite.

It is sometimes said that probabilistic combinatorics uses the fact that whatever happens with probability greater than 0 must happen sometimes; one may say with equal justice that many applications of probabilistic number theory hinge on the fact that whatever is unusual must be rare. If certain algebraic objects (say, rational or integer solutions to certain equations) can be shown to be in the tail of certain sensibly defined distributions, it follows that there must be few of them; this is a very concrete non-probabilistic statement following from a probabilistic one.

[edit] Computations in number theory

While the word algorithm goes back only to certain readers of al-Khwārizmī, careful descriptions of methods of solution are older than proofs: such methods - that is, algorithms - are as old as any recognisable mathematics - ancient Egyptian, Babylonian, Vedic, Chinese - whereas proofs appeared only with the Greeks of the classical period.

There are two main questions: "can we compute this?" and "can we compute it rapidly?". Anybody can test whether a number is prime or, if it is not, split it into prime factors; doing so rapidly is another matter. We now know fast algorithms for testing primality, but, in spite of much work, no truly fast algorithm for factoring.

The difficulty of a computation can be useful: modern protocols for encrypting messages depend on functions that are known to all, but whose inverses (a) are known only to a chosen few, and (b) would take one too long a time to figure out on one's own. For example, these functions can be such that their inverses can be computed only if certain large integers are factorised. While many difficult computational problems outside number theory are known, most working encryption protocols nowadays are based on the difficulty of a few number-theoretical problems.

On a different note - some things may not be computable at all; in fact, this can be proven. For instance, Turing showed in 1936 that there is no algorithm for deciding in finite time whether a given algorithm ends in finite time. In 1970, it was proven that there is no algorithm for solving any and all Diophantine equations. There are thus some problems in number theory that will never be solved. We even know the shape of some of them, viz., Diophantine equations in nine variables; we simply do not know, and cannot know, which coefficients give us equations for which the following two statements are both true: there are no solutions, and we shall never know that there are no solutions.

[edit] Problems solved and unsolved

[edit] The beginnings

What are the integers x, y, z such that \scriptstyle x^2 + y^2 = z^2?

A scribe from Larsa (1800 BCE) almost certainly had a full solution. The late source Proclus credits Pythagoras with the partial solution \scriptstyle (x,(x^2-1)/2,(x^2+1)/2)
, where x ranges on the odd integers. He also credits Plato with a closely related rule. A general solution makes its first fully explicit appearance in Euclid's Elements (Book X, Lemma 1).

Are there incommensurable line segments? (In our language: are there irrational numbers?)

Yes (early Pythagoreans, before Plato's day). The question belongs to the history of number systems at least as much as it belongs here. The proof in Euclid's Elements is purely arithmetical; nothing besides the "theory of the odd and the even" (likely early Pythagorean) is needed.

Are there infinitely many prime numbers?

Yes (Euclid).

Given two integers, find the largest integer that divides them both.

Euclid's algorithm does the job. It also provides the basis for the standard method for finding integer solutions to linear equations in two variables. Such equations, however, were not addressed by Euclid; the first algorithm found for solving them was Āryabhaṭa's kuṭṭaka (see below).

[edit] Diophantus

[edit] India

[edit] Fermat

[edit] Euler

[edit] Questions directing current research

[edit] References

  1. Sir Thomas Heath, A History of Greek Mathematics, vol. 1, Dover, 1981, p. 13.
  2. Take, e.g., Serre's A Course in Arithmetic (1970; translated into English in 1973). In 1952, Davenport still had to specify that he meant The Higher Arithmetic. Hardy and Wright wrote in the introduction to An Introduction to the Theory of Numbers (1938): "We proposed at one time to change [the title] to An introduction to arithmetic, a more novel and in some ways a more appropriate title; but it was pointed out that this might lead to misunderstandings about the content of the book."
  3. Neugebauer and Sachs, Mathematical cuneiform texts, American Oriental Series, Vol. 29, 1945, p. 40. The term takiltum is problematic. Robson (E. Robson, "Neither Sherlock Holmes nor Babylon: a reassessment of Plimpton 322", Historia Math. 28 (3), p. 192) prefers the rendering "The holding-square of the diagonal from which 1 is torn out, so that the short side comes up...".
  4. Robson, op. cit., p. 189. Other sources give the modern formula \scriptstyle (p^2-q^2,2pq,p^2+q^2). Van der Waerden, Science Awakening, Oxford University Press, New York, 1961, p. 79, gives both the modern formula and what amounts to the form preferred by Robson.
  5. Eleanor Robson, "Neither Sherlock Holmes nor Babylon: a reassessment of Plimpton 322", Historia Math. 28 (3), p. 184.
  6. Neugebauer (O. Neugebauer, The exact sciences in antiquity, Dover, 1969 (corrected preprint of the 1957 ed.), p. 36-40) discusses the table in detail and mentions in passing Euclid's method in modern notation (Neugebauer, op. cit., p. 39)
  7. J. Friberg, Methods and traditions of Babylonian mathematics: Plimpton 322, Pythagorean triples and the Babylonian triangle parameter equations, Historia Math. 8, p. 302.
  8. Eleanor Robson, "Neither Sherlock Holmes nor Babylon: a reassessment of Plimpton 322", Historia Math. 28 (3), p. 201. This is controversial. See Plimpton 322.
  9. B. L. van der Waerden, Science awakening, Oxford University Press, New York, 1961, p. 43.
  10. Iamblichus, Life of Pythagoras, cited in van der Waerden, op. cit., p. 108. See also Porphyry, Life of Pythagoras, paragraph 6, in Kenneth Sylvan Guthrie, The Pythagorean Sourcebook and Library, Phanes Press, 1987. Van der Waerden (op. cit., p. 87-90) sustains the view that Thales knew Babylonian mathematics.
  11. Herodotus (II. 81) and Isocrates (Busiris 28), cited in: C. A. Huffman, "Pythagoras", Stanford Encyclopaedia of Philosophy. On Thales, see Eudemus ap. Proclus, 65.7, cited in: Patricia O'Grady, "Thales of Miletus", The Internet Encyclopaedia of Philosophy.
  12. O. Becker, Quellen und Studien, B 4, p. 533, cited in: van der Waerden, op. cit., p. 108.
  13. O. Becker, op. cit.
  14. van der Waerden, op. cit., p. 109.
  15. Plato, Theaetetus, p. 147 B, cited in: Kurt von Fritz, "The discovery of incommensurability by Hippasus of Metapontum", p. 212, in: J. Christianidis (ed.), Classics in the History of Greek Mathematics, Kluwer, 2004.
  16. von Fritz, op. cit.
  17. T. Heath, A History of Greek Mathematics: From Thales to Euclid, vol. 1, Dover, 1981, p. 76.
  18. Sun Zi, Suan Ching, Chapter 3, Problem 26. This can be found in pp. 219-220 of Lam Lay Yong and Ang Tian Se, Fleeting Footsteps: Tracing the conception of arithmetic and algebra in ancient China, rev. ed., World Scientific, 2004, which contains a full translation of the Suan Ching (based on Qian Baocong (ed.) Suanjing shi shu (Ten mathematical classics). Beijing: Zhonghua shuju, 1963). See also the discussion in Yong and Se, op. cit., pp. 138-140.
  19. The date of the text has been narrowed down to 220-420 AD (Yan Dunjie) or 280-473 AD (Wang Ling) due to internal evidence (= taxation systems assumed in the text). See Yong and Se, op. cit., pp. 27-28.
  20. Sun Zi, Suan Ching, Ch. 3, Problem 26, in Yong and Se, op. cit., pp. 219-220:
    [26] Now there are an unknown number of things. If we count by threes, there is a remainder 2; if we count by fives, there is a remainder 3; if we count by sevens, there is a remainder 2. Find the number of things. Answer: 23. Method: If we count by threes and there is a remainder 2, put down 140. If we count by fives and there is a remainder 3, put down 63. If we count by sevens and there is a remainder 2, put down 30. Add them to obtain 233 and subtract 210 to get the answer. If we count by threes and there is a remainder 1, put down 70. If we count by fives and there is a remainder 1, put down 21. If we count by sevens and there is a remainder 1, put down 15. When [a number] exceeds 106, the result is obtained by subtracting 105.
  21. See, e.g., Sun Zi, Suan Ching, Ch. 3, Problem 36, in Yong and Se, op. cit., pp. 223-224:
    [36] Now there is a pregnant woman whose age is 29. If the gestation period is 9 months, determine the sex of the unborn child. Answer: Male. Method: Put down 49, add the gestation period and subtract the age. From the remainder take away 1 representing the heaven, 2 the earth, 3 the man, 4 the four seasons, 5 the five phases, 6 the six pitch-pipes, 7 the seven stars [of the Dipper], 8 the eight winds, and 9 the nine divisions [of China under Yu the Great]. If the remainder is odd, [the sex] is male and if the remainder is even, [the sex] is female.
    This is the last problem in Sun Zi's otherwise matter-of-fact treatise.
  22. See, e.g., C. Boyer, A History of Mathematics, 2nd. ed., revised by Merzbach, Wiley, 1991, p. 82.
  23. Āryabhaṭa, Āryabhatīya, Chapter 2, verses 32-33, cited in: K. Plofker, Mathematics in India, Princeton University Press, 2008, pp. 134-140. See also W. E. Clark, The Āryabhaṭīya of Āryabhaṭa: An ancient Indian work on Mathematics and Astronomy, University of Chicago Press, 1930, pp. 42-50. A slightly more explicit description of the kuṭṭaka was later given in Brahmagupta, Brāhmasphuṭasiddhānta, XVIII, 3-5 (in Colebrooke, Algebra, with Arithmetic and Mensuration, from the Sanscrit of Brahmegupta and Bháscara, London, 1817, p. 325, cited in: Clark, op. cit., p. 42).
  24. Plofker, op. cit., p. 119.
  25. Colebrook, op. cit., lxv, cited in J. F. P. Hopkins, Geographical and navigational literature, Ch. 17 of M. J. L. Young, J. D. Latham and R. B. Serjeant, eds., The Cambridge history of Arabic literature: Religion, learning and science in the `Abbasid period, Cambridge University Press, 1990, p. 302. See also the preface in E. Sachau, Alberuni's India: An account of the religion, philosophy, literature, geography, chronology, astronomy and astrology of India, Vol I, 1888, cited in: D. E. Smith, History of Mathematics, Vol I., Dover, 1958, p. 168.
  26. Roshdi Rashed, Ibn al-Haytham el le théorème de Wilson, Arch. Hist. Exact Sci. 22 (1980), no. 4, pp. 305-321.
  27. André Weil, Number theory: an approach through history - from Hammurapi to Legendre, Birkhäuser, 1984, pp. 45-46.
  28. Weil, op. cit., p. 118. This was more so in number theory than in other areas (remark in M. S. Mahoney, The mathematical career of Pierre de Fermat, 1601-1665, Princeton Univ. Press, 1994, 2nd ed., p. 284). Bachet's own proofs were "ludicrously clumsy" (Weil, op. cit., p. 33).
  29. Perfect and especially amicable numbers are of little or no interest nowadays. The same was not true in medieval times - whether in the West or the Arab-speaking world - due in part to the importance given to them by the Neopythagorean (and hence mystical) Nicomachus (ca. 100 CE), who wrote a primitive but influential "Introduction to Arithmetic". See Van der Waerden, Science Awakening, Oxford, 1961, Ch. IV.
  30. Mahoney, op. cit., p. 48 and pp. 53-54. The initial subjects of Fermat's correspondence included divisors ("aliquot parts") and many subjects outside number theory; see the list in the letter from Fermat to Roberval, 22.IX.1636, Ch. Henry and Tannery (eds.), Oeuvres de Fermat, Paris, 1891-1912, vol. II, pp. 72 and 74, cited in: Mahoney, op. cit., p. 54.
  31. Weil, op. cit., pp. 1-2.
  32. Weil, op. cit., p. 53
  33. Letter from Fermat to Frenicle, 1640, cited in: Weil, op. cit., p. 56
  34. Here, as usual, given two integers a and b and a non-zero integer m, we write \scriptstyle a \equiv b \text{ mod } m (read "a is congruent to b modulo m") to mean that m divides a-b, or, what is the same, a and b leave the same residue when divided by m. This notation is actually much later than Fermat's; it first appears in section 1 of Gauss's Disquisitiones Arithmeticae.
  35. Weil, op. cit., p. 7. Weil goes on to say that Fermat would have recognised that Bachet's argument is essentially Euclid's algorithm.
  36. The Galois group of an extension K/L consists of the operations (isomorphisms) that send elements of L to other elements of L while leaving all elements of K fixed. Thus, for instance, Gal(C/R) consists of two elements: the identity element (taking every element x+iy of C to itself) and complex conjugation (the map taking each element x+iy to x-iy). The Galois group of an extension tells us many of its crucial properties. The study of Galois groups started with Evariste Galois; in modern language, the main outcome of his work is that an equation f(x)=0 can be solved by radicals (that is, x can be expressed in terms of the four basic operations together with square roots, cubic roots, etc.) if and only if the extension of the rationals by the roots of the equation f(x)=0 has a Galois group that is solvable in the sense of group theory. ("Solvable", in the sense of group theory, is a simple property that can be checked easily for finite groups.)
  37. It may be useful to look at an example here. Say we want to study the curve y2 = x3 + 7. We allow x and y to be complex numbers: (a + bi)2 = (c + di)3 + 7. This is, in effect, a set of two equations on four variables, since both the real and the imaginary part on each side must match. As a result, we get a surface (two-dimensional) in four dimensional space. After we choose a convenient hyperplane on which to project the surface (meaning that, say, we choose to ignore the coordinate a), we can plot the resulting projection, which is a surface in ordinary three-dimensional space. It then becomes clear that the result is a torus, i.e., the surface of a doughnut (somewhat stretched). A doughnut has one hole; hence the genus is 1.

[edit] External links

Information.svg Some content on this page may previously have appeared on Citizendium.
Personal tools
Variants
Actions
Navigation
Community
Toolbox