Number Theory

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 10

number theory

Go to navigationGo to search

Our theory of numbers is derived from the ancient Greek arithmetic of Diophantus.1 Cover of the arithmetic of
Diophantus translated into Latin by Bachet de Méziriac, edition with comments by Pierre de Fermat published in
1670.

The distribution of prime numbers is a central point of study in number theory. This Ulam spiral serves to illustrate
this, hinting, in particular, at the conditional independence between being prime and being a value of certain
quadratic polynomials.
Number theory is the branch of mathematics that studies the properties of numbers, particularly integers, but
more generally, it studies the properties of rings of numbers: integer rings containing {\displaystyle \mathbb {Z} }\
mathbb{Z} through a finite and injective morphism {\displaystyle \mathbb {Z} \hookrightarrow A}{\displaystyle \
mathbb {Z} \hookrightarrow A}. It contains a considerable number of problems that could be understood by "non-
mathematicians." More generally, this field studies the problems that arise with the study of integers. As Jürgen
Neukirch quotes:

Number theory occupies an idealized position among mathematical disciplines analogous to that occupied by
mathematics itself among other sciences.2
Integers can be considered in themselves or as solutions of equations (Diophantine geometry). Issues in number
theory are often best understood through the study of analytic objects (e.g., the Riemann zeta function) that
encode properties of integers, primes, or other objects of number theory. somehow (Analytic number theory).
Real numbers can also be studied in relation to rational numbers, for example, as an approximation of the latter
(Diophantine approximation).

The term "arithmetic" was also used to refer to number theory. This is a fairly old term, although not so popular
anymore. Hence number theory is usually called high arithmetic,3 although the term has also fallen into disuse.
This sense of the term arithmetic should not be confused with elementary arithmetic, or with the branch of logic
that studies Peano arithmetic as a formal system. Mathematicians who study number theory are called number
theorists.

Index
1 History
1.1 origins
1.1.1 Dawn of Arithmetic
1.1.2 Classical Greece and the early Hellenistic period
1.1.3 Diophantus of Alexandria
1.1.4 Āryabhaṭa, Brahmagupta, Bhāskara
1.1.5 Arithmetic in the Islamic Golden Age
1.1.6 Europe Western in the Middle Ages
2 Fields
2.1 Theory elementary of numbers
2.2 Theory number analysis
2.3 Theory of additive numbers
2.4 Theory algebraic of numbers
2.5 Theory geometric numbers
2.6 Combinatorial number theory
2.7 Computational number theory

3 History
4 See also
5 Grades
6 References
7 Bibliography
8 external links
History
origins
The Plimpton 322 tablet
Dawn of Arithmetic
The oldest historical find of an arithmetic nature is a fragment of a table: the broken clay tablet Plimpton 322
(Larsa, Mesopotamia, ca. 1800 BC) contains a list of "Pythagorean triples", that is, integers {\displaystyle (a,b,c)}
{\displaystyle (a,b,c)} such that {\displaystyle a^{2}+b ^{2}=c^{2}} {\displaystyle a^{2}+b^{2}=c^{2}}. The triples are
too many and too big to have been obtained by brute force. The title above the first column reads: "The takiltum
of the diagonal that has been subtracted such that the width..."4

The arrangement of the table suggests5 that it was constructed by what amounts, in modern language, to the
identity

{\displaystyle \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},}{\displaystyle \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 the routine exercises of ancient Babylon.6 If some other method was used,7 the triples were
constructed first and then rearranged by {\displaystyle c/a}{\displaystyle c/a}, presumably for use real as "table",
for example, with views to applications.

It is not known what these applications could have been, or if there could have been any; Babylonian astronomy,
for example, really developed only later. It has been suggested instead that the table was a source of numerical
examples for school problems,89 which is controversial. Robson's article is written polemically 10 in order to
"perhaps [...] knock [Plimpton 322] off his pedestal" 11; At the same time, he settles on the conclusion that:
[...] the question "how was the table calculated?" does not have to have the same answer as the question "what
problems does the table raise?" The first can be answered more satisfactorily by reciprocal pairs, as first
suggested half a century ago, and the second by some form of right triangle problems 10.
Robson takes issue with the idea that the scribe who produced Plimpton 322, who had to "work for a living", and
would not have belonged to a "well-to-do middle class", could be motivated by his own "idle curiosity" in the
absence of a "market for new mathematics."12

While Babylonian number theory - or what survives of Babylonian mathematics can be called that - consists of
this single striking fragment, Babylonian algebra (in the secondary sense of "algebra") was exceptionally well
developed.13 Late Neoplatonic Sources14 They claim that Pythagoras learned mathematics from the
Babylonians. Much earlier sources15 state that Thales and Pythagoras traveled and studied in Egypt.

Euclid IX 21-34 is most probably Pythagorean;16 it is very simple material ("odd times even is even", "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 show that {\displaystyle {\sqrt {2}}}{\sqrt {2}} is an irrational.17 The Pythagorean mystics gave great
importance to odd and even numbers.18The discovery that {\displaystyle {\sqrt {2}}}{\sqrt {2}} is irrational is
attributed to the early Pythagoreans (pre-Theodorus).19 By revealing (in modern terms) that numbers could be
irrational, this discovery seems having caused the first founding crisis in the history of mathematics; Its
demonstration or disclosure is sometimes attributed to Hypasus, who was expelled or split from the Pythagorean
sect.20 This forced us to distinguish between numbers (integers and rational numbers - the subjects of arithmetic
-), on the one hand, and lengths and proportions (which we would identify with real numbers, whether rational or
not), on the other.

The Pythagorean tradition also spoke of the so-called polygonal or figurative numbers.21 While square, cubic,
etc. numbers are now seen as more natural than triangular, pentagonal, etc. numbers, the study of the sums of
triangular and pentagonals would prove fruitful in the early modern era (17th century to early 19th century).

We do not know of any clearly arithmetical material in ancient Egyptian or Vedic sources, although there is some
algebra in both. The Chinese remainder theorem appears as an exercise 22 in Sunzi Suanjing (3rd, 4th, or 5th
centuries CE). 23 (There is an important step overlooked in Sunzi's solution: note 1 is the problem later solved by
Kuṭṭaka's Āryabhaṭa - see below).

There is also some numerical mysticism in Chinese mathematics,note 2 but, unlike that of the Pythagoreans, it
seems to have led nowhere. Like the perfect numbers of the Pythagoreans, magic squares have moved from
superstition to recreation.

Classical Greece and the early Hellenistic period


Main article: Greek mathematics
Apart from a few fragments, the mathematics of classical Greece is known to us either from the reports of
contemporary non-mathematicians or from the mathematical works of the early Hellenistic period.24 In the case
of number theory, this means, in general, Plato and Euclid, respectively.

Although Asian mathematics influenced Greek and Hellenistic learning, it appears that Greek mathematics is also
an indigenous tradition.

Eusebius of Caesarea, PE X, in chapter 4 mentions Pythagoras:

In fact, the said Pythagoras, while eagerly studying the wisdom of each nation, visited Babylon, and Egypt, and
all Persia, being instructed by the Magi and the priests: and in addition to these it is said that he studied with the
Brahmans (these are Indian philosophers). ); and from some he collected astrology, from others geometry, and
from others arithmetic and music, and different things from different nations, and only from the wise men of
Greece he obtained nothing, married as they were to a poverty and scarcity of wisdom: so, on the contrary, he
himself became the author of the instruction of the Greeks in the learning he had obtained from abroad. 25
Aristotle stated that Plato's philosophy closely followed the teachings of the Pythagoreans,26 and Cicero repeats
this statement: Platonem ferunt didicisse Pythagorea omnia ("They say that Plato learned everything
Pythagorean").27

Plato had a great interest in mathematics, and clearly distinguished between arithmetic and calculus. (By
arithmetic he meant, in part, theorizing about number, rather than what arithmetic or number theory has come to
mean.) It is through one of Plato's dialogues - namely, the Tetetus' - that we know that Theodore had shown that
{\displaystyle {\sqrt {3}},{\sqrt {5}},\dots ,{\sqrt {17}}}{\displaystyle {\sqrt {3}},{\sqrt {5}},\dots ,{\sqrt

{17}}} are irrational. Tetetus was, like Plato, a disciple of Theodore; He worked on the distinction of the different
types of incommensurables, so it could be said that he was a pioneer in the study of numerical systems. Book X
of Euclid's Elements is described by Pappus as based largely on the work of Theaetetus.

Euclid dedicated part of his Elements to prime numbers and divisibility, topics that unequivocally belong to
number theory and are basic to it (books VII to IX of Euclid's Elements). In particular, he gave an algorithm for
calculating the greatest common divisor of two numbers (Euclid's algorithm; Elements, Prop. VII.2) and the first
known proof of the infinity of prime numbers (Elements, Prop. IX.20).

In 1773, Lessing published an epigram that he had found in a manuscript during his work as a librarian; It was
intended to be a letter sent by Archimedes to Eratosthenes.2829 The epigram proposed what is known as
Archimedes' cattle problem; Its solution, absent in the manuscript, requires solving an indeterminate quadratic
equation, which reduces to what would later be mistakenly called Pell's equation. As far as we know, such
equations were first treated successfully by the Indian school. It is not known if Archimedes himself had a
solution method.

Diophantus of Alexandria

Cover of the 1621 edition of Diophantus's Arithmetica, translated into Latin by Claude Gaspard Bachet de
Méziriac.
Very little is known about Diophantus of Alexandria; He probably lived in the 3rd century AD, that is, about five
hundred years after Euclid. Six of the thirteen books of Diophantus' Arithmetic are preserved in the original Greek
and four more in an Arabic translation. Arithmetics is a collection of elaborate problems in which the task
invariably consists of finding rational solutions to a system of polynomial equations, usually of the form {\
displaystyle f(x,y)=z^{2}}{\displaystyle f (x,y)=z^{2}} or {\displaystyle f(x,y,z)=w^{2}}{\displaystyle f(x,y,z)=w^{2}}.
Thus, today, we speak of "Diophantine equations" when we talk about polynomial equations to which rational or
integer solutions must be found.

It can be said that Diophantus studied rational points, that is, points whose coordinates are rational, in curves and
algebraic varieties; However, unlike the Greeks of classical times, who did what we would today call basic
algebra in geometric terms, Diophantus did what we would today call basic algebraic geometry in purely
algebraic terms. In modern language, what Diophantus did was find rational parameterizations of varieties; that
is, given an equation of the form (say) {\displaystyle f(x_{1},x_{2},x_{3})=0}{\displaystyle f(x_{1},x_{2}, x_{3})=0},
his goal was to find (essentially) three rational functions {\displaystyle g_{1},g_{2},g_{3}} {\displaystyle g_{1},g_{2}
,g_{3}} such that, for all values of {\displaystyle r}r and {\displaystyle s}s, set

{\displaystyle x_{i}=g_{i}(r,s)}{\displaystyle x_{i}=g_{i}(r,s)} for {\displaystyle i=1,2,3}{\ displaystyle i=1,2,3} gives a


solution to {\displaystyle f(x_{1},x_{2},x_{3})=0.}{\displaystyle f(x_{1},x_{2) },x_{3})=0.}

Diophantus also studied the equations of some non-rational curves, for which a rational parameterization is not
possible. He managed to find some rational points on these curves (elliptic curve, in what appears to be their first
known appearance) by means of what amounts to a tangent construction: translated into coordinate geometry
(which did not exist in Diophantus' time), his method It would be visualized 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. (Diofanto also resorted to what might be called a special case of secant construction.)

Although Diophantus dealt largely with rational solutions, he assumed some results about integers, in particular
that every integer is the sum of four squares, although he never said so explicitly.
Āryabhaṭa, Brahmagupta, Bhāskara
Although Greek astronomy probably influenced Indian learning, to the point of introducing trigonometry,30 it
appears to be the case that Indian mathematics is otherwise an indigenous tradition;31 in particular, there is no
evidence that the Elements of Euclid reached India before the 18th century.32
Āryabhaṭa (476-550 AD) showed that pairs of simultaneous congruencies {\displaystyle nequiva _{1}{\bmod
{m}}_{1}}{\displaystyle nequiva _{1}{\bmod {m}}_{1 }}, {\displaystyle nequiva _{2}{\bmod {m}}_{2}}{\displaystyle
nequiva _{2} {\bmod {m}}_{2}} could be solved by a method called kuṭṭaka, or sprayer; 33 this is a procedure
close to (a generalization of) the Euclidean algorithm, which was probably discovered independently in India.34
Āryabhaṭa appears to have had applications to astronomical calculations in mind.30

Brahmagupta (628 AD) began the systematic study of indefinite quadratic equations - in particular, the misnamed
Pell's Equation, in which Archimedes may have first been interested, and which did not begin to be solved in the
West until the time of Fermat and Euler. Later Sanskrit authors would follow, using the technical terminology of
Brahmagupta. A general procedure (the chakravala, or "cyclic method") for solving Pell's equation was finally
found by Jayadeva (cited in the 11th century; his work is otherwise lost); the oldest extant exposition appears in
the Bīja-gaṇita of Bhāskara II (12th century).35

Indian mathematics remained largely unknown in Europe until the late 18th century;36 The work of Brahmagupta
and Bhāskara was translated into English in 1817 by Henry Colebrooke.37

Arithmetic in the Islamic Golden Age

Al-Haytham seen by the West: on the frontispiece of Selenographia Alhasen [sic] represents knowledge through
reason and Galileo represents knowledge through the senses.
In the early 9th century, the Caliph Al-Ma'mun ordered the translation of many Greek mathematical works and at
least one Sanskrit work (the Sindhind, which may or may not39 be Brāhmasphuṭasiddhānta's Brahmagupta).
Diophantus's main work, Arithmetic, was translated into Arabic by Qusta ibn Luqa (820-912). Part of the treatise
al-Fakhri (by al-Karajī, 953 - ca. 1029) is based on it to some extent. According to Rashed Roshdi, Al-Karajī's
contemporary Ibn al-Haytham knew40 what would later be called Wilson's theorem.

Western Europe in the Middle Ages


Apart from a treatise on squares in the arithmetic progression by Fibonacci - who traveled and studied in North
Africa and Constantinople - there was no number theory in Western Europe during the Middle Ages. Things
began to change in Europe at the end of the Renaissance, thanks to a renewed study of the works of Greek
antiquity. A catalyst was the textual amendment and Latin translation of Diophantus's Arithmetica.41

Fields
Depending on the methods used and the questions to be answered, number theory is subdivided into various
branches.

Elementary number theory


In elementary number theory, integers are studied without using techniques from other fields of mathematics.
Questions of divisibility, Euclid's algorithm to calculate the greatest common divisor, the factorization of integers
as products of prime numbers, the search for perfect numbers and congruences belong to elementary number
theory. Typical statements are Fermat's little theorem and Euler's theorem that extends it, the Chinese remainder
theorem and the law of quadratic reciprocity. In this branch, the properties of multiplicative functions such as the
Möbius function and the Euler φ function are investigated; as well as sequences of integers such as factorials
and F numbers.

Various questions within elementary number theory seem simple, but require very deep considerations and new
approaches, including the following:

Goldbach's conjecture that all even numbers (starting with 4) are the sum of two prime numbers.
Twin prime number conjecture about the infinity of the so-called twin prime numbers.
Fermat's Last Theorem (proved in 1995 by Andrew Wiles).
Riemann hypothesis on the distribution of the zeros of the Riemann zeta function, closely connected to the
problem of the distribution of prime numbers among others.
Analytical number theory
An analytic number theory uses calculus and complex analysis as tools to address questions about integers.
Some examples of this are the prime number theorem and the Riemann hypothesis. Waring's problem, the twin
prime conjecture, and the Goldbach conjecture are also being attacked through analytical methods.

Additive number theory


Additive number theory deals more deeply with the problems of number representation. Typical problems are
those already mentioned, Waring's problem and Goldbach's conjecture. This branch usually uses some results
referring to analytical number theory, such as the Hardy-Littlewood circle method, sometimes it is complemented
with sieve theory and in some cases topological methods are usually used.

Algebraic number theory


Algebraic number theory is a branch of number theory in which the concept of number is expanded to algebraic
numbers, which are the roots of polynomials with rational coefficients.

Geometric number theory


Geometric number theory (traditionally called number geometry) incorporates all forms of geometry. It begins
with Minkowski's theorem about common points on convex sets and investigations on spherical surfaces.

Combinatorial number theory


Combinatorial number theory deals with problems in number theory involving combinatorial ideas and their
formulations or solutions. Paul Erdős is the creator of this branch of number theory. Typical topics include
covered systems, zero-sum problems, various restricted sets, and arithmetic progressions on a set of integers.
Algebraic or analytical methods are quite powerful in this field.

Computational number theory


Computational number theory studies the relevant algorithms of number theory. Fast algorithms for evaluating
prime numbers and integer factorization have important applications in cryptography.

«The evolution of computing has made arithmetic stop being a contemplative and specialist science and become
a true applied branch. The need for new computing algorithms requires - as Enzo R. Gentile- vast and profound
arithmetic knowledge.

History
Mathematicians in India were interested in finding integer solutions to the Diophantine equations from the middle
of the 1st millennium BC. C. The first geometric use of Diophantine equations dates back to the Shulba-sutras,
which were written between the 5th and 3rd centuries BC. C. The religious Baudhaiana (in the 4th century BC.
C.) found two sets of positive integers to a set of simultaneous Diophantine equations, and simultaneous
Diophantine equations with more than four unknowns are also used. Apastamba (in the 3rd century BC). C.)
used simultaneous Diophantine equations with more than five unknowns.*

The Yaina mathematicians were the first to discard the idea that all infinities are the same or equal, but they had
already been studying them for years. They recognize five different types of infinities: infinity in one or two
directions (one-dimensional), infinity on surfaces (two-dimensional), infinity everywhere (three-dimensional), and
perpetually infinite (in an infinite number of dimensions).
Number theory was one of the favorite disciplines of study among the Greek mathematicians of Alexandria (in
Egypt) from the 3rd century BC. C., who were aware of the concept of the Diophantine equation in their particular
cases. The first Hellenistic mathematician to study these equations was Diophantus.

Diofanto investigated a method for finding integer solutions to indeterminate linear equations,42 equations in
which sufficient information is missing to produce a unique set of discrete answers. The equation x + y = 5 is an
example of them. Diophantus discovered that many indeterminate equations can be reduced to a form where a
certain category of solutions are known, even through a solution that is not.

The Diophantine equations were studied intensively by medieval Indian mathematicians, who were the first to
systematically search for methods for the determination of integer solutions. Ariabhata (476-550) gave the first
explicit description of the general integer solution of the linear Diophantine equation ay + bx = c, which appears in
his text Ariabhatíia. The kuttaka algorithm is considered one of Ariabhata's most significant contributions in pure
mathematics, which finds the integer solutions of a system of linear Diophantine equations, a problem of
important application in astronomy. Also find the general solution of the indeterminate linear equation using this
method.

Brahmagupta (598-668) worked on the most difficult Diophantine equations, which appear in his 18th book
dedicated to algebra and indeterminate equations. He used the chakravala method to solve the quadratic
Diophantine equations, including those of the form of Pell's equation such that 61x2 + 1 = y2. His Brahma-
sphuta-siddhanta was translated into Arabic in 773 and into Latin in 1126. The equation 61x2 + 1 = y2 was
proposed as a problem by the French mathematician Pierre de Fermat. The general solution to this particular
form of Pell's equation was found 70 years later by Leonhard Euler, although the general solution to Pell's
equation was found 100 years later by Joseph-Louis de Lagrange in 1767. However, several centuries earlier,
Pell's equation was worked out by Bhaskara II in 1150 using a modified version of Brahmagupta's chakravala
method, finding the general solution of other indeterminate intermediate quadratic equations and diophantine
quadratic equations. The Chakravala method for finding the general solution to Pell's equation was simpler than
the method used by Lagrange 600 years later. Bhaskara also finds the solution of other indeterminate quadratic,
cubic, quartic and polynomial equations of higher degrees. Naraian Pandit further refined the other indeterminate
quadratics for equations of higher degrees.

See also
Hilbert problems
Fibonacci sequence
golden ratio
Chinese remainder theorem
Sophie Germain
Grades
Sunzi Suanjing, chap. 3, problem 26, in Lam and Ang, 2004, pp. 219–20:
26] Now there are an unknown number of things. If we count in 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 in threes and there is a remainder of 2, write down 140. If we count by fives and have 3 left
over, write down 63. If we count by sevens and there is a 2 left over, we write down 30. Add to get 233 and
subtract 210 to get the answer. Yeah
We count by three and 1 is left over, we put 70. Yeah we count of five in five and
1 left over, score 21. If we count by sevens seven andone left over 1, we score 15. When
[a number] exceeds 106, the result is gets subtracting he 105.

See, for example, Sunzi Suanjing, chap. 3, problem 36, in Lam and Ang, 2004, pp. 223-24:
36] Now there is a pregnant woman whose age is 29 years old. If the gestation period is 9 months, determine the
sex of the unborn child. Answer: Male. Method: Enter 49, add the gestation period and subtract the age. From
the rest, remove 1 that represents the sky, 2 the earth, 3 the man, 4 the four seasons, 5 the five phases, 6 the six
trumpets, 7 the seven stars [of the Big Dipper], 8 the eight winds and 9 the nine divisions [of China under Yu the
Great]. If the rest is odd, [the sex] is male and if the rest is even, [the sex] is female.

This is the last problem in Sunzi's otherwise practical treatise.

References
Jean-Paul Collette (1985), History of mathematics (volumes 1 and 2). Translation by Alfonso Casal, Madrid:
Siglo XXI Editores SA ISBN 84-323-0526-4
Introduction to the work Cohomology of number fields:
Die Zahlentheorie nimmt unter den mathematischen Disziplinen eine ähnlich idealisierte Stellung ein wie die
Mathematik selbst unter den anderen Wissenschaften.
Davenport, Harold (1999). The Higher Arithmetic: An Introduction to the Theory of Numbers (7th edition).
Cambridge, England: Cambridge University Press. ISBN 0-521 63446-6.
Neugebauer & Sachs, 1945, p. 40. The term takiltum is problematic. Robson prefers the interpretation "The
diagonal support square from which the 1 is torn, so that the short side comes up...".Robson, 2001, p. 192
Robson, 2001, p. 189. Other sources give the modern formula {\displaystyle (p^{2}- q^{2},2pq,p^{2}+q^{2})}{\
displaystyle (p^{2}-q^{ 2},2pq,p^{2}+q^{2})}. Van der Waerden gives both the modern formula and what is
equivalent to Robson's preferred form. (van der Waerden, 1961, p. 79)
van der Waerden, 1961, p. 184.

Neugebauer (Neugebauer, 1969, pp. 36-40) discusses the table in detail and mentions in passing Euclid's
method in modern notation (Neugebauer, 1969, p. 39).
Friberg, 1981, p. 302.
Robson, 2001, p. 201.
Robson, 2001, p. 202.
Robson, 2001, p. 167.
Robson, 2001, pp. 199-200.
van der Waerden, 1961, p. 43.
Iamblichus, Life of Pythagoras, (trans., e.g., ) cited in . See also Porphyry, Life of Pythagoras, paragraph 6, in .
Van der Waerden (van der Waerden, 1961, pp. 87-90) supports the opinion that Thales knew Babylonian
mathematics.
Herodotus (II. 81) and Isocrates (Busiris 28), cited in: . On Thales, see Eudemus ap. Proclus, 65.7, (e.g.) cited
in: . Proclus used a work by Eudemus of Rhodes (now lost), the Catalog of Geometers. See also the introduction,
Morrow, 1992, p. xxx on the reliability of Proclus.
Becker, 1936, p. 533, cited in: van der Waerden, 1961, p. 108.
Becker, 1936.
van der Waerden, 1961, p. 109.
Plato, Theaetetus, p. 147 B, (e.g. Jowett, 1871), cited in von Fritz, 2004, p. 212: "Teodoro wrote to us
something about roots, like the roots of three or five, showing that they are incommensurable through unity;..."
See also Theodore's Spiral.
von Fritz, 2004.
Heath, 1921, p. 76.
Sunzi Suanjing, chapter 3, problem 26. It can be found in Lam and Ang, 2004, pp. 219-20, containing a
complete translation of the Suan Ching (based on Qian, 1963). See also the discussion in Lam and Ang, 2004,
pp. 138-140.
The date of the text has been reduced to 220-420 CE (Yan Dunjie) or 280-473 CE (Wang Ling) through internal
evidence (= taxation systems assumed in the text). See Lam and Ang, 2004, pp. 27-28.
Boyer and Merzbach, 1991, p. 82.
«Eusebius of Caesarea: Praeparatio Evangelica (Preparation for the Gospel). Tr. HEY Gifford (1903) - Book 10.
Metaphysics, 1.6.1 (987a)
Tusc. Dispute 1.17.39.
Vardi, 1998, pp. 305-19.
Weil, 1984, pp. 17-24.
Plofker, 2008, p. 119.
Any early contact between Babylonian and Indian mathematics remains conjectural (Plofker, 2008, p. 42).
Mumford, 2010, p. 387.
Āryabhatīya, Āryabhatīya, Chapter 2, verses 32-33, cited in: Plofker, 2008, pp. 134–40. See also Clark, 1930,
pp. 42-50. A slightly more explicit description of the kuṭṭaka was given later in Brahmagupta,
Brāhmasphuṭasiddhānta, XVIII, 3-5 (in Colebrooke, 1817, p. 325, cited in Clark, 1930, p. 42).
Mumford, 2010, p. 388.
Plofker, 2008, p. 194.
Plofker, 2008, p. 283.
Colebrooke, 1817.
Colebrooke, 1817, p. lxv, cited in Hopkins, 1990, p. 302. See also the preface in Sachau, 188 cited in Smith,
1958, pp. 168
Pingree, 1968, pp. 97-125, and Pingree, 1970, pp. 103-23, cited in Plofker, 2008, p. 256.
Rashed, 1980, pp. 305-21.
Bachet, 1621, after a first attempt by Xylander, 1575
"Brief history of number theory" Archived May 28, 2007 in
Wayback Machine., article on the Lycos website. Accessed June 5, 2007. Cite error: The <ref> tag defined in
<references> belongs to the group
«» not declared in the previous text.
Cite error: The <ref> tag defined in <references> belongs to the group "" not declared in the previous text.
Cite error: The <ref> tag defined in <references> belongs to the group "" not declared in the previous text.
Cite error: The <ref> tag defined in <references> belongs to the group "" not declared in the previous text.
Cite error: The <ref> tag defined in <references> belongs to the group "" not declared in the previous text.
Cite error: The <ref> tag defined in <references> belongs to the group "" not declared in the previous text.
Bibliography
Apostol, Tom M. (1976). Introduction to analytical number theory. Undergraduate Texts in Mathematics. Springer.
ISBN 978-0-387-90163-3. Accessed February 28, 2016.
Apostol, Tom M. (n.d.). An Introduction to the Theory of Numbers. (Review of Hardy & Wright.) Mathematical
Reviews (MathSciNet). American Mathematical Society. MR 0568909. Accessed February 28, 2016.
(Subscription needed)
Becker, Oskar (1936). «Die Lehre von Geraden und Ungeraden im neunten Buch der euklidischen Elemente».
Quellen und Studien zur Geschichte der Mathematik, Astronomie und Physik. Abteilung B:Studien (in German) 3:
533-53.
Boyer, Carl Benjamin; Merzbach, Uta C. (1991). A History of Mathematics (2nd edition). New York: Wiley. ISBN
978-0-471-54397-8. 1968 edition at archive.org Clark, Walter Eugene (trans.) (1930). The Āryabhaṭīya of
Āryabhaṭa: An ancient Indian work on Mathematics and Astronomy. University of Chicago Press. Accessed
February 28, 2016.
Colebrooke, Henry Thomas (1817). Algebra, with Arithmetic and Mensuration, from the Sanscrit of Brahmegupta
and Bháscara.. London: J. Murray. Accessed February 28, 2016.
Davenport, Harold; Montgomery, Hugh L. (2000). Multiplicative Number Theory.
Graduate texts in mathematics 74 (revised 3rd edition). Springer. ISBN 978-0-387 95097-6.
Edwards, Harold M. (November 1983). «Euler and Quadratic Reciprocity». Mathematics Magazine 56 (5): 285-
91. JSTOR 2690368. doi:10.2307/2690368.
Edwards, Harold M. (2000). Fermat's Last Theorem: a Genetic Introduction to Algebraic Number Theory.
Graduate Texts in Mathematics 50 (reprint of 1977 edition). Springer Verlag. ISBN 978-0-387-95002-0.
Fermat, Pierre de (1679). Varia Opera Mathematica (in French, la). Toulouse: Joannis Pech. Accessed February
28, 2016.
Friberg, Jöran (August 1981). "Methods and Traditions of Babylonian Mathematics: Plimpton 322, Pythagorean
Triples and the Babylonian Triangle Parameter Equations." Historia Mathematica 8 (3): 277-318.
doi:10.1016/0315-0860(81)90069-0.
von Fritz, Kurt (2004). "The Discovery of Incommensurability by Hippasus of Metapontum." In Christianidis, J.,
ed. Classics in the History of Greek Mathematics. Berlin: Kluwer (Springer). ISBN 978-1-4020-0081-2.
Gauss, Carl Friedrich; Waterhouse, William C. (trans.) (1966). Disquisitiones Arithmeticae. Springer. ISBN 978-0-
387-96254-2.
Goldfeld, Dorian M. (2003). "Elementary Proof of the Prime Number Theorem: a Historical Perspective."
Accessed February 28, 2016.
Goldstein, Catherine; Schappacher, Norbert (2007). «A book in search of a discipline». In Goldstein, C.;
Schappacher, N.; Schwermer, Joachim, eds. The Shaping of Arithmetic after CF Gauss's "Disquisitiones
Arithmeticae". Berlin & Heidelberg: Springer. pp. 3-66. ISBN 978-3-540-20441-1. Accessed February 28, 2016.
Granville, Andrew (2008). "Analytic number theory". In Gowers, Timothy; Barrow Green, June; Leader, Imre, eds.
The Princeton Companion to Mathematics. Princeton University Press. ISBN 978-0-691-11880-2. Accessed
February 28, 2016. Unknown parameter |title-link= ignored (help)
Porphyry; Guthrie, K.S. (trans.) (1920). Life of Pythagoras. Alpine, New Jersey:
Platonist Press.
Guthrie, Kenneth Sylvan (1987). The Pythagorean Sourcebook and Library. Grand Rapids, Michigan: Phanes
Press. ISBN 978-0-933999-51-0.
Hardy, Godfrey Harold; Wright, E.M. (2008). An Introduction to the Theory of Numbers (Sixth edition). Oxford
University Press. ISBN 978-0-19-921986-5. MR 2445243. Unknown parameter |title-link= ignored (help)
Heath, Thomas L. (1921). A History of Greek Mathematics, Volume 1: From Thales to Euclid. Oxford: Clarendon
Press. Accessed February 28, 2016.
Hopkins, J.F.P. (1990). «Geographical and Navigational Literature». In Young, M.J.L. Latham, J.D.; Serjeant,
R.B., eds. Religion, Learning and Science in the 'Abbasid Period. The Cambridge history of Arabic literature.
Cambridge University Press. ISBN 978-0-521-32763-3.
Huffman, Carl A. (8 August 2011). «Pythagoras». In Zalta, Edward N., ed. Stanford Encyclopaedia of Philosophy
(Fall 2011 edition). Consulted on February 7, 2012.
Iwaniec, Henryk; Kowalski, Emmanuel (2004). Analytic Number Theory. American Mathematical Society
Colloquium Publications 53. Providence, RI: American Mathematical Society. ISBN 978-0-8218-3633-0.
Dish; Jowett, Benjamin (trans.) (1871). Theaetetus.
Lam, Lay Yong; Ang, Tian Se (2004). Fleeting Footsteps: Tracing the Conception of Arithmetic and Algebra in
Ancient China (revised edition). Singapore: World Scientific. ISBN 978-981-238-696-0. Accessed February 28,
2016.
Long, Calvin T. (1972). Elementary Introduction to Number Theory (2nd edition). Lexington, VA: DC Heath and
Company. LCCN 77171950.
Mahoney, M.S. (1994). The Mathematical Career of Pierre de Fermat, 1601–1665 (Reprint, 2nd edition).
Princeton University Press. ISBN 978-0-691-03666-3. Accessed February 28, 2016.
Milne, J. S. (March 18, 2017). "Algebraic Number Theory" (3.07). Consulted on April 7, 2020.
Montgomery, Hugh L.; Vaughan, Robert C. (2007). Multiplicative Number Theory: I, Classical Theory. Cambridge
University Press. ISBN 978-0-521-84903-6. Accessed February 28, 2016.
Morrow, Glenn Raymond (trans., ed.); Proclus (1992). A Commentary on Book 1 of Euclid's Elements. Princeton
University Press. ISBN 978-0-691-02090-7.
Mumford, David (March 2010). «Mathematics in India: reviewed by David Mumford». Notices of the American
Mathematical Society 57 (3): 387. ISSN 1088-9477.
Neugebauer, Otto E. (1969). "The Exact Sciences in Antiquity." Acta Historica Scientiarum Naturalium et
Medicinalium 9 (corrected reprint of the 1957 edition) (New York: Dover Publications). pp. 1-191. ISBN 978-0-
486-22332-2. PMID 14884919. Accessed March 2, 2016.
Neugebauer, Otto E.; Sachs, Abraham Joseph; Götze, Albrecht (1945). Mathematical Cuneiform Texts.
American Oriental Series 29. American Oriental Society etc.
O'Grady, Patricia (September 2004). «Thales of Miletus». The Internet Encyclopaedia of Philosophy. Consulted
on February 7, 2012.
Pingree, David; Ya'qub, ibn Tariq (1968). "The Fragments of the Works of Ya'qub ibn Tariq." Journal of Near
Eastern Studies 26.
Pingree, D.; al-Fazari (1970). "The Fragments of the Works of al-Fazari." Journal of Near Eastern Studies 28.
Plofker, Kim (2008). Mathematics in India. Princeton University Press. ISBN 978-0 691-12067-6.
Qian, Baocong, ed. (1963). Suanjing shi shu (Ten Mathematical Classics) (in Chinese). Beijing: Zhonghua shuju.
Accessed February 28, 2016.
Rashed, Roshdi (1980). "Ibn al-Haytham et le théorème de Wilson." Archive for History of Exact Sciences 22 (4):
305-21. S2CID 120885025. doi:10.1007/BF00717654.
Robson, Eleanor (2001). "Neither Sherlock Holmes nor Babylon: a Reassessment of Plimpton 322." Historia
Mathematica 28 (3): 167-206. doi:10.1006/hmat.2001.2317. Archived from the original on October 21, 2014.
Sachau, Eduard; Bīrūni, ̄Muḥammad ibn Aḥmad (1888). Alberuni's India: An Account of the Religion, Philosophy,
Literature, Geography, Chronology, Astronomy and Astrology of India, Vol. 1. London: Kegan, Paul, Trench,
Trübner & Co. Accessed February 28, 2016.
Serre, Jean-Pierre (1996). A Course in Arithmetic. Graduate texts in mathematics 7. Springer. ISBN 978-0-387-
90040-7.
Smith, D.E. (1958). History of Mathematics, Vol I. New York: Dover Publications. Tannery, Paul; Henry, Charles
(eds.); Fermat, Pierre de (1891). Fermat's Oeuvres. (4 Vols.) (in fr, la). Paris: Imprimerie Gauthier-Villars et Fils.
Volume 1 Volume 2 Volume 3 Volume 4 (1912)
Iamblichus; Taylor, Thomas (trans.) (1818). Life of Pythagoras or, Pythagoric Life. London: J.M. Watkins.
Archived from the original on July 21, 2011. For other editions, see Iamblichus#List of editions and translations
Truesdell, CA (1984). "Leonard Euler, Supreme Geometer." In Hewlett, John (trans.), ed. Leonard Euler,
Elements of Algebra (reprint of 1840 5th edition). New York: Springer-Verlag. ISBN 978-0-387-96014-2. This
Google books preview of Elements of algebra lacks Truesdell's intro, which is reprinted (slightly abridged) in the
following book:
Truesdell, CA (2007). "Leonard Euler, Supreme Geometer." In Dunham, William, ed. The Genius of Euler:
reflections on his life and work. Volume 2 of MAA tercentenary Euler celebration. New York: Mathematical
Association of America. ISBN 978-0-88385 558-4. Accessed February 28, 2016.
Varadarajan, V.S. (2006). Euler Through Time: A New Look at Old Themes. American Mathematical Society.
ISBN 978-0-8218-3580-7. Accessed February 28, 2016. Vardi, Ilan (April 1998). «Archimedes' Cattle Problem».
American Mathematical Monthly 105 (4): 305-19. JSTOR 2589706. doi:10.2307/2589706. Unknown parameter |
citeseerx= ignored (help)
van der Waerden, Bartel L.; Dresden, Arnold (trans) (1961). Science Awakening. Vol. 1 or Vol 2. New York:
Oxford University Press.
Weil, André (1984). Number Theory: an Approach Through History – from Hammurapi to Legendre. Boston:
Birkhäuser. ISBN 978-0-8176-3141-3. Accessed February 28, 2016.

You might also like