Jump to content

Number theory: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Reverted edits by 174.117.152.251 (talk) to last version by D.Lazard
Garald (talk | contribs)
fresh restart
Line 1: Line 1:
'''Number theory''' is a branch of [[pure 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 [[prime numbers]] (the building blocks of integers) as well
[[Image:Ulam Spiral Divisors 100000.png|300px|right|thumb|In this graphic the [[natural numbers]] are arranged as in the [[Ulam spiral]] and a disk of size proportional to the number of [[divisor]]s is drawn for each number: this yields an intriguing, yet not fully understood pattern.]]
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]]).
'''Number theory''' is the branch of [[pure mathematics]] concerned with the properties of [[number]]s in general, and [[integer]]s in particular, as well as the wider classes of problems that arise from their study.


Integers can be considered either in themselves or as solutions to equations
Number theory may be subdivided into several fields, according to the methods used and the type of questions investigated. (''See the [[list of number theory topics]]''.)
([[diophantine geometry]]). Questions in number theory are often best understood through
the study of [[Complex analysis|analytical]] objects (e.g., the [[Riemann zeta function]]) that encode properties of the integers, primes or other number-theoretic objects in some fashion ([[analytic 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The older term for number theory is ''arithmetic''; it was superseded by "number theory"
The terms "[[arithmetic]]" or "the higher arithmetic" as [[nouns]] are also used to refer to elementary number theory. These are somewhat older terms, which are no longer as popular as they once were. However the word "arithmetic" is popularly used as an [[adjective]] rather than the more cumbersome phrase "number-theoretic", and also "arithmetic of" rather than "number theory of"; e.g., [[arithmetic geometry]], [[arithmetic function]]s, [[arithmetic of elliptic curves]].
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."<ref name=HeathAr>Sir Thomas Heath, A History of Greek Mathematics, vol. 1, Dover,
1981, p. 13.</ref> 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.<ref>Take, e.g.,
[[Jean-Pierre Serre|Serre]]'s ''A Course in Arithmetic'' (1970; translated into
English in 1973). In 1952, [[Harold Davenport|Davenport]] still had to specify that he
meant ''The Higher Arithmetic''. [[G. H. Hardy|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."</ref> 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]]''.)


==Fields==
== History ==
===Elementary number theory===
In '''elementary number theory''', integers are studied without use of techniques from other mathematical fields. Questions of [[divisibility]], use of the [[Euclidean algorithm]] to compute the [[greatest common divisor]] of numbers, [[integer factorization]]s, investigation of [[perfect number]]s, [[modular arithmetic|congruences]], etc. belong here. Several important discoveries of this field are [[Fermat's little theorem]], [[Euler's theorem]], the [[Chinese remainder theorem]] and the law of [[quadratic reciprocity]]. The properties of [[multiplicative function]]s such as the [[Möbius function]] and [[Euler's phi function|Euler's φ function]], [[integer sequence]]s, [[factorial]]s, and [[Fibonacci number]]s all also fall into this area.


{{main|User:Garald/History of number theory}}
Many questions in number theory can be stated in elementary number theoretic terms, but they may require very deep consideration and new approaches outside the realm of elementary number theory to solve. Examples include:
* [[Goldbach's conjecture]] concerning the expression of [[even and odd numbers|even]] numbers as sums of two primes.
* [[Mihăilescu's theorem]] (formerly Catalan's conjecture) regarding successive integer powers.
* The twin prime conjecture about the infinitude of [[Twin prime|prime pairs]].
* The [[Collatz conjecture]] concerning a simple iteration.
* [[Fermat's Last Theorem]] (stated in 1637, but not proven until 1994) concerning the impossibility of finding nonzero integers ''x, y, z'' such that <math>x^n + y^n = z^n</math> for some integer ''n'' greater than ''2''.


===The beginnings===
The theory of [[Diophantine equation]]s has even been shown to be ''[[Decision problem|undecidable]]'' (see [[Hilbert's tenth problem]]).


While there are elements of what in retrospect can be seen as number theory
===Analytic number theory===
in [[Babylonian mathematics|Babylonian]] and ancient Chinese mathematics
'''[[Analytic number theory]]''' employs the machinery of [[calculus]] and [[complex analysis]] to tackle questions about integers. The [[prime number theorem]] (PNT) and the related [[Riemann hypothesis]] are examples. [[Waring's problem]] (representing a given integer as a sum of [[square number|squares]], [[Cube (arithmetic)|cubes]] etc.), the [[Twin prime|twin prime conjecture]] (finding infinitely many prime pairs with difference 2) and [[Goldbach's conjecture]] (writing even integers as sums of two primes) are being attacked with analytical methods as well. [[Mathematical proof|Proofs]] of the [[Transcendental number|transcendence]] of mathematical constants, such as [[pi|π]] or ''[[e (mathematical constant)|e]]'', are also classified as analytical number theory. While statements about transcendental numbers may seem to be removed from the study of integers, they really study the possible values of [[polynomials]] with integer coefficients evaluated at, say, ''e''; they are also closely linked to the field of [[Diophantine approximation]], where one investigates "how well" a given real number may be approximated by a [[rational number|rational]] one.
(see [[Plimpton 322]] and the [[Chinese Remainder Theorem]], respectively), the history of number theory truly starts with the Greek and Indian traditions.


The [[irrational number|irrationality]] of <math>\scriptstyle \sqrt{2}</math> is credited to
===Algebraic number theory===
the early [[Pythagoreans]].<ref>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. [[Plato]] reports on further work by Theodorus on
In '''[[algebraic number theory]]''', the concept of a number is expanded to the [[algebraic number]]s which are [[Root of a function|roots]] of polynomials with [[rational number|rational]] coefficients. These domains contain elements analogous to the integers, the so-called [[algebraic integer]]s. In this setting, the familiar features of the integers (e.g., unique factorization) need not hold. The virtue of the machinery employed—[[Galois theory]], [[group cohomology]], [[class field theory]], [[group representation]]s and [[L-function]]s—is that it allows one to recover that order partly for this new class of numbers.
irrationality.</ref> [[Euclid]] gave an algorithm for computing the
greatest common divisor of two numbers ([[Euclid's Elements]], Prop. VII.2) and a proof
that there are infinitely many primes (Elements, Prop. IX.20).
Much later in the Hellenistic period, [[Diophantus]] studied rational solutions to equations
and systems of equations.


Results in number theory within [[Indian mathematics]] date from the period that would correspond to the medieval era in Europe. [[Aryabhata]] gave an algorithm for solving<ref name="Aryabhata">Āryabhaṭa,
Many number theoretic questions are best attacked by studying them ''modulo p'' for all primes ''p'' (see [[finite field]]s). This leads to the construction of the [[p-adic number]]s; this field of study is called [[local analysis]] and it arises from algebraic number theory.
''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).</ref>
pairs of [[congruences]]
<math>\scriptstyle n\equiv a_1 \text{ mod } m_1</math>,
<math>\scriptstyle n\equiv a_2 \text{ mod } m_2</math>
apparently with astronomical applications in mind.<ref name="Plofker">K. Plofker, ''Mathematics in India'', Princeton University Press, 2008, p. 119.</ref>
[[Brahmagupta]] started the systematic study of indefinite quadratic equations, including
what would later be misnamed [[Pell's equation]]. A general procedure (the [[chakravala method|chakravala]], or "cyclic method") for solving Pell's equation was
finally found by Jayadeva (cited in the eleventh century; his work is otherwise lost); the earliest surviving exposition appears in [[Bhāskara II]]'s
Bīja-gaṇita (twelfth century).<ref name="PlofBha">Plofker, op. cit., p. 194</ref>
Unfortunately, these achievements were largely unknown in the West until the late eighteenth century.<ref name="Ploper">Plofker, op. cit., p. 283</ref>


Much Greek mathematics and some Indian mathematics was available to Arabic scholars from the early ninth century onwards. Part of the treatise ''al-Fakhri'' (by [[al-Karaji|al-Karajī]], 953 - ca. 1029) builds on Diophantus's work to some extent.
===Geometry of numbers===
The '''[[geometry of numbers]]''' incorporates some basic geometric concepts, such as [[Lattice_(group)|lattices]], into number-theoretic questions. It starts with [[Minkowski's theorem]] about [[lattice point]]s in [[convex set]]s, and leads to basic proofs of the finiteness of the [[Ideal class group|class number]] and [[Dirichlet's unit theorem]], two fundamental theorems in algebraic number theory.


===Combinatorial number theory===
=== Modern number theory ===
'''Combinatorial number theory''' deals with number theoretic problems which involve [[combinatorial]] ideas in their formulations or solutions. [[Paul Erdős]] is the main founder of this branch of number theory. Typical topics include [[Partition (number theory)|partitions]], [[covering system]], [[zero-sum problems]], various [[restricted sumset]]s, and [[arithmetic progressions]] in a set of integers. Algebraic or analytic methods are powerful in this field. See also [[arithmetic combinatorics]].


Modern number theory begins with [[Pierre de Fermat]], inspired in part by his study of Diophantus. Continuous activity on the subject started almost a century later with [[Euler]].<ref>A. Weil, 'Number theory: an approach through history - from Hammurapi to Legendre'', Birkhäuser, 1984, pp. 1-2.</ref> [[Lagrange]] provided proofs of some of Fermat's and Euler's key statements. He and [[Legendre]] also set the basis
===Computational number theory===
of the study of quadratic forms; Legendre was the first to state the law of [[quadratic reciprocity]]. In [[Disquisitiones Arithmeticae]], Gauss gave the first valid proof of this law, developed the theory of quadratic forms further, and started the modern study of [[cyclotomy]].
'''[[Computational number theory]]''' studies [[algorithms]] relevant in number theory. Fast algorithms for [[prime testing]] and [[integer factorization]] have important applications in [[cryptography]].


Starting early in the nineteenth century, the following developments gradually took place:
===Arithmetic algebraic geometry===
* The rise to self-consciousness of number theory (or ''higher arithmetic'') as a field of study.<ref>See the discussion in section 5 of C. Goldstein and N. Schappacher, "A book in search of a discipline (1801-1860)', in C. Goldstein, N. Schappacher and J. Schwermer (eds.), "The shaping of arithmetic after C. F. Gauss's Disquisitiones Arithmeticae", Springer, 2007. Early signs of self-consciousness are present already in letters by Fermat: thus his remarks on what number theory is, and how "Diophantus's work [...] does not really belong to [it]" (quoted in A. Weil, op. cit., p. 25).</ref>
See [[arithmetic geometry]].
* The development of much of modern mathematics necessary for basic modern number theory: complex analysis, group theory, Galois theory -- accompanied by greater rigor in analysis and abstraction in algebra.
* The rough subdivision of number theory into its modern subfields - in particular, [[analytic number theory|analytic]] and [[algebraic number theory]].


Algebraic number theory may be said to start with the study of reciprocity and cyclotomy, but truly came into its own with the development of abstract algebra and early ideal theory and valuation theory; see below. An obvious conventional starting point for analytic number theory would be [[Riemann]]'s memoir on the [[Riemann zeta function]] (1859); there is also
===Arithmetic topology===
[[Dirichlet's theorem on arithmetic progressions]], which preceded it in the study of the
'''[[Arithmetic topology]]''' developed from a series of analogies between [[number field]]s and [[3-manifolds]]; [[primes]] and [[Knot (mathematics)|knots]] pointed out by [[Barry Mazur]] and by [[Yuri Manin]] in the 1960s.
[[Riemann zeta function|zeta function]] (for <math>\scriptstyle Re(s)>1</math>), or
[[Jacobi]]'s work on the four square theorem, which connected arithmetical questions with [[elliptic functions]]. The first use of analytical arguments in number theory goes further back,
to [[Euler]].<ref>H. Iwaniec and E. Kowalski, Analytic number theory, AMS Colloquium Pub., Vol. 53, 2004, p. 1.</ref>


The history of each subfield is sketched in its own section below. Many of the most interesting questions in each area remain open and are being actively worked on.
===Arithmetic dynamics===
'''[[Arithmetic dynamics]]''' is a field that emerged in the 1990s that amalgamates two areas of mathematics, [[dynamical systems]] and number theory. Classically, discrete dynamics refers to the study of the [[Iterated function|iteration]] of self-maps of the [[complex plane]] or [[real line]]. Arithmetic dynamics is the study of the number-theoretic properties of integer, rational, {{math|<var>p</var>}}-adic, and/or algebraic points under repeated application of a [[polynomial]] or [[rational function]].


== Approaches and subfields ==
===Modular forms===
[[Modular forms]] are (complex) [[analytic function]]s on the [[upper half-plane]] satisfying a certain kind of [[functional equation]] and growth condition. The theory of modular forms therefore belongs to [[complex analysis]] but the main importance of the theory has traditionally been in its connections with number theory. Modular forms appear in other areas, such as [[algebraic topology]] and [[string theory]].


===Introductory texts and elementary tools===
==History==
===Greek number theory===


Two of the most popular introductions to the subject are:
The foundation of Number Theory was stated by Euclid in his Elements. Euclid has provided an exceptionally beautiful [[Euclid's theorem|proof of the infinitude of the set of prime numbers]]. He also characterized the prime numbers by showing that indecomposability is equivalent to primality in the domain of positive integers. His so-called Euclid algorithm plays a major role in Number Theory; it led centuries later to the definition of so-called euclidean rings. Euclid provided one half of the characterization of even perfect numbers (the other half was done by Euler).
* [[G_H_Hardy|G. H. Hardy]] and E. M. Wright, ''An introduction to the theory of numbers'', 6th ed., rev. by D. R. Heath-Brown and J. H. Silverman, Oxford University Press, Oxford, 2008 (first published in 1938).
* [[Ivan_Matveyevich_Vinogradov|I. M. Vinogradov]], ''Elements of Number Theory'', Mineola, NY: Dover Publications, 2003, reprint of the 1954 edition.


Hardy and Wright's book is a comprehensive classic, though its clarity sometimes suffers due to the authors' insistence on elementary methods.<ref name="MR0568909">T. M. Apostol,
Eratosthenes has invented his algorithm for compiling lists of prime numbers efficiently. His algorithm, Eratosthenes Sieve, is used to this day, these days too.
Review of ''An introduction to the theory of numbers'', Mathematical Reviews, MR0568909.</ref>
Vinogradov's main attraction consists in its set of problems, which quickly lead to Vinogradov's own research interests; the text itself is very basic and close to minimal.


The term ''[[elementary proof|elementary]]'' generally denotes a method that does not use [[complex analysis]]. For example, the [[prime number theorem]] was first proven in 1896, but an elementary proof was found only in 1949. The term is somewhat ambiguous: for example, proofs based on [[Tauberian theorem|Tauberian theorems]] are often seen as quite enlightening but not elementary, in spite of using Fourier analysis, rather than complex analysis as such. Here as elsewhere, an ''elementary'' proof may be longer and more difficult for most readers than a non-elementary one.
Newer historical findings show that also Archimedes was studying difficult number theoretical problems.


Number theory has the reputation of being a field many of whose results can be stated to the layperson. At the same time, the proofs of these results are not particularly accessible, in part because the range of tools they use is, if anything, unusually broad within mathematics.
Number theory was a favorite study among the [[Greek mathematics|Greek mathematicians]] of the late Hellenistic period (3rd century AD) in [[Alexandria]], [[Egypt]], who were aware of the [[Diophantine equation]] concept in numerous special cases. The first Greek mathematician to study these equations was [[Diophantus]].


Popular choices for a second textbook include [[Borevich]] and [[Igor_Shafarevich|Shafarevich]]'s ''Number theory'' and [[Jean-Pierre_Serre|Serre]]'s ''Cours d'arithmetique''. Textbooks for later stages in one's study tend to branch into analytic and algebraic number theory, among other subfields.
Diophantus also looked for a method of finding integer solutions to [[linear equation|linear]] [[indeterminate equation]]s, equations that lack sufficient information to produce a single discrete set of answers. The equation <math>x + y = 5</math> is such an equation. Diophantus discovered that many indeterminate equations can be reduced to a form where a certain category of answers is known even though a specific answer is not.


===Main fields===
===Classical Indian number theory===
====Analytic number theory====
[[Diophantine equations]] were extensively studied by mathematicians in medieval India, who were the first to systematically investigate methods for the determination of integral solutions of Diophantine equations. [[Aryabhata]] (499) gave the first explicit description of the general integral solution of the linear Diophantine equation <math>ay + bx = c</math>, which occurs in his text ''Aryabhatiya''. This ''kuttaka'' algorithm is considered to be one of the most significant contributions of Aryabhata in pure mathematics, which found solutions to Diophantine equations by means of [[continued fraction]]s. The technique was applied by Aryabhata to give integral solutions of simultaneous linear Diophantine equations, a problem with important applications in astronomy. He also found the general solution to the [[indeterminate equation|indeterminate]] [[linear equation]] using this method.{{Citation needed|date=September 2010}}


{{main|Analytic number theory}}
[[Brahmagupta]] in 628 handled more difficult Diophantine equations. He used the [[Chakravala method|''chakravala'' method]] to solve [[quadratic equation|quadratic]] Diophantine equations, including forms of [[Pell's equation]], such as <math>61x^2 + 1 = y^2</math>. His [[Brahmasphutasiddhanta|''Brahma Sphuta Siddhanta'']] was translated into [[Arabic]] in 773 and was subsequently translated into [[Latin]] in 1126. The equation <math>61x^2 + 1 = y^2</math> was later posed as a problem in 1657 by the [[France|French]] mathematician [[Pierre de Fermat]]. The general solution to this particular form of Pell's equation was found over 70 years later by [[Leonhard Euler]], while the general solution to Pell's equation was found over 100 years later by [[Joseph Louis Lagrange]] in 1767. Meanwhile, many centuries ago, the general solution to Pell's equation was recorded by [[Bhaskara II]] in 1150, using a modified version of Brahmagupta's ''chakravala'' method, which he also used to find the general solution to other indeterminate quadratic equations and quadratic Diophantine equations. Bhaskara's ''chakravala'' method for finding the general solution to Pell's equation was much simpler than the method used by Lagrange over 600 years later. Bhaskara also found solutions to other indeterminate quadratic, [[cubic equation|cubic]], [[quartic equation|quartic]], and higher-order [[polynomial]] equations. [[Narayana Pandit]] further improved on the ''chakravala'' method and found more general solutions to other indeterminate quadratic and higher-order polynomial equations.{{Citation needed|date=September 2010}}


''Analytic number theory'' is generally held to denote the study of problems in number theory by analytic means, i.e., by the tools of
===Islamic number theory===
[[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 [[prime numbers|primes]]) is crucial. Several different senses of the word ''analytic'' are thus conflated in the designation ''analytic number theory'' as it is commonly used.
From the 9th century, [[Islamic mathematics|Islamic mathematicians]] had a keen interest in number theory. The first of these mathematicians was [[Thabit ibn Qurra]], who discovered an algorithm which allowed pairs of [[amicable number]]s to be found, that is two numbers such that each is the sum of the proper divisors of the other. In the 10th century, [[Ibn Tahir al-Baghdadi]] looked at a slight variant of Thabit ibn Qurra's method.{{Citation needed|date=September 2010}}


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).
In the 10th century, [[Alhazen|al-Haitham]] seems to have been the first to attempt to classify all even [[perfect number]]s (numbers equal to the sum of their proper divisors) as those of the form <math>2^{k-1}(2^k - 1)</math> where <math>2^k - 1</math> is prime. Al-Haytham is also the first person to state [[Wilson's theorem]], namely that if p is prime then <math>1+(p-1)!</math> is divisible by ''p''. It is unclear whether he knew how to prove this result. It is called Wilson's theorem because of a comment made by [[Edward Waring]] in 1770 that [[John Wilson (mathematician)|John Wilson]] had noticed the result. There is no evidence that Wilson knew how to prove it and most certainly Waring did not. Lagrange gave the first proof in 1771.{{Citation needed|date=September 2010}}


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 number|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 function]]s, which are generalisations of the [[Riemann zeta function]], an all-important analytic object that controls the distribution of prime numbers.
Amicable numbers played a large role in Islamic mathematics. In the 13th century, [[Persian people|Persian]] mathematician [[Al-Farisi]] gave a new proof of [[Thabit number|Thabit ibn Qurra's theorem]], introducing important new ideas concerning factorisation and combinatorial methods. He also gave the pair of amicable numbers 17296, 18416 which have been attributed to Euler, but we know that these were known earlier than al-Farisi, perhaps even by Thabit ibn Qurra himself. In the 17th century, [[Muhammad Baqir Yazdi]] gave the pair of amicable numbers 9,363,584 and 9,437,056 still many years before Euler's contribution.{{Citation needed|date=September 2010}}


===Early European number theory===
====Algebraic number theory====
In the 13th century, Leonardo de Pisa (better known as [[Fibonacci]]), wrote one of his greatest works, the ''[[Liber Quadratorum]]''. In this work he deals with [[Pythagorean triples|Pythagorean triple]]. He noted that square numbers can be constructed as sums of odd numbers. He defined the concept of a congruum, a number of the form ab(a + b)(a - b), if a + b is even, and 4 times this if a + b is odd. Fibonacci proved that a congruum must be divisible by 24 and he also showed that for x, c such that x2 + c and x2 - c are both squares, then c is a congruum. He also proved that a square cannot be a congruum.<ref>O'Connor, John J.; Robertson, Edmund F, [http://www-history.mcs.st-andrews.ac.uk/Biographies/Fibonacci.html Fibonacci], [[MacTutor History of Mathematics archive]], [[University of St Andrews]]</ref> His contribution to number theory were so great that it has been said that "''the Liber quadratorum alone ranks Fibonacci as the major contributor to number theory between [[Diophantus]] and the 17th-century French mathematician [[Pierre de Fermat]]''".<ref>[http://books.google.es/books?id=sIk2_5kLwqIC&pg=PA2&dq=squaring+the+circle.thinking&cd=1#v=onepage&q=Fibonacci&f=false Duthel,Heinz:''Squaring the circle-thinking the unthinkable",p.84'']</ref>


{{main|Algebraic number theory}}
Further advances were done in the 16th and 17th centuries, with [[Franciscus Vieta|Vieta]], [[Bachet de Meziriac]], and especially [[Fermat]], whose [[infinite descent]] approach was an important general method of solving diophantine questions. [[Fermat's Last Theorem]] was posed as a problem in 1637, a proof of which wasn't found until 1994. Fermat also posed the equation <math>61x^2 + 1 = y^2</math> as a problem in 1657. Fermat's theorem about prime numbers congruent to 1 mod 4 being a sum of two squares prompted the development of algebraic number theory. Number Theory gained a status as a separate mathematical domain chiefly thanks to Fermat's contributions and his relentless promotion of Number Theory.


''Algebraic number theory'' studies algebraic properties and algebraic objects of interest in number theory. (Thus, analytic and algebraic number theory can and do overlap:
In the eighteenth century, Euler and Lagrange made important contributions to number theory. Euler created [[analytic number theory]], he showed that the sum of inverses of prime numbers is infinite. He found a general solution to the equation <math>61x^2 + 1 = y^2</math>. Euler and Lagrange solved these more general Pell equations by means of [[continued fraction]]s. Euler completed the characterization of perfect numbers (the first half of which was obtained by Euclid). He also showed that the sum of inverses of squares of all positive integers is equal to pi^2/6, and more advanced results.
the former is defined by its methods, the latter by its objects of study.)
A key topic is that of the [[algebraic number|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 <math>\scriptstyle f(x)=0</math> with rational coefficients;
for example, every solution <math>x</math> of <math>\scriptstyle x^5 + (11/2) x^3 - 7 x^2 + 9 = 0
</math> (say) is an algebraic number. Fields of algebraic numbers are also called ''[[algebraic number field]]s''.


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 [[Ideal (mathematics)|ideals]] and
Fermat, Descartes, Euler and others found many pluperfect numbers.
[[Norm (mathematics)|norms]] in quadratic fields. (A ''quadratic field'' consists of all
numbers of the form <math>\scriptstyle a + b \sqrt{d}</math>, where
<math>a</math> and <math>b</math> are rational numbers and <math>d</math>
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 II|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
===Beginnings of modern number theory===
and <math>\scriptstyle \sqrt{-5}</math>, the number <math>6</math> can be factorised both as <math>\scriptstyle 6 = 2 \cdot 3</math> and
Around the beginning of the nineteenth century books of [[Adrien-Marie Legendre|Legendre]] (1798), and [[Carl Friedrich Gauss|Gauss]] put together the first systematic theories in Europe. Gauss's ''[[Disquisitiones Arithmeticae]]'' (1801) may be said to begin the modern theory of numbers.
<math>\scriptstyle 6 = (1 + \sqrt{-5}) ( 1 - \sqrt{-5})</math>; all of <math>2</math>, <math>3</math>, <math>\scriptstyle 1 + \sqrt{-5}</math> and
<math>\scriptstyle 1 - \sqrt{-5}</math>
are irreducible, and thus, in a naïve sense, analogous to primes among the integers.)
The initial impetus for the development of ideal numbers (by [[Ernst Kummer|Kummer]]) seems to have come from the study
of higher reciprocity laws,<ref name="Edwards">H. M. Edwards, Fermat's Last Theorem: a genetic introduction to algebraic number theory, Springer Verlag, 1977, p. 79.</ref> i.e., generalisations of [[quadratic reciprocity]].


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''.
The formulation of the theory of [[congruence relation|congruences]] starts with Gauss's ''Disquisitiones''. He introduced the notation
(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]]<ref>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.)</ref> ''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.
:<math>a \equiv b \pmod c,</math>


====Diophantine geometry====
and explored most of the field. [[Pafnuty Chebyshev|Chebyshev]] published in 1847 a work in Russian on the subject, and in France [[Joseph Alfred Serret|Serret]] popularised it.


{{main|Diophantine geometry}}
Besides summarizing previous work, Legendre stated the [[law of quadratic reciprocity]]. This law, discovered by [[Mathematical induction|induction]] and enunciated by Euler, was first proved by Legendre in his ''Théorie des Nombres ''(1798) for special cases. Independently of Euler and Legendre, Gauss discovered the law about 1795, and was the first to give a general proof. The following have also contributed to the subject: [[Augustin Louis Cauchy|Cauchy]]; [[Dirichlet]] whose ''[[Vorlesungen über Zahlentheorie]]'' is a classic; [[Carl Gustav Jakob Jacobi|Jacobi]], who introduced the [[Jacobi symbol]]; [[Liouville]], [[Christian Zeller|Zeller]], [[Gotthold Eisenstein|Eisenstein]], [[Ernst Kummer|Kummer]], and [[Kronecker]]. The theory extends to include [[Cubic reciprocity|cubic]] and [[quartic reciprocity]], (Gauss, Jacobi who first proved the law of cubic reciprocity, and Kummer).
{{main|Glossary of arithmetic and Diophantine geometry}}


The central problem of ''Diophantine geometry'' is to determine when a [[Diophantine equation]] has solutions, and if it does, how many. The approach taken is to think of the solutions of an equation as a geometric object.
To Gauss is also due the representation of numbers by binary [[quadratic form]]s.


For example, 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 [[algebraic curve|curve]], a [[algebraic surface|surface]] or some other such object in ''n''-dimensional space. In Diophantine geometry, one asks whether there are any ''[[rational points]]'' (points all of whose coordinates are rationals) or
===Prime number theory===
''integral points'' (points all of whose coordinates are integers) on the curve or surface. If there are any such points, the next step is to ask how many there are and how they are distributed. A basic question in this direction is: are there finitely
A recurring and productive theme in number theory is the study of the distribution of prime numbers. [[Carl Friedrich Gauss]] conjectured an [[Asymptotic analysis|asymptotic]] rule for such behaviour (the [[prime number theorem]]) as a teenager.
or infinitely many rational points on a given curve (or surface)? What about integer points?


An example here may be helpful. Consider the equation <math>x^2+y^2 = 1</math>;
[[Dirichlet]] (1837) proved that every eligible arithmetic progression contains infinitely many prime numbers. [[Chebyshev]] (1850) gave useful upper and lower bounds for the number of primes in initial intervals of natural numbers. Riemann introduced [[complex analysis]] into the theory of the [[Riemann zeta function]]. This led to a relation between the zeros of the zeta function and the distribution of primes, eventually leading to a proof of prime number theorem independently by [[Jacques Hadamard|Hadamard]] and [[Charles Jean de la Vallée-Poussin|de la Vallée Poussin]] in 1896. However, an elementary proof was given later by [[Paul Erdős]] and [[Atle Selberg]] in 1949. Here elementary means that it does not use techniques of complex analysis; however, the proof is still very ingenious and difficult. The [[Riemann hypothesis]], which would give much more accurate information, is still an open question.
we would like to study its rational solutions, i.e., its solutions
<math>(x,y)</math> such that
''x'' and ''y'' are both rational. This is the same as asking for all integer solutions
to <math>a^2 + b^2 = c^2</math>; any solution to the latter equation gives
us a solution <math>x = a/c</math>, <math>y = b/c</math> to the former. It is also the
same as asking for all points with rational coordinates on the curve
described by <math>x^2 + y^2 = 1</math>. (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 <math>f(x,y)=0</math>, where <math>f</math> is a polynomial in two variables - turns out to depend crucially on the ''genus'' of the curve. The ''genus'' can be defined as follows:<ref>It may be useful to look at an example here. Say we want to study the curve <math>y^2 = x^3 + 7</math>. We allow ''x'' and ''y'' to be complex numbers: <math>(a + b i)^2 = (c + d i)^3 + 7</math>. This is, in effect, a set of two equations on four variables, since both the real
===Nineteenth-century developments===
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
[[Augustin Louis Cauchy]], [[Louis Poinsot]] (1845), and notably [[Charles Hermite]] have added to the subject. In the theory of ternary forms, [[Gotthold Eisenstein]] has been a leader, and to him and [[H. J. S. Smith]] is also due a noteworthy advance in the theory of forms in general. Smith gave a complete classification of ternary quadratic forms, and extended Gauss's researches concerning real [[quadratic form]]s to complex forms. The investigations concerning the representation of numbers by the sum of 4, 5, 6, 7, 8 squares were advanced by Eisenstein and the theory was completed by Smith.
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.</ref> allow the variables in <math>f(x,y)=0</math> to be complex numbers; then <math>f(x,y)=0</math> defines a 2-dimensional surface in (projective) 4-dimensional space (since two complex variables can be decomposed into four real variables, i.e., four dimensions). Count
the number of (doughnut) holes in the surface; call this number the ''genus'' of <math>f(x,y)=0</math>. Other geometrical notions turn out to be just as crucial.


There is also the closely linked area of [[diophantine approximations]]: given a number <math>x</math>, 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 <math>a/q</math> (with <math>gcd(a,q)=1</math>) a good approximation to <math>x</math> if <math>\scriptstyle |x-a/q|<\frac{1}{q^c}</math>, where <math>c</math> is large.) This question is of special interest if <math>x</math> is an algebraic number. If <math>x</math> 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.
[[Johann Peter Gustav Lejeune Dirichlet|J. P. G. Lejeune Dirichlet]] was the first to lecture upon the subject in a German university. Among his contributions is the proof of [[Fermat's Last Theorem]]:
:<math>x^n+y^n \neq z^n, (x,y,z \neq 0, n > 2)</math>


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. ''Arithmetic geometry'', on the other hand, is a contemporary term
for the cases ''n'' = 5 and ''n'' = 14 (Euler and Legendre had already proved the cases ''n'' = 3 and ''n'' = 4 and therefore by implication, all multiples of 3 and 4). Among the later French writers are [[Émile Borel]]; [[Henri Poincaré]], whose memoirs are numerous and valuable; [[Jules Tannery]], and [[Thomas Jan Stieltjes]]. Among the leading contributors in Germany were [[Leopold Kronecker]], [[Ernst Kummer]], [[Ernst Christian Julius Schering]], [[Paul Bachmann]], and [[Richard Dedekind]]. In Austria [[Otto Stolz]]'s ''Vorlesungen über allgemeine Arithmetik ''(1885–86), and in England [[George Ballard Mathews]]' Theory of Numbers (Part I, 1892) were scholarly general works. [[Angelo Genocchi]], [[James Joseph Sylvester]], and [[J. W. L. Glaisher]] have also added to the theory.
for much the same domain as that covered by the term ''diophantine geometry''. The term ''arithmetic geometry'' is arguably used
most often when one wishes to emphasise the connections to modern algebraic geometry (as in, for instance, [[Faltings' theorem]]) rather than to techniques in diophantine approximations.


===Recent approaches and subfields===
===Late nineteenth- and early twentieth-century developments===
It was the time of major advancements in number theory due to the work of [[Axel Thue]] on [[Diophantine equation]]s, of [[David Hilbert]] in [[algebraic number theory]] and the solution of [[Waring's problem]], and to the creation of [[geometric number theory]] by [[Hermann Minkowski]], but also thanks to [[Adolf Hurwitz]], [[Georgy F. Voronoy]], [[Waclaw Sierpinski]], [[Derrick Norman Lehmer]] and several others.


The areas below date as such from no earlier than the mid-twentieth century, even if they are based
===Twentieth-century developments===
on older material. For example, as is explained below, the matter of algorithms in number theory is very old, in some sense older than the concept of proof; at the same time, the modern study of computational complexity dates only from the 1930s and 1940s.
Major figures in twentieth-century number theory include [[Hermann Weyl]], [[Nikolai Chebotaryov]], [[Emil Artin]], [[Erich Hecke]], [[Helmut Hasse]], [[Alexander Gelfond]], [[Yuri Linnik]], [[Paul Erdős]], [[Gerd Faltings]], [[G. H. Hardy]], [[Edmund Landau]], [[Louis Mordell]], [[John Edensor Littlewood]], [[Ivan Niven]], [[Srinivasa Ramanujan]], [[André Weil]], [[Ivan Vinogradov]], [[Atle Selberg]], [[Carl Ludwig Siegel]], [[Igor Shafarevich]], [[John Tate]], [[Robert Langlands]], [[Goro Shimura]], [[Kenkichi Iwasawa]], [[Jean-Pierre Serre]], [[Pierre Deligne]], [[Enrico Bombieri]], [[Alan Baker (mathematician)|Alan Baker]], [[Peter Swinnerton-Dyer]], [[Bryan John Birch]], [[Vladimir Drinfel'd]], [[Laurent Lafforgue]], [[Andrew Odlyzko]], [[Andrew Wiles]], and [[Richard Taylor (mathematician)|Richard Taylor]], [[Henryk Iwaniec]].


Milestones in twentieth-century number theory include:
====Probabilistic number theory ====
*The development of class field theory, its completion—[[Teiji Takagi]], Emil Artin, [[Philipp Furtwängler]] in the 1920s—and its extensions and reformulations— Helmut Hasse, [[Claude Chevalley]] in the 1930s.
*The [[Weil conjectures]] introduced by André Weil in the 1940s, and their proof in the work of [[Bernard Dwork]], [[Alexander Grothendieck]], Pierre Deligne, and others.
*Barban's theorem in 1961 and its 1965 refinement, the [[Bombieri–Vinogradov theorem]].
*The creation of the [[Langlands program]] by Robert Langlands in the late 1960s, and subsequent progress by many mathematicians.
*[[Chen's theorem]] stated in 1966 and proved in 1973
*The proof of [[Fermat's Last Theorem]] by Andrew Wiles, and the proof of the full [[Taniyama–Shimura conjecture]] in 1999 by [[Christophe Breuil]], [[Brian Conrad]], [[Fred Diamond]], and Richard Taylor.


{{main|Probabilistic number theory}}
Notable current work - that is, new results proven in the first decade of the twentieth-first century - includes progress by Clozel, Taylor, Harris, Shepherd-Barron and others on the [[Sato-Tate conjecture]], the results by Green, Tao, Ziegler and many others on linear forms in the primes - based on [[Timothy Gowers|Gowers]]'s work - and Goldston, Pintz and Yildirim's theorem on small gaps between primes.


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?
==Applied number theory==
The book ''Number theory for computing''<ref>[http://books.google.co.uk/books?id=lIvPz7k41SEC Number theory for computing]</ref> says that number theory has been applied to physics<ref>See, for example, ''Number theory and physics archive'', http://empslocal.ex.ac.uk/people/staff/mrwatkin/zeta/physics.htm</ref>, chemistry, biology, computing, engineering, coding and cryptography, random number generation, acoustics, communications, graphic design and even music and business. It also says that [[Shiing-Shen Chern]] "considers number theory as a branch of applied mathematics because of its strong applicability in other fields."


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 [[statistical independence|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.
===Public-key cryptography===
Many public-key cryptography schemes use number theory; e.g., [[RSA Cryptosystem]], [[Elliptic curve cryptography]].


It is sometimes said that probabilistic combinatorics uses the fact that whatever happens with probability greater than <math>0</math> 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.
===Residue number system===
A [[residue number system]] (RNS) represents a large integer using a set of smaller integers, so that computation may be performed more efficiently. It relies on the [[Chinese remainder theorem]] of [[modular arithmetic]] for its operation. RNS have applications in the field of digital computer arithmetic. By decomposing in this a large integer into a set of smaller integers, a large calculation can be performed as a series of smaller calculations that can be performed independently and in parallel. Because of this, it is particularly popular in hardware implementations.


====Arithmetic combinatorics====
==Quotations==
*"Mathematics is the queen of the sciences and number theory is the queen of mathematics."&nbsp;— [[Carl Friedrich Gauss|Gauss]]<ref>Quoted in ''Gauss zum Gedächtniss'' (1856) by Wolfgang Sartorius von Waltershausen</ref>
*"God invented the integers; all else is the work of man."&nbsp;— [[Leopold Kronecker|Kronecker]]<ref>"Die ganzen Zahlen hat der liebe Gott gemacht, alles andere ist Menschenwerk" Heinrich Weber: ''Leopold Kronecker. Jahresberichte'' D.M.V 2 (1893) 5-31</ref>
*"Number is the within of all things."&nbsp;— Attributed to [[Pythagoras]]<ref>L.A. Michael: The Principles of Existence & Beyond (Dec 2007) ISBN 1847991998, ISBN 978-1847991997</ref>


{{main|Arithmetic combinatorics}}
==Notes==
{{Reflist}}


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

*{{Apostol IANT}}
If we begin from a fairly "thick" infinite set <math>A</math>, does it contain many elements in arithmetic progression: <math>a</math>,
*{{Cite book| author=Dedekind, Richard | title=Essays on the Theory of Numbers | publisher=Cambridge University Press | year=1963 | isbn=0-486-21010-3}}
<math>a+b</math>, <math> a+2 b</math>, <math>a+3 b</math>, ... , <math>a+10b</math>, say? Should it be possible to write large integers as sums of elements of <math>A</math>?
*{{Cite book| author=Davenport, Harold | title=The Higher Arithmetic: An Introduction to the Theory of Numbers (7th ed.) | publisher=Cambridge University Press | year=1999 | isbn=0-521-63446-6}}

*{{Cite book| author=Friedberg, Richard | title=An Adventurer's Guide to Number Theory | publisher=McGraw-Hill | year=1968 | location=New York | isbn=0486281337 | url = http://books.google.com/books?id=2alNGWQIOjAC&printsec=frontcover}}
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 <math>A</math> of arithmetic significance, such as the primes or the squares) and, arguably, some of the ''geometry of numbers'',
*{{Cite book| author=Guy, Richard K. | title=Unsolved Problems in Number Theory | publisher=Springer-Verlag | year=1981 | isbn=0-387-90593-6}}
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 <math>A</math> 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 [[ring|rings]], in which case the growth of <math>A+A</math> and <math>A</math>·<math>A</math> may be
*{{Cite book| author=Hardy, G. H. and Wright, E. M. | title=An Introduction to the Theory of Numbers (5th ed.) | publisher=Oxford University Press | year=1980 | isbn=0-19-853171-0}}
compared.
*{{Cite book| author=Niven, Ivan, Zuckerman, Herbert S. and Montgomery, Hugh L. | title=An Introduction to the Theory of Numbers (5th ed.) | publisher=Wiley Text Books | year=1991 | isbn=0-471-62546-9}}

*{{Cite book| authorlink=Oystein Ore |author=Ore, Oystein | title=Number Theory and Its History | publisher=Dover Publications, Inc. | year=1948 | isbn=0-486-65620-9}}
====Computations in number theory====
*Smith, David. [http://www.gutenberg.org/dirs/etext05/hsmmt10p.pdf ''History of Modern Mathematics'' (1906)] (adapted public domain text)

*Dutta, Amartya Kumar (2002). 'Diophantine equations: The Kuttaka', [http://www.iisc.ernet.in/academy/resonance/Oct2002 ''Resonance - Journal of Science Education''].
{{main|Computational number theory}}
*O'Connor, John J. and Robertson, Edmund F. (2004). [http://turnbull.mcs.st-and.ac.uk/~history/Indexes/Arabs.html 'Arabic/Islamic mathematics'], ''[[MacTutor History of Mathematics archive]]''.

*O'Connor, John J. and Robertson, Edmund F. (2004). [http://turnbull.mcs.st-and.ac.uk/~history/Indexes/Indians.html 'Index of Ancient Indian mathematics'], ''MacTutor History of Mathematics archive''.
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.
*O'Connor, John J. and Robertson, Edmund F. (2004). [http://turnbull.mcs.st-and.ac.uk/~history/Indexes/Number_Theory.html 'Numbers and Number Theory Index'], ''MacTutor History of Mathematics archive''.

*[[List of important publications in mathematics#Number theory|Important publications in number theory]]
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 [[primality test|testing primality]], but, in spite of much work, no truly fast algorithm for factoring.
*{{Citation

| first=Carl B.
The difficulty of a computation can be useful: modern protocols
| last=Boyer
for [[cryptography|encrypting messages]] (e.g., [[RSA]])
| authorlink=Carl Benjamin Boyer
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
| title=A History of Mathematics
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.
| edition=Second Edition
| publisher=John Wiley & Sons, Inc.
On a different note - some things may not be computable at all; in fact, this can be proven. For instance, [[Alan Turing|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
| year=1991
was proven that there is [[Hilbert's 10th problem|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
| isbn=0471543977
the following two statements are both true: there are no solutions, and we shall never know that there are no solutions.
}}

== References ==

<references/>

{{Citizendium}}


==External links==
==External links==
{{wikibooks|Discrete mathematics|Number theory}}
{{Portal|Number theory}}
{{Portal|Number theory}}
* [http://www.numbertheory.org Number Theory Web]
* [http://www.numbertheory.org Number Theory Web]
* [http://www.shoup.net/ntb/ A Computational Introduction to Number Theory and Algebra] by Victor Shoup
* [http://www.math.niu.edu/~rusin/known-math/index/11-XX.html The Mathematical Atlas - 11: Number theory]


{{Mathematics-footer}}
{{Mathematics-footer}}
{{Number theory-footer}}
{{Number theory-footer}}


<!--
{{DEFAULTSORT:Number Theory}}
do not use categories on user pages
[[Category:Number theory| ]]
[[Category:Number theory| ]]

-->


[[ar:نظرية الأعداد]]
[[ar:نظرية الأعداد]]
[[an:Teoría de numers]]
[[an:Teoría de numeros]]
[[ast:Teoría de númberos]]
[[ast:Teoría de númberos]]
[[bn:সংখ্যাতত্ত্ব]]
[[bn:সংখ্যাতত্ত্ব]]
Line 195: Line 262:
[[el:Θεωρία αριθμών]]
[[el:Θεωρία αριθμών]]
[[es:Teoría de números]]
[[es:Teoría de números]]
[[eo:Nombroteorio]]
[[eu:Zenbaki-teoria]]
[[fa:نظریه اعداد]]
[[hif:Number theory]]
[[fr:Théorie des nombres]]
[[gl:Teoría dos números]]
[[gan:數論]]
[[ko:정수론]]
[[hi:संख्या सिद्धान्त]]
[[hr:Teorija brojeva]]
[[id:Teori bilangan]]
[[ia:Theoria de numeros]]
[[is:Talnafræði]]
[[it:Teoria dei numeri]]
[[he:תורת המספרים]]
[[jv:Téori angka]]
[[ka:რიცხვთა თეორია]]
[[la:Theoria numerorum]]
[[lv:Skaitļu teorija]]
[[lb:Zuelentheorie]]
[[lt:Skaičių teorija]]
[[jbo:nacycmaci]]
[[hu:Számelmélet]]
[[ml:സംഖ്യാസിദ്ധാന്തം]]
[[mt:Teorija tan-numri]]
[[mr:अंक प्रवाद]]
[[ms:Teori nombor]]
[[mn:Тооны онол]]
[[nl:Getaltheorie]]
[[ja:数論]]
[[no:Tallteori]]
[[nn:Talteori]]
[[oc:Teoria dels nombres]]
[[pnb:نمبر تھیوری]]
[[pms:Teorìa dij nùmer]]
[[nds:Tallentheorie]]
[[pl:Teoria liczb]]
[[pt:Teoria dos números]]
[[ro:Teoria numerelor]]
[[ru:Теория чисел]]
[[scn:Tiurìa dî nùmmura]]
[[simple:Number theory]]
[[sk:Teória čísel]]
[[sl:Teorija števil]]
[[sr:Теорија бројева]]
[[sh:Teorija brojeva]]
[[fi:Lukuteoria]]
[[sv:Talteori]]
[[ta:எண் கோட்பாடு]]
[[th:ทฤษฎีจำนวน]]
[[tr:Sayılar teorisi]]
[[tk:Sanlar teoriýasy]]
[[uk:Теорія чисел]]
[[ur:نظریۂ عدد]]
[[vi:Lý thuyết số]]
[[vo:Numateor]]
[[fiu-vro:Arvoteooria]]
[[war:Teyorya han ihap]]
[[zh:数论]]

Revision as of 09:50, 10 October 2011

Number theory is a branch of pure 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 prime numbers (the building blocks of integers) 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 in themselves or as solutions to equations (diophantine geometry). Questions in number theory are often best understood through the study of analytical objects (e.g., the Riemann zeta function) that encode properties of the integers, primes or other number-theoretic objects in some fashion (analytic 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.)

History

The beginnings

While there are elements of what in retrospect can be seen as number theory in Babylonian and ancient Chinese mathematics (see Plimpton 322 and the Chinese Remainder Theorem, respectively), the history of number theory truly starts with the Greek and Indian traditions.

The irrationality of is credited to the early Pythagoreans.[3] Euclid gave an algorithm for computing the greatest common divisor of two numbers (Euclid's Elements, Prop. VII.2) and a proof that there are infinitely many primes (Elements, Prop. IX.20). Much later in the Hellenistic period, Diophantus studied rational solutions to equations and systems of equations.

Results in number theory within Indian mathematics date from the period that would correspond to the medieval era in Europe. Aryabhata gave an algorithm for solving[4] pairs of congruences , apparently with astronomical applications in mind.[5] Brahmagupta started the systematic study of indefinite quadratic equations, including what would later be misnamed Pell's equation. A general procedure (the chakravala, or "cyclic method") for solving Pell's equation was finally found by Jayadeva (cited in the eleventh century; his work is otherwise lost); the earliest surviving exposition appears in Bhāskara II's Bīja-gaṇita (twelfth century).[6] Unfortunately, these achievements were largely unknown in the West until the late eighteenth century.[7]

Much Greek mathematics and some Indian mathematics was available to Arabic scholars from the early ninth century onwards. Part of the treatise al-Fakhri (by al-Karajī, 953 - ca. 1029) builds on Diophantus's work to some extent.

Modern number theory

Modern number theory begins with Pierre de Fermat, inspired in part by his study of Diophantus. Continuous activity on the subject started almost a century later with Euler.[8] Lagrange provided proofs of some of Fermat's and Euler's key statements. He and Legendre also set the basis of the study of quadratic forms; Legendre was the first to state the law of quadratic reciprocity. In Disquisitiones Arithmeticae, Gauss gave the first valid proof of this law, developed the theory of quadratic forms further, and started the modern study of cyclotomy.

Starting early in the nineteenth century, the following developments gradually took place:

  • The rise to self-consciousness of number theory (or higher arithmetic) as a field of study.[9]
  • The development of much of modern mathematics necessary for basic modern number theory: complex analysis, group theory, Galois theory -- accompanied by greater rigor in analysis and abstraction in algebra.
  • The rough subdivision of number theory into its modern subfields - in particular, analytic and algebraic number theory.

Algebraic number theory may be said to start with the study of reciprocity and cyclotomy, but truly came into its own with the development of abstract algebra and early ideal theory and valuation theory; see below. An obvious conventional starting point for analytic number theory would be Riemann's memoir on the Riemann zeta function (1859); there is also Dirichlet's theorem on arithmetic progressions, which preceded it in the study of the zeta function (for ), or Jacobi's work on the four square theorem, which connected arithmetical questions with elliptic functions. The first use of analytical arguments in number theory goes further back, to Euler.[10]

The history of each subfield is sketched in its own section below. Many of the most interesting questions in each area remain open and are being actively worked on.

Approaches and subfields

Introductory texts and elementary tools

Two of the most popular introductions to the subject are:

  • G. H. Hardy and E. M. Wright, An introduction to the theory of numbers, 6th ed., rev. by D. R. Heath-Brown and J. H. Silverman, Oxford University Press, Oxford, 2008 (first published in 1938).
  • I. M. Vinogradov, Elements of Number Theory, Mineola, NY: Dover Publications, 2003, reprint of the 1954 edition.

Hardy and Wright's book is a comprehensive classic, though its clarity sometimes suffers due to the authors' insistence on elementary methods.[11] Vinogradov's main attraction consists in its set of problems, which quickly lead to Vinogradov's own research interests; the text itself is very basic and close to minimal.

The term elementary generally denotes a method that does not use complex analysis. For example, the prime number theorem was first proven in 1896, but an elementary proof was found only in 1949. The term is somewhat ambiguous: for example, proofs based on Tauberian theorems are often seen as quite enlightening but not elementary, in spite of using Fourier analysis, rather than complex analysis as such. Here as elsewhere, an elementary proof may be longer and more difficult for most readers than a non-elementary one.

Number theory has the reputation of being a field many of whose results can be stated to the layperson. At the same time, the proofs of these results are not particularly accessible, in part because the range of tools they use is, if anything, unusually broad within mathematics.

Popular choices for a second textbook include Borevich and Shafarevich's Number theory and Serre's Cours d'arithmetique. Textbooks for later stages in one's study tend to branch into analytic and algebraic number theory, among other subfields.

Main fields

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.

Algebraic number theory

Algebraic number theory studies algebraic properties and algebraic objects of interest in number theory. (Thus, analytic and algebraic number theory can and do overlap: the former is defined by its methods, the latter by its objects of study.) A key topic 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 with rational coefficients; for example, every solution of (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 , where and are rational numbers and 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 , the number can be factorised both as and ; all of , , and are irreducible, and thus, in a naïve sense, analogous to primes among the integers.) The initial impetus for the development of ideal numbers (by Kummer) seems to have come from the study of higher reciprocity laws,[12] i.e., generalisations of quadratic reciprocity.

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[13] 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.

Diophantine geometry

The central problem of Diophantine geometry is to determine when a Diophantine equation has solutions, and if it does, how many. The approach taken is to think of the solutions of an equation as a geometric object.

For example, 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. In Diophantine geometry, one asks whether there are any rational points (points all of whose coordinates are rationals) or integral points (points all of whose coordinates are integers) on the curve or surface. If there are any such points, the next step is to ask how many there are and how they are distributed. A basic question in this direction is: 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 ; we would like to study its rational solutions, i.e., its solutions such that x and y are both rational. This is the same as asking for all integer solutions to ; any solution to the latter equation gives us a solution , to the former. It is also the same as asking for all points with rational coordinates on the curve described by . (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 , where is a polynomial in two variables - turns out to depend crucially on the genus of the curve. The genus can be defined as follows:[14] allow the variables in to be complex numbers; then defines a 2-dimensional surface in (projective) 4-dimensional space (since two complex variables can be decomposed into four real variables, i.e., four dimensions). Count the number of (doughnut) holes in the surface; call this number the genus of . Other geometrical notions turn out to be just as crucial.

There is also the closely linked area of diophantine approximations: given a number , 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 (with ) a good approximation to if , where is large.) This question is of special interest if is an algebraic number. If 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. Arithmetic geometry, on the other hand, is a contemporary term for much the same domain as that covered by the term diophantine geometry. The term arithmetic geometry is arguably used most often when one wishes to emphasise the connections to modern algebraic geometry (as in, for instance, Faltings' theorem) rather than to techniques in diophantine approximations.

Recent approaches and subfields

The areas below date as such from no earlier than the mid-twentieth century, even if they are based on older material. For example, as is explained below, the matter of algorithms in number theory is very old, in some sense older than the concept of proof; at the same time, the modern study of computational complexity dates only from the 1930s and 1940s.

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 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.

Arithmetic combinatorics

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

If we begin from a fairly "thick" infinite set , does it contain many elements in arithmetic progression: , , , , ... , , say? Should it be possible to write large integers as sums of elements of ?

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 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 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 and · may be compared.

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 (e.g., RSA) 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.

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. ^ 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. Plato reports on further work by Theodorus on irrationality.
  4. ^ 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).
  5. ^ K. Plofker, Mathematics in India, Princeton University Press, 2008, p. 119.
  6. ^ Plofker, op. cit., p. 194
  7. ^ Plofker, op. cit., p. 283
  8. ^ A. Weil, 'Number theory: an approach through history - from Hammurapi to Legendre, Birkhäuser, 1984, pp. 1-2.
  9. ^ See the discussion in section 5 of C. Goldstein and N. Schappacher, "A book in search of a discipline (1801-1860)', in C. Goldstein, N. Schappacher and J. Schwermer (eds.), "The shaping of arithmetic after C. F. Gauss's Disquisitiones Arithmeticae", Springer, 2007. Early signs of self-consciousness are present already in letters by Fermat: thus his remarks on what number theory is, and how "Diophantus's work [...] does not really belong to [it]" (quoted in A. Weil, op. cit., p. 25).
  10. ^ H. Iwaniec and E. Kowalski, Analytic number theory, AMS Colloquium Pub., Vol. 53, 2004, p. 1.
  11. ^ T. M. Apostol, Review of An introduction to the theory of numbers, Mathematical Reviews, MR0568909.
  12. ^ H. M. Edwards, Fermat's Last Theorem: a genetic introduction to algebraic number theory, Springer Verlag, 1977, p. 79.
  13. ^ 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.)
  14. ^ It may be useful to look at an example here. Say we want to study the curve . We allow x and y to be complex numbers: . 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.

This article incorporates material from the Citizendium article "Number theory", which is licensed under the Creative Commons Attribution-ShareAlike 3.0 Unported License but not under the GFDL.

External links