Prime number: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Line 347: Line 347:
Fermat primes are primes of the form
Fermat primes are primes of the form
:{{nowrap|''F''<sub>''k''</sub> {{=}} 2<sup>2<sup>''k''</sup></sup> + 1}},
:{{nowrap|''F''<sub>''k''</sub> {{=}} 2<sup>2<sup>''k''</sup></sup> + 1}},
with ''k'' a natural number. They are named after [[Pierre de Fermat]], who conjectured that all such numbers are prime. This was based on the evidence of the first five numbers in this series—3, 5, 17, 257, and 65,537—being prime. However, ''F''<sub>5</sub> is composite and so are all other Fermat numbers that have been verified as of 2015. A [[regular polygon|regular ''n''-gon]] is [[constructible polygon|constructible using straightedge and compass]] if and only if the odd prime factors of ''n'' (if any) are distinct Fermat primes.
with ''k'' a natural number. They are named after [[Pierre de Fermat]], who conjectured that all such numbers are prime. This was based on the evidence of the first five numbers in this series—3, 5, 17, 257, and 65,537—being prime.<ref name="kls">{{cite book | last1 = Křížek | first1 = Michal | last2 = Luca | first2 = Florian | last3 = Somer | first3 = Lawrence | doi = 10.1007/978-0-387-21850-2 | isbn = 0-387-95332-9 | location = New York | mr = 1866957 | pages = 1–2 | publisher = Springer-Verlag | series = CMS Books in Mathematics | title = 17 Lectures on Fermat Numbers: From Number Theory to Geometry | url = https://books.google.com/books?id=hgfSBwAAQBAJ&pg=PA1 | volume = 9 | year = 2001}}</ref> However, ''F''<sub>5</sub> is composite and so are all other Fermat numbers that have been verified as of 2017.<ref>{{cite journal | last1 = Boklan | first1 = Kent D. | last2 = Conway | first2 = John H. | author2-link = John Horton Conway | arxiv = 1605.01371 | date = January 2017 | doi = 10.1007/s00283-016-9644-3 | issue = 1 | journal = [[The Mathematical Intelligencer]] | pages = 3–5 | title = Expect at most one billionth of a new Fermat prime! | volume = 39}}</ref> A [[regular polygon|regular ''n''-gon]] is [[constructible polygon|constructible using straightedge and compass]] if and only if the odd prime factors of ''n'' (if any) are distinct Fermat primes.<ref name="kls"/>


===Other mathematical occurrences of primes===
===Other mathematical occurrences of primes===

Revision as of 08:10, 24 January 2018

The prime numbers are the numbers greater than one that are not products of two smaller numbers

A prime number (or a prime) is a natural number greater than 1 that cannot be formed by multiplying two smaller natural numbers. A natural number greater than 1 that is not prime is called a composite number. For example, 5 is prime because the only ways of writing it as a product, 1 × 5 or 5 × 1, involve 5 itself. However 6 is composite because it is the product 2 × 3 of two numbers that are both smaller than 6. Primes are central in number theory because of the fundamental theorem of arithmetic: every natural number greater than 1 is either a prime itself or can be factorized as a product of primes that is unique up to their order. For this theorem to be true, 1 should not be prime because there are multiple ways of writing any number as a product of primes and 1: for instance, 3, 1 × 3, 1 × 1 × 3, etc. are all valid factorizations of 3.

The property of being prime is called primality. A simple but slow method of checking the primality of a given number n is known as trial division. It consists of testing whether n is a multiple of any integer between 2 and n. Algorithms much more efficient than trial division can test the primality of large numbers. These include the Miller–Rabin primality test, which is fast but has a small chance of error, and the AKS primality test, which always produces the correct answer in polynomial time but is too slow to be practical. Particularly fast methods are available for numbers of special forms, such as Mersenne numbers. As of January 2018, the largest known prime number has 23,249,425 decimal digits.

There are infinitely many primes, as demonstrated by Euclid around 300 BC. No known simple formula separates prime numbers from composite numbers. However, the distribution of primes, that is to say, the statistical behaviour of primes in the large, can be modelled. The first result in that direction is the prime number theorem, proven at the end of the 19th century, which says that the probability that a randomly chosen number is prime is inversely proportional to its number of digits, or to its logarithm.

Many questions regarding prime numbers remain unsolved. These include Goldbach's conjecture that every even integer greater than 2 can be expressed as the sum of two primes, and the twin prime conjecture that there are infinitely many pairs of primes differing from each other by 2. Such questions spurred the development of various branches of number theory, focusing on analytic or algebraic aspects of numbers. Primes are used in several routines in information technology, such as public-key cryptography, which relies on the difficulty of factoring large numbers into their prime factors. In abstract algebra, objects that behave like the prime numbers include prime elements and prime ideals.

Definition and examples

A natural number (i.e. 1, 2, 3, 4, 5, 6, etc.) is called a prime number (or a prime) if it is greater than 1 and cannot be written as a product of a × b of two integers and , both of which are smaller than . The numbers greater than 1 that are not prime are called composite numbers.[1] In other words, is prime if items cannot be divided up into smaller equal-size groups of more than one item,[2] or if it is not possible to arrange dots into a rectangular grid that is more than one dot wide and more than one dot high.[3] For example, among the numbers 1 through 6, the numbers 2, 3, and 5 are the prime numbers,[4] as there are no other numbers that divide them evenly. 1 is not prime, as it is specifically excluded in the definition. 4 = 2 × 2 and 6 = 2 × 3 are both composite.

The first 168 prime numbers (all the prime numbers less than 1000) are:[5]

2, 3, 5, 7, 11, 13, 17, 19, 23, 29, 31, 37, 41, 43, 47, 53, 59, 61, 67, 71, 73, 79, 83, 89, 97, 101, 103, 107, 109, 113, 127, 131, 137, 139, 149, 151, 157, 163, 167, 173, 179, 181, 191, 193, 197, 199, 211, 223, 227, 229, 233, 239, 241, 251, 257, 263, 269, 271, 277, 281, 283, 293, 307, 311, 313, 317, 331, 337, 347, 349, 353, 359, 367, 373, 379, 383, 389, 397, 401, 409, 419, 421, 431, 433, 439, 443, 449, 457, 461, 463, 467, 479, 487, 491, 499, 503, 509, 521, 523, 541, 547, 557, 563, 569, 571, 577, 587, 593, 599, 601, 607, 613, 617, 619, 631, 641, 643, 647, 653, 659, 661, 673, 677, 683, 691, 701, 709, 719, 727, 733, 739, 743, 751, 757, 761, 769, 773, 787, 797, 809, 811, 821, 823, 827, 829, 839, 853, 857, 859, 863, 877, 881, 883, 887, 907, 911, 919, 929, 937, 941, 947, 953, 967, 971, 977, 983, 991, 997 (sequence A000040 in the OEIS).

No even number greater than 2 is prime because any such number can be expressed as the product . Therefore, the prime numbers other than two are all odd numbers. They are often called odd primes.[6] Similarly, when written in the usual decimal system, all prime numbers larger than 5 must end in 1, 3, 7, or 9. The numbers that end with other digits are all composite: decimal numbers that end in 0, 2, 4, 6, or 8 are even, and decimal numbers that end in 0 or 5 are divisible by 5.[7]

Demonstration, with Cuisenaire rods, that 7 is prime, because none of 2, 3, 4, 5, or 6 divide it evenly

The divisors of a natural number are the numbers that divide evenly (without a remainder). Every natural number has both 1 and itself as a divisor. If it has any other divisor, it cannot be prime. Based on this idea, the primes can be defined as the numbers without other divisors: a number is prime if it has exactly two positive divisors, 1 and the number itself.[8] Yet another way to say the same thing is that a number is prime if it is greater than one and if none of the numbers divides evenly.[9]

The set of all primes is often denoted by (a boldface capital P)[10] or by (a blackboard bold capital P).[11]

Unique factorization and the primality of one

Fundamental theorem of arithmetic

The crucial importance of prime numbers to number theory and mathematics in general stems from the fundamental theorem of arithmetic,[12] which states that every integer larger than 1 can be written as a product of one or more primes in a way that is unique except for the order of the prime factors.[13] Primes can thus be considered the "basic building blocks" of the natural numbers.[14] For example:

23244 = 2 · 2 · 3 · 13 · 149
= 22 · 3 · 13 · 149. (22 denotes the square or second power of 2.)

As in this example, the same prime factor may occur multiple times. A decomposition:

n = p1 · p2 · ... · pt

of a number n into (finitely many) prime factors p1, p2, ... to pt is called prime factorization of n. The fundamental theorem of arithmetic can be rephrased so as to say that any two factorizations of the same number into primes will be identical except for the order of the factors.[15] So, although there are many different ways of finding a factorization using an integer factorization algorithm, they all must produce the same result.

If p is a prime number and p divides a product ab of integers, then p divides a or p divides b.[16] This proposition is known as Euclid's lemma. It is used in some proofs of the uniqueness of prime factorizations.[17]

Primality of one

Most early Greeks did not even consider 1 to be a number,[18][19] so they could not consider it to be a prime. A few mathematicians from this time also considered the prime and composite numbers to be subdivisions of the odd numbers, so they also did not consider 2 to be prime. However, Plato, Aristotle, Euclid, and a majority of the other Greek mathematicians considered 2 as prime. The medieval Islamic mathematicians largely followed the Greeks in viewing 1 as not being a number.[18]

By the Middle Ages and Renaissance mathematicians began treating 1 as a number, and some of them included it as the first prime number.[20] In the mid-18th century Christian Goldbach listed 1 as the first prime in his famous correspondence with Leonhard Euler; however, Euler himself did not consider 1 to be a prime number.[21] In the 19th century many mathematicians still considered the number 1 to be a prime.[22] For example, Derrick Norman Lehmer's list of primes up to 10,006,721, reprinted as late as 1956,[23] started with 1 as its first prime.[24] Henri Lebesgue has been said to be the last professional mathematician to call 1 prime,[25] but G. H. Hardy did so even later. By the early 20th century, mathematicians began to arrive at the consensus that 1 is not a prime number, but rather forms its own special category as a "unit".[22]

If the definition of a prime number were changed to call 1 a prime, many statements involving the prime numbers would not hold in the form they are usually stated, but would instead require special treatment for the number 1. For example, the number 15 can be factored as 3 · 5 and 1 · 3 · 5, so if 1 were defined to be prime, the number 15 would have two different factorizations into prime numbers. In order for the fundamental theorem of arithmetic to remain valid with this definition, it would need to be rephrased in terms of factorizations into primes greater than 1.[22] Similarly, the sieve of Eratosthenes would not work correctly if it handled 1 as a prime, because it would eliminate all multiples of 1 (that is, all other numbers) and produce as output only the single number 1.[24] Some other more technical properties of prime numbers also do not hold for the number 1: for instance, the formulas for Euler's totient function or for the sum of divisors function are different for prime numbers than they are for 1.[26]

History

The Ishango bone, from the Upper Paleolithic era, is suggested to include a list of primes
The Rhind Mathematical Papyrus

There are hints that the ancient Egyptian mathematicians had some knowledge of prime numbers: the Egyptian fraction expansions in the Rhind Mathematical Papyrus, for instance, have quite different forms for primes and for composites.[27] It has also been suggested that the engravings on the Ishango bone record a list of prime numbers.[28] However, the earliest surviving records of the explicit study of prime numbers come from Ancient Greek mathematics. Euclid's Elements (circa 300 BC) contain important theorems about primes, including the infinitude of primes and the fundamental theorem of arithmetic. Euclid also showed how to construct a perfect number from a Mersenne prime.[29] The Sieve of Eratosthenes, attributed to Eratosthenes, is a simple method to compute primes, although the large primes found today with computers are not generated this way.[30][31]

Wilson's theorem, characterizing the prime numbers as the solutions to the equation (n - 1)! ≡ -1 (mod n), was found around 1000 AD by an Islamic mathematician, Ibn al-Haytham (Alhazen). Ibn al-Haytham also investigated the perfect numbers formed from Mersenne primes, and conjectured that all perfect numbers arose in this way, but was unable to prove it.[32] Another Islamic mathematician, Ibn al-Banna' al-Marrakushi, observed that the sieve of Eratosthenes can be sped up by testing only the divisors up to the square root of the largest number to be tested. Fibonacci brought Islamic mathematics back to Europe. His book Liber Abaci (1202) was the first to describe trial division for testing primality, again using divisors only up to the square root.[31]

The next significant developments took place in 17th and 18th century Europe. In 1640 Pierre de Fermat stated (without proof) Fermat's little theorem (later proved by Leibniz and Euler).[33] Fermat also conjectured that all numbers of the form 22n + 1 are prime (they are called Fermat numbers) and he verified this up to n = 4 (or 216 + 1). However, the very next Fermat number 232 + 1 is composite (one of its prime factors is 641), as Euler discovered later, and in fact no further Fermat numbers are known to be prime.[34] The French monk Marin Mersenne looked at primes of the form 2p - 1, with p a prime. They are called Mersenne primes in his honor.[35] Christian Goldbach formulated Goldbach's conjecture, that every even number is the sum of two primes, in a 1742 letter to Euler.[36] Euler's work in number theory included many results about primes. He showed the infinite series 1/2 + 1/3 + 1/5 + 1/7 + 1/11 + … is divergent.[37] In 1747 he proved Ibn al-Haytham's conjecture (now the Euclid–Euler theorem) that the even perfect numbers are precisely the integers of the form 2p−1(2p − 1), where the second factor is a Mersenne prime.[29]

At the start of the 19th century, Legendre and Gauss independently conjectured that as x tends to infinity, the number of primes up to x is asymptotic to x/ln x, where ln x is the natural logarithm of x. Ideas of Riemann in his 1859 paper on the zeta-function sketched an outline for proving this. Although the closely related Riemann hypothesis remains unproven, Riemann's outline was completed in 1896 by Hadamard and de la Vallée Poussin, independently of each other, and the result is now known as the prime number theorem.[38] Another important 19th-century result was Dirichlet's theorem on arithmetic progressions, that every arithmetic progression contains infinitely many primes.[39]

Many mathematicians have worked on primality tests for larger numbers than would be possible by trial division. Some of these methods are restricted to specific number forms; this includes Pépin's test for Fermat numbers (1877),[40] Proth's theorem (around 1878),[41] the Lucas–Lehmer primality test (originated 1856), and the generalized Lucas primality test.[31] More recent algorithms like the Adleman–Pomerance–Rumely primality test,[30] Elliptic curve primality proving, and the AKS primality test work on arbitrary numbers[42] but are slower than the algorithms for specific number forms.[43] Since 1951 all the largest known primes have been found by computers.[44] The search for ever larger primes has generated interest outside mathematical circles. The Great Internet Mersenne Prime Search and other distributed computing projects to find large primes have become popular,[5][45] while mathematicians continue to struggle with the theory of primes.

For a long time, prime numbers were thought to have extremely limited application outside of pure mathematics.[46] This changed in the 1970s when the concepts of public-key cryptography were invented, in which prime numbers formed the basis of the first algorithms such as the RSA cryptosystem.[47] Important recent developments in the theory of prime numbers include the Green–Tao theorem (2004) on long arithmetic progressions of prime numbers, and Yitang Zhang's 2013 proof that there exist infinitely many prime gaps of bounded size.[48]

Number of prime numbers

There are infinitely many prime numbers. Another way of saying this is that the sequence

2, 3, 5, 7, 11, 13, ...

of prime numbers never ends. This statement is referred to as Euclid's theorem in honor of the ancient Greek mathematician Euclid, since the first known proof for this statement is attributed to him. Many more proofs of the infinitude of primes are known, including an analytical proof by Euler, Goldbach's proof based on Fermat numbers,[49] Furstenberg's proof using general topology,[50] and Kummer's elegant proof.[51]

Euclid's proof

Euclid's proof that there are infinitely many primes[52] shows equivalently that every finite set S of primes misses at least one prime. The key idea is to multiply together the numbers in S and add one:

Because the resulting number N is greater than one, it has at least one prime number (possibly N itself) in its prime factorization. But this prime cannot be in S, because dividing N by any one of the primes in S leaves a remainder of 1. Therefore, S does not contain all the primes.

It is often erroneously reported that Euclid begins with the assumption that the set S contains all prime numbers, leading to a contradiction, or that S contains all the primes up to some threshold rather than any arbitrary finite collection of primes.[53] The numbers formed by adding one to the products of the smallest primes are called the Euclid numbers.[54]

Euler's analytical proof

Euler's proof uses the partial sums of the reciprocals of primes,

For any arbitrary real number x, there exists a prime p for which this partial sum is bigger than x.[55] This shows that there are infinitely many primes, because if there were finitely many primes the sum would reach its maximum value at the biggest prime rather than being unbounded. More precisely, the growth rate of S(p) is doubly logarithmic, as quantified by Mertens' second theorem.[56] For comparison, the sum

does not grow to infinity as n goes to infinity (see the Basel problem). In this sense, prime numbers occur more often than squares of natural numbers.[57] Brun's theorem states that the sum of the reciprocals of twin primes,

is finite. Because of Brun's theorem, it is not possible to use Euler's method to solve the twin prime conjecture, that there exist infinitely many twin primes.[57]

Testing primality and integer factorization

There are various methods to determine whether a given number n is prime. The most basic routine, trial division, is too slow to be useful for large numbers. One group of modern primality tests is applicable to arbitrary numbers, while more efficient tests are available for numbers of special types. Most primality tests only tell whether their argument is prime or not. Routines that also provide a prime factor of composite arguments (or all of its prime factors) are called factorization algorithms.

Trial division

The most basic method of checking the primality of a given integer n is called trial division. This method divides n by each integer from 2 to the square root of n (inclusive). If any of these numbers divides n without a remainder, n is composite; otherwise it is prime. Factors larger than the square root do not need to be checked because, whenever n has a factor m bigger than the square root, it will also have another factor n/m smaller than the square root. Another optimization is to check only prime divisors in this range.[58] For instance, to check whether 37 is prime, this method divides it by the primes in the range from 2 to √37, which are 2, 3, and 5. Each division produces a nonzero remainder, so 37 is indeed prime.

Although this method is simple to describe, it is impractical for testing the primality of large integers because the number of divisions needed grows too rapidly.[59] However, trial division is still used, with a smaller limit on the divisor size than the square root, to quickly filter out composite numbers with small factors before using more complicated methods on the numbers that pass this test.[60]

Sieves

The sieve of Eratosthenes, marking in sequence unmarked numbers and, in slightly brighter hue, their square and the multiples above, starting at 2=red, and proceeding with the next uncolored number (3=green, 5=blue, 7=yellow) in the table. Finally, 11 and all unmarked numbers get colored as primes in magenta, since 11x11=121 is already beyond the table.

Before computers, mathematical tables listing all of the primes or prime factorizations up to a given limit were commonly printed. With them, one could determine whether a given number was prime by checking the table rather than doing any calculations.[61] Although these are no longer used, it is still sometimes useful to generate lists of all small prime numbers. The oldest method for this is called the sieve of Eratosthenes; variants of this method are still commonly used.[62] It operates by maintaining a table of Boolean values indicating whether each position in the table is divisible by any of the primes found so far. It loops through the table and, when it finds a position p that is not divisible by any earlier primes, outputs p as its next prime number. Then, it marks each multiple of p as being divisible by p.[63] Another more efficient sieving method for the same problem is the sieve of Atkin.[64] In advanced mathematics, sieve theory applies similar methods to other problems.[65]

Primality testing versus primality proving

Some of the fastest modern tests for whether an arbitrary given number n is prime are probabilistic (or Monte Carlo) algorithms, meaning that they have a small random chance of producing an incorrect answer.[66] For example, a given test might have the property that prime numbers will always pass the test, and that composite numbers will usually fail the test, but that a composite number could erroneously pass with some small probability ɛ. A composite number that passes such a test is called a pseudoprime.[67] If the test is repeated n times on the same number, the probability that a composite number could pass the test every time is ɛn, which decreases exponentially with the number of tests, providing high confidence (although not certainty) that a number that passes the repeated test is prime. On the other hand, if the test ever fails, then the number is certainly composite.[68] Probabilistic tests with this behavior include the Solovay–Strassen primality test and the Miller–Rabin primality test.[67]

In contrast, some other algorithms guarantee that their answer will always be correct: primes will always be determined to be prime and composites will always be determined to be composite. For instance, this is true of trial division. The algorithms with guaranteed-correct output include both deterministic (non-random) algorithms, such as the AKS primality test,[69] and randomized Las Vegas algorithms where the random choices made by the algorithm do not affect its final answer, such as some variations of elliptic curve primality proving.[66] The elliptic curve primality test is the fastest in practice of the guaranteed-correct primality tests, but its runtime analysis is based on heuristic arguments rather than rigorous proofs. The AKS primality test has mathematically proven time complexity, but is slower than elliptic curve primality proving in practice.[70] When using these methods to generate large random prime numbers, one can speed them up in practice by performing a faster probabilistic test to quickly eliminate most composite numbers before switching to a guaranteed-correct algorithm to verify that the remaining numbers are prime.[71]

The following table lists some of these tests. Their running time is given in terms of n, the number to be tested and, for probabilistic algorithms, the number k of tests performed. Moreover, ε is an arbitrarily small positive number, and log is the logarithm to an unspecified base. The big O notation means that each time bound should be multiplied by a constant factor to convert it from dimensionless units to units of time; this factor depends on implementation details such as the type of computer used to run the algorithm, but not on the input parameters n and k.

Test Developed in Type Running time Notes References
AKS primality test 2002 deterministic O((log n)6+ε) [69][72]
Elliptic curve primality proving 1977 Las Vegas O((log n)4+ε) heuristically [70]
Miller–Rabin primality test 1980 Monte Carlo O(k · (log n)2+ε) error probability 4k [73]
Solovay–Strassen primality test 1977 Monte Carlo O(k · (log n)2+ε) error probability 2k [73]

Special-purpose algorithms and the largest known prime

In addition to the aforementioned tests that apply to any natural number, some numbers of a special form can be tested for primality more easily. For example, the Lucas–Lehmer primality test can determine whether a Mersenne number (one less than a power of two) is prime, deterministically, in the same time as a single iteration of the Miller–Rabin test.[74] This is why the largest known prime has frequently been a Mersenne prime.[75]

The following table gives the largest known primes of various types. Some of these primes have been found using distributed computing. In 2009, the Great Internet Mersenne Prime Search project was awarded a US$100,000 prize for first discovering a prime with at least 10 million digits.[76] The Electronic Frontier Foundation also offers $150,000 and $250,000 for primes with at least 100 million digits and 1 billion digits, respectively.[77]

Type Prime Number of decimal digits Date Found by
Mersenne prime 277,232,917 − 1 23,249,425 December 26, 2017[78] Jonathan Pace, Great Internet Mersenne Prime Search
not a Mersenne prime (Proth number) 10,223 × 231,172,165 + 1 9,383,761 October 31, 2016[79] Péter Szabolcs, PrimeGrid[80]
factorial prime 208,003! − 1 1,015,843 July 2016 Sou Fukui[81]
primorial prime 1,098,133# − 1 476,311 March 2012 James P. Burt, PrimeGrid[82]
twin primes 2,996,863,034,895  × 21,290,000 ± 1 388,342 September 2016 Tom Greer, PrimeGrid[83]

Integer factorization

Given a composite integer n, the task of providing one (or all) prime factors is referred to as factorization of n. It is significantly more difficult than primality testing,[84] and although many factorization algorithms are known, they are slower than the fastest primality testing methods. Trial division and Pollard's rho algorithm can be used to find very small factors of n,[85] and elliptic curve factorization can be effective when n has factors of moderate size.[86] Methods suitable for arbitrary large numbers that do not depend on the size of its factors include the quadratic sieve and general number field sieve. As with primality testing, there are also factorization algorithms that require their input to have a special form, including the special number field sieve.[87] The largest number to be factored by a general-purpose algorithm is RSA-768, which has 232 decimal digits and is the product of two primes.[88]

Shor's algorithm can factor any integer in a polynomial number of steps on a quantum computer.[89] However current technology can only run this algorithm for very small numbers. The largest number that has been factored by a quantum computer running this algorithm is 21.[90]

Distribution

In 1975, number theorist Don Zagier commented that primes both

grow like weeds among the natural numbers, seeming to obey no other law than that of chance [but also] exhibit stunning regularity [and] that there are laws governing their behavior, and that they obey these laws with almost military precision.[91]

The distribution of primes in the large, such as the question how many primes are smaller than a given, large threshold, is described by the prime number theorem, but no efficient formula for the n-th prime is known.

There are arbitrarily long sequences of consecutive non-primes, as for every positive integer the consecutive integers from to (inclusive) are all composite (as is divisible by for between and ).

Dirichlet's theorem on arithmetic progressions, in its basic form, asserts that linear polynomials

with coprime integers a and b take infinitely many prime values. Stronger forms of the theorem state that the sum of the reciprocals of these prime values diverges, and that different such polynomials with the same b have approximately the same proportions of primes. The corresponding question for quadratic polynomials is less well understood.

Formulas for primes

There is no known efficient formula for primes. For example, there is no non-constant polynomial, even in several variables, that takes only prime values.[92] However, there are numerous expressions that do encode all primes, or only primes. One possible formula is based on Wilson's theorem and generates the number 2 many times and all other primes exactly once.[93] There is also a set of Diophantine equations in nine variables and one parameter with the following property: the parameter is prime if and only if the resulting system of equations has a solution over the natural numbers. This can be used to obtain a single formula with the property that all its positive values are prime.[92]

Other examples of prime-generating formulas come from Mills' theorem and a theorem of Wright. These assert that there are real constants and such that

are prime for any natural number in the first formula, and any number of exponents in the second formula.[94] Here represents the floor function, the largest integer less than or equal to the number in question. However, these are not useful for generating primes, as the primes must be generated first in order to compute the values of or .[92]

Number of prime numbers below a given number

Graphs of the prime counting function (blue), (green) and (red). The red and blue curves almost coincide.

The prime counting function is defined as the number of primes not greater than .[95] For example, , since there are five primes less than or equal to 11. Methods such as the Meissel–Lehmer algorithm can compute exact values of faster than it would be possible to list each prime up to .[96] The prime number theorem states that satisfies

which means that the ratio of to the right hand fraction approaches 1 as grows to infinity.[97] This implies that the likelihood that a randomly chosen number less than is prime is (approximately) inversely proportional to the number of digits in .[98] A more accurate estimate for is given by the offset logarithmic integral[97]

The prime number theorem implies that the th prime number (where , , etc.) is proportional to .[99] In order for the primes to grow this quickly, some prime gaps, the differences of two consecutive primes, must be large.[100] The existence of arbitrarily large prime gaps can also be seen in a more elementary way by noting that the sequence consists of composite numbers, for any natural number .[101] However, large prime gaps occur much earlier than this argument shows.[100] For example, the first prime gap of length 8 is between the primes 89 and 97,[102] much smaller than .

Arithmetic progressions

An arithmetic progression is a finite or infinite sequence of numbers such that consecutive numbers in the sequence all have the same difference.[103] This difference is called the modulus of the progression.[104] For example,

3, 12, 21, 30, 39, ...,

is an infinite arithmetic progression with modulus 9. In an arithmetic progression, all the numbers have the same remainder when divided by the modulus; in this example, the remainder is 3. Because both the modulus 9 and the remainder 3 are multiples of 3, so is every element in the sequence. Therefore, this progression contains only one prime number, 3 itself. In general, the infinite progression

can have more than one prime only when its remainder and modulus are coprime, which happens when their greatest common divisor is one. If they are coprime, Dirichlet's theorem on arithmetic progressions asserts that the progression contains infinitely many primes.[105]

Primes in the arithmetic progressions modulo 9. Each row of the thin horizontal band shows one of the nine possible progressions mod 9, with prime numbers marked in red. The progressions of numbers that are 0, 3, or 6 mod 9 contain at most one prime number (the number 3); the remaining progressions of numbers that are 2, 4, 5, 7, and 8 mod 9 have infinitely many prime numbers, with similar numbers of primes in each progression.

The Green–Tao theorem shows that there are arbitrarily long finite arithmetic progressions consisting only of primes.[48][106]

Some other results about prime numbers depend on their membership in arithmetic progressions. For instance, according to Fermat's theorem on sums of two squares, an odd prime is expressible as the sum of two squares, , exactly when belongs to the progression 1, 5, 9, 13, ...[107]

Prime values of quadratic polynomials

The Ulam spiral. Prime numbers (red) cluster on some diagonals and not others. Prime values of are shown in blue.

Euler noted that the function

yields prime numbers for , although composite numbers appear among its later values.[108][109] The search for an explanation for this phenomenon led to the deep algebraic number theory of Heegner numbers and the class number problem.[110] The Hardy-Littlewood conjecture F predicts the density of primes among the values of quadratic polynomials with integer coefficients in terms of the logarithmic integral and the polynomial coefficients. No quadratic polynomial has been proven to take infinitely many prime values.[111]

The Ulam spiral arranges the natural numbers in a two-dimensional grid, spiraling in concentric squares surrounding the origin with the prime numbers highlighted. Visually, the primes appear to cluster on certain diagonals and not others, suggesting that some quadratic polynomials take prime values more often than others.[111]

Open questions

Zeta function and the Riemann hypothesis

Plot of the absolute values of the zeta function, showing some of its features

The Riemann zeta function is an analytic function on the complex numbers. For complex numbers with real part greater than one it equals both an infinite sum over all integers, and an infinite product over the prime numbers,

This equality between a sum and a product, discovered by Euler, is called an Euler product.[112] The Euler product can be derived from the fundamental theorem of arithmetic, and shows the close connection between the zeta function and the prime numbers.[113] It leads to another proof that there are infinitely many primes: if there were only finitely many, then the sum-product equality would also be valid at , but the sum would diverge (it is the harmonic series ) while the product would be finite, a contradiction.[114] In fact the zeta function has its only pole at ; it has finite values at all other complex numbers, even the ones with real part less than one where the sum and product formulas diverge.[115] For instance, , a fact that is sometimes used to assign a finite value to the divergent sum 1 + 2 + 3 + 4 + ⋯.[116]

The Basel problem asked to evaluate , a sum now recognizable as . Euler's first major result was its solution, .[117] The reciprocal of this number, , is the limiting probability that two random numbers selected uniformly from a large range are relatively prime.[118]

The unproven Riemann hypothesis, dating from 1859, states that the zeroes of the zeta-function are all either negative even numbers, or complex numbers with real part equal to 1/2.[119] The original proof of the prime number theorem was based on a weak form of this hypothesis, that there are no zeros with real part equal to 1,[120][121] although other more elementary proofs have been found.[122] The prime-counting function can be expressed by Riemann's explicit formula as a sum in which each term comes from one of the zeros of the zeta function; the main term of this sum is the logarithmic integral, and the remaining terms cause the sum to fluctuate above and below the main term.[123] In this sense, the zeros control how regularly the prime numbers are distributed. If the Riemann hypothesis is true, these fluctuations will be small, and the asymptotic distribution of primes given by the prime number theorem will also hold also hold over much shorter intervals (of length about the square root of for intervals near a number ).[121]

Other conjectures

In addition to the Riemann hypothesis, many more conjectures revolving about primes have been posed. Often having an elementary formulation, many of these conjectures have withstood proof for decades: all four of Landau's problems from 1912 are still unsolved.[124] One of them is Goldbach's conjecture, which asserts that every even integer n greater than 2 can be written as a sum of two primes.[125] As of 2014, this conjecture has been verified for all numbers up to n = 4 · 1018.[126] Weaker statements than this have been proven, for example Vinogradov's theorem says that every sufficiently large odd integer can be written as a sum of three primes.[127] Chen's theorem says that every sufficiently large even number can be expressed as the sum of a prime and a semiprime, the product of two primes.[128] Also, any even integer can be written as the sum of six primes.[129] The branch of number theory studying such questions is called additive number theory.[130]

Other conjectures deal with existence of prime numbers in other integer sequences. It is conjectured that there are infinitely many Fibonacci primes[131] and infinitely many Mersenne primes, but not Fermat primes.[132] It is not known whether or not there are an infinite number of Wieferich primes[133] and of prime Euclid numbers.[134]

A third type of conjecture concerns prime gaps. It is conjectured that there are infinitely many twin primes, pairs of primes with difference 2; this is the twin prime conjecture. Polignac's conjecture states more generally that for every positive integer k, there are infinitely many pairs of consecutive primes that differ by 2k.[135] Andrica's conjecture,[135] Brocard's conjecture,[136] Legendre's conjecture,[137] and Oppermann's conjecture[136] all suggest that the prime gaps up to should be at most approximately , a result that is known to follow from the Riemann hypothesis, while the much stronger Cramér's conjecture sets the largest gap size at .[135]

Applications

For a long time, number theory in general, and the study of prime numbers in particular, was seen as the canonical example of pure mathematics, with no applications outside of the self-interest of studying the topic with the exception of use of prime numbered gear teeth to distribute wear evenly. In particular, number theorists such as British mathematician G. H. Hardy prided themselves on doing work that had absolutely no military significance.[138] However, this vision was shattered in the 1970s, when it was publicly announced that prime numbers could be used as the basis for the creation of public key cryptography algorithms.[47] Prime numbers are also used for hash tables and pseudorandom number generators.

Arithmetic modulo a prime and finite fields

Modular arithmetic modifies usual arithmetic by only using the numbers , for a natural number called the modulus. Any other natural number can be mapped into this system by replacing it by its remainder after division by .[139] Modular sums, differences and products are calculated by performing the same replacement by the remainder on the result of the usual sum, difference, or product of integers.[140] Equality of integers corresponds to congruence in modular arithmetic: We write that and are congruent ( mod ) when they have the same remainder after division by .[141] However, in this system of numbers, division is not as easy to define as the other operations. Division by all nonzero numbers is possible if and only if the modulus is prime. In the terminology of abstract algebra, the ability to perform division means that modular arithmetic modulo a prime number forms a field or, more specifically, a finite field, while other moduli only give a ring but not a field.[142]

Several theorems about primes can be viewed through the lens of modular arithmetic. For instance, Fermat's little theorem states that if (mod ), then (mod ).[143] Summing this over all choices of gives the equation

valid whenever is prime. Giuga's conjecture says that this equation is also a sufficient condition for to be prime.[144] Wilson's theorem says that an integer is prime if and only if the factorial is congruent to mod . For composite numbers , is instead congruent to 0 mod .[145]

Fermat primes and constructible polygons

Construction of a regular pentagon using straightedge and compass. This is only possible because 5 is a Fermat prime.

Fermat primes are primes of the form

Fk = 22k + 1,

with k a natural number. They are named after Pierre de Fermat, who conjectured that all such numbers are prime. This was based on the evidence of the first five numbers in this series—3, 5, 17, 257, and 65,537—being prime.[146] However, F5 is composite and so are all other Fermat numbers that have been verified as of 2017.[147] A regular n-gon is constructible using straightedge and compass if and only if the odd prime factors of n (if any) are distinct Fermat primes.[146]

Other mathematical occurrences of primes

Many mathematical domains make great use of prime numbers. An example from the theory of finite groups are the Sylow theorems: if G is a finite group and pn is the highest power of the prime p that divides the order of G, then G has a subgroup of order pn. Also, any group of prime order is cyclic (Lagrange's theorem) and any group whose order is divisible by only two primes is solvable (the Burnside theorem).

Public-key cryptography

Several public-key cryptography algorithms, such as RSA and the Diffie–Hellman key exchange, are based on large prime numbers (2048-bit primes are common). RSA relies on the assumption that it is much easier (i.e., more efficient) to perform the multiplication of two (large) numbers x and y than to calculate x and y (assumed coprime) if only the product xy is known.[47] The Diffie–Hellman key exchange relies on the fact that there are efficient algorithms for modular exponentiation, while the reverse operation the discrete logarithm is thought to be a hard problem.

Prime numbers in nature

The evolutionary strategy used by cicadas of the genus Magicicada makes use of prime numbers.[148] These insects spend most of their lives as grubs underground. They only pupate and then emerge from their burrows after 7, 13 or 17 years, at which point they fly about, breed, and then die after a few weeks at most. Biologists theorize that these prime-numbered breeding cycle lengths have evolved in order to prevent predators to synchronize with these cycles.[149][150]

There is speculation[by whom?] that the zeros of the zeta function are connected to the energy levels of complex quantum systems.[151]

Generalizations

The concept of prime number is so important that it has been generalized in different ways in various branches of mathematics. Generally, "prime" indicates minimality or indecomposability, in an appropriate sense. For example, the prime field is the smallest subfield of a field F containing both 0 and 1. It is either Q or the finite field with p elements, whence the name.[152] Often a second, additional meaning is intended by using the word prime, namely that any object can be, essentially uniquely, decomposed into its prime components. For example, in knot theory, a prime knot is a knot that is indecomposable in the sense that it cannot be written as the knot sum of two nontrivial knots. Any knot can be uniquely expressed as a connected sum of prime knots.[153] Prime models and prime 3-manifolds are other examples of this type.

Prime elements in rings

Prime numbers give rise to two more general concepts that apply to elements of any commutative ring R, an algebraic structure where addition, subtraction and multiplication are defined: prime elements and irreducible elements. An element p of R is called prime element if it is neither zero nor a unit (i.e., does not have a multiplicative inverse) and satisfies the following requirement: given x and y in R such that p divides the product xy, then p divides x or y. An element is irreducible if it is not a unit and cannot be written as a product of two ring elements that are not units. In the ring Z of integers, the set of prime elements equals the set of irreducible elements, which is

In any ring R, any prime element is irreducible. The converse does not hold in general, but does hold for unique factorization domains.

The fundamental theorem of arithmetic continues to hold in unique factorization domains. An example of such a domain is the Gaussian integers Z[i], that is, the set of complex numbers of the form a + bi where i denotes the imaginary unit and a and b are arbitrary integers. Its prime elements are known as Gaussian primes. Not every prime (in Z) is a Gaussian prime: in the bigger ring Z[i], 2 factors into the product of the two Gaussian primes (1 + i) and (1 − i). Rational primes (i.e. prime elements in Z) of the form 4k + 3 are Gaussian primes, whereas rational primes of the form 4k + 1 are not.

Prime ideals

In ring theory, the notion of number is generally replaced with that of ideal. Prime ideals, which generalize prime elements in the sense that the principal ideal generated by a prime element is a prime ideal, are an important tool and object of study in commutative algebra, algebraic number theory and algebraic geometry. The prime ideals of the ring of integers are the ideals (0), (2), (3), (5), (7), (11), … The fundamental theorem of arithmetic generalizes to the Lasker–Noether theorem, which expresses every ideal in a Noetherian commutative ring as an intersection of primary ideals, which are the appropriate generalizations of prime powers.[154]

Prime ideals are the points of algebro-geometric objects, via the notion of the spectrum of a ring.[155] Arithmetic geometry also benefits from this notion, and many concepts exist in both geometry and number theory. For example, factorization or ramification of prime ideals when lifted to an extension field, a basic problem of algebraic number theory, bears some resemblance with ramification in geometry. Such ramification questions occur even in number-theoretic questions solely concerned with integers. For example, prime ideals in the ring of integers of quadratic number fields can be used in proving quadratic reciprocity, a statement that concerns the solvability of quadratic equations

where x is an integer and p and q are (usual) prime numbers.[156] Early attempts to prove Fermat's Last Theorem climaxed when Kummer introduced regular primes, primes satisfying a certain requirement concerning the failure of unique factorization in the ring consisting of expressions

where a0, ..., ap−1 are integers and ζ is a complex number such that ζp = 1.[157]

Valuations

Valuation theory studies certain functions from a field K to the real numbers R called valuations.[158] Every such valuation yields a topology on K, and two valuations are called equivalent if they yield the same topology. A prime of K (sometimes called a place of K) is an equivalence class of valuations. For example, the p-adic valuation of a rational number q is defined to be the integer vp(q), such that

where both r and s are not divisible by p. For example, v3(18/7) = 2. The p-adic norm is defined as [159]

In particular, this norm gets smaller when a number is multiplied by p, in sharp contrast to the usual absolute value (also referred to as the infinite prime). While completing Q (roughly, filling the gaps) with respect to the absolute value yields the field of real numbers, completing with respect to the p-adic norm |−|p yields the field of p-adic numbers.[160] These are essentially all possible ways to complete Q, by Ostrowski's theorem. Certain arithmetic questions related to Q or more general global fields may be transferred back and forth to the completed (or local) fields. This local-global principle again underlines the importance of primes to number theory.

In games, arts, and literature

Prime numbers have influenced many artists and writers. The French composer Olivier Messiaen used prime numbers to create ametrical music through "natural phenomena". In works such as La Nativité du Seigneur (1935) and Quatre études de rythme (1949–50), he simultaneously employs motifs with lengths given by different prime numbers to create unpredictable rhythms: the primes 41, 43, 47 and 53 appear in the third étude, "Neumes rythmiques". According to Messiaen this way of composing was "inspired by the movements of nature, movements of free and unequal durations".[161]

In his science fiction novel Contact, NASA scientist Carl Sagan suggested that prime numbers could be used as a means of communicating with aliens, an idea that he had first developed informally with American astronomer Frank Drake in 1975.[162] In the novel The Curious Incident of the Dog in the Night-Time by Mark Haddon, the narrator arranges the sections of the story by consecutive prime numbers.[163]

Many films, such as Cube, Sneakers, The Mirror Has Two Faces and A Beautiful Mind reflect a popular fascination with the mysteries of prime numbers and cryptography. Prime numbers are used as a metaphor for loneliness and isolation in the Paolo Giordano novel The Solitude of Prime Numbers, in which they are portrayed as "outsiders" among integers.[164]

See also

References

  1. ^ Gardiner, Anthony (1997). The Mathematical Olympiad Handbook: An Introduction to Problem Solving Based on the First 32 British Mathematical Olympiads 1965–1996. Oxford University Press. p. 26. ISBN 9780198501053.
  2. ^ Henderson, Anne (2014). Dyslexia, Dyscalculia and Mathematics: A practical guide (2nd ed.). Routledge. p. 62. ISBN 9781136636622.
  3. ^ Adler, Irving (1960). The Giant Golden Book of Mathematics: Exploring the World of Numbers and Space. Golden Press. p. 16.
  4. ^ Leff, Lawrence S. (2000). Math Workbook for the SAT I. Barron's Educational Series. p. 360. ISBN 9780764107689.
  5. ^ a b Ziegler, Günter M. (2004). "The great prime number record races". Notices of the American Mathematical Society. 51 (4): 414–416. MR 2039814.
  6. ^ Stillwell, John (1997). Numbers and Geometry. Undergraduate Texts in Mathematics. Springer. p. 9. ISBN 9780387982892.
  7. ^ Sierpiński, Wacław (1964). A Selection of Problems in the Theory of Numbers. New York: Macmillan. p. 40. MR 0170843.
  8. ^ Dudley, Underwood (1978). "Section 2: Unique factorization". Elementary number theory (2nd ed.). W. H. Freeman and Co. p. 10. ISBN 978-0-7167-0076-0. {{cite book}}: Invalid |ref=harv (help)
  9. ^ Sierpiński, Wacław (1988). Elementary Theory of Numbers. North-Holland Mathematical Library. Vol. 31 (2nd ed.). Elsevier. p. 113. ISBN 9780080960197. {{cite book}}: Invalid |ref=harv (help)
  10. ^ Nathanson, Melvyn B. (2000). "Notations and Conventions". Elementary Methods in Number Theory. Graduate Texts in Mathematics. Vol. 195. Springer. ISBN 9780387227382. MR 1732941. {{cite book}}: Invalid |ref=harv (help)
  11. ^ Faticoni, Theodore G. (2012). The Mathematics of Infinity: A Guide to Great Ideas. Pure and Applied Mathematics: A Wiley Series of Texts, Monographs and Tracts. Vol. 111 (2nd ed.). John Wiley & Sons. p. 44. ISBN 9781118243824.
  12. ^ Smith, Karl J. (2011). The Nature of Mathematics (12th ed.). Cengage Learning. p. 188. ISBN 9780538737586.
  13. ^ Dudley 1978, Section 2, Theorem 2, p. 16
  14. ^ du Sautoy, Marcus (2003). The Music of the Primes: Searching to Solve the Greatest Mystery in Mathematics. Harper Collins. p. 23. ISBN 9780060935580.
  15. ^ Neale, Vicky (2017). Closing the Gap: The Quest to Understand Prime Numbers. Oxford University Press. p. 107. ISBN 9780191092435. {{cite book}}: Invalid |ref=harv (help)
  16. ^ Dudley 1978, Section 2, Lemma 5, p. 15
  17. ^ Higgins, Peter M. (1998). Mathematics for the Curious. Oxford University Press. pp. 77–78. ISBN 9780191500503.
  18. ^ a b Caldwell, Chris K.; Reddick, Angela; Xiong, Yeng; Keller, Wilfrid (2012). "The history of the primality of one: a selection of sources". Journal of Integer Sequences. 15 (9): Article 12.9.8. MR 3005523. {{cite journal}}: Invalid |ref=harv (help) For a selection of quotes from and about the ancient Greek positions on this issue, see in particular pp. 3–4. For the Islamic mathematicians, see p. 6.
  19. ^ Tarán, Leonardo (1981). Speusippus of Athens: A Critical Study With a Collection of the Related Texts and Commentary. Philosophia Antiqua : A Series of Monographs on Ancient Philosophy. Vol. 39. BRILL. pp. 35–38. ISBN 9789004065055.
  20. ^ Caldwell et al. 2012, pp. 7–13. See in particular the entries for Stevin, Brancker, Wallis, and Prestet.
  21. ^ Caldwell et al. 2012, p. 15.
  22. ^ a b c Caldwell, Chris K.; Xiong, Yeng (2012). "What is the smallest prime?" (PDF). Journal of Integer Sequences. 15 (9): Article 12.9.7. MR 3005530.
  23. ^ Riesel, Hans (1994). Prime Numbers and Computer Methods for Factorization (2nd ed.). Basel, Switzerland: Birkhäuser. p. 36. ISBN 978-0-8176-3743-9. MR 1292250. {{cite book}}: Invalid |ref=harv (help)
  24. ^ a b Conway, John Horton; Guy, Richard K. (1996). The Book of Numbers. New York: Copernicus. pp. 129–130. ISBN 978-0-387-97993-9. MR 1411676. {{cite book}}: Invalid |ref=harv (help)
  25. ^ Derbyshire, John (2003). "The Prime Number Theorem". Prime Obsession: Bernhard Riemann and the Greatest Unsolved Problem in Mathematics. Washington, D.C.: Joseph Henry Press. p. 33. ISBN 978-0-309-08549-6. MR 1968857. OCLC 249210614.
  26. ^ For the totient, see Sierpiński 1988, p. 245. For the sum of divisors, see Sandifer, C. Edward (2007). How Euler Did It. MAA Spectrum. Mathematical Association of America. p. 59. ISBN 9780883855638. {{cite book}}: Invalid |ref=harv (help)
  27. ^ Bruins, Evert Marie, review in Mathematical Reviews of Gillings, R. J. (1974). "The recto of the Rhind Mathematical Papyrus. How did the ancient Egyptian scribe prepare it?". Archive for History of Exact Sciences. 12: 291–298. doi:10.1007/BF01307175. MR 0497458.
  28. ^ Everett, Caleb (2017). Numbers and the Making of Us: Counting and the Course of Human Cultures. Harvard University Press. p. 35. ISBN 9780674504431.
  29. ^ a b Stillwell, John (2010). Mathematics and Its History. Undergraduate Texts in Mathematics (3rd ed.). Springer. p. 40. ISBN 9781441960528.
  30. ^ a b Pomerance, Carl (December 1982). "The Search for Prime Numbers". Scientific American. 247 (6): 136–147. JSTOR 24966751.
  31. ^ a b c Mollin, Richard A. (2002). "A brief history of factoring and primality testing B. C. (before computers)". Mathematics Magazine. 75 (1): 18–29. doi:10.2307/3219180. MR 2107288.
  32. ^ O'Connor, John J.; Robertson, Edmund F., "Abu Ali al-Hasan ibn al-Haytham", MacTutor History of Mathematics Archive, University of St Andrews
  33. ^ Sandifer 2007, 8. Fermat's Little Theorem (November 2003), p. 45
  34. ^ Sandifer, C. Edward (2014). How Euler Did Even More. Mathematical Association of America. p. 42. ISBN 9780883855843.
  35. ^ Koshy, Thomas (2002). Elementary Number Theory with Applications. Academic Press. p. 369. ISBN 9780124211711. {{cite book}}: Invalid |ref=harv (help)
  36. ^ Yuan, Wang (2002). Goldbach Conjecture. Series In Pure Mathematics. Vol. 4 (2nd ed.). World Scientific. p. 21. ISBN 9789814487528.
  37. ^ Narkiewicz, Wladyslaw (2000). "1.2 Sum of Reciprocals of Primes". The Development of Prime Number Theory: From Euclid to Hardy and Littlewood. Springer Monographs in Mathematics. Springer. p. 11. ISBN 9783540662891.
  38. ^ Apostol, Tom M. (2000). "A centennial history of the prime number theorem". In Bambah, R. P.; Dumir, V. C.; Hans-Gill, R. J. (eds.). Number Theory. Trends in Mathematics. Basel: Birkhäuser. pp. 1–14. MR 1764793.
  39. ^ Apostol, Tom M. (1976). "7. Dirichlet's Theorem on Primes in Arithmetical Progressions". Introduction to Analytic Number Theory. New York and Heidelberg: Springer-Verlag. pp. 146–156. MR 0434929. {{cite book}}: Invalid |ref=harv (help)
  40. ^ Chabert, Jean-Luc (2012). A History of Algorithms: From the Pebble to the Microchip. Springer. p. 261. ISBN 9783642181924.
  41. ^ Rosen, Kenneth H. (2000). "Theorem 9.20. Proth's Primality Test". Elementary Number Theory and Its Applications (4th ed.). Addison-Wesley. p. 342. ISBN 9780201870732. {{cite book}}: Invalid |ref=harv (help)
  42. ^ Bauer, Craig P. (2013). Secret History: The Story of Cryptology. Discrete Mathematics and Its Applications. CRC Press. p. 468. ISBN 9781466561861.
  43. ^ Klee, Victor; Wagon, Stan (1991). Old and New Unsolved Problems in Plane Geometry and Number Theory. Dolciani mathematical expositions. Vol. 11. Cambridge University Press. p. 224. ISBN 9780883853153.
  44. ^ A 44-digit prime number found in 1951 by Aimé Ferrier with a mechanical calculator remains the largest prime not to have been found with the aid of electronic computers. See e.g. Cooper, S. Barry; Hodges, Andrew (2016). The Once and Future Turing. Cambridge University Press. pp. 37–38. ISBN 9781107010833.
  45. ^ Rosen 2000, p. 245.
  46. ^ For instance, Beiler writes that number theorist Ernst Kummer loved his ideal numbers, closely related to the primes, "because they had not soiled themselves with any practical applications", and Katz writes that Edmund Landau, known for his work on the distribution of primes, "loathed practical applications of mathematics", and for this reason avoided subjects such as geometry that had already shown themselves to be useful. Beiler, Albert H. (1966). Recreations in the Theory of Numbers: The Queen of Mathematics Entertains. Dover. p. 2. ISBN 9780486210964. Katz, Shaul (2004). "Berlin roots—Zionist incarnation: the ethos of pure mathematics and the beginnings of the Einstein Institute of Mathematics at the Hebrew University of Jerusalem". Science in Context. 17 (1–2): 199–234. doi:10.1017/S0269889704000092. MR 2089305.
  47. ^ a b c Kraft, James S.; Washington, Lawrence C. (2014). Elementary Number Theory. Textbooks in mathematics. CRC Press. p. 7. ISBN 9781498702690. {{cite book}}: Invalid |ref=harv (help)
  48. ^ a b Neale 2017, pp. 18 and 47.
  49. ^ Letter in Latin from Goldbach to Euler, July 1730.
  50. ^ Furstenberg, Harry (1955). "On the infinitude of primes". American Mathematical Monthly. 62 (5). Mathematical Association of America: 353. doi:10.2307/2307043. MR 0068566.
  51. ^ Ribenboim, Paulo (2004). The little book of bigger primes. Berlin, New York: Springer-Verlag. p. 4. ISBN 978-0-387-20169-6. {{cite book}}: Invalid |ref=harv (help)
  52. ^ Euclid's Elements, Book IX, Proposition 20. See David Joyce's English translation of Euclid's proof or Williamson, James (1782). The Elements of Euclid, With Dissertations. Oxford: Clarendon Press. p. 63.
  53. ^ Hardy, Michael; Woodgold, Catherine (2009). "Prime Simplicity". Mathematical Intelligencer. 31 (4): 44–52. doi:10.1007/s00283-009-9064-8.
  54. ^ Vardi, Ilan (1991). Computational Recreations in Mathematica. Addison-Wesley. pp. 82–89. ISBN 9780201529890.
  55. ^ Apostol 1976, Section 1.6, Theorem 1.13
  56. ^ Apostol 1976, Section 4.8, Theorem 4.12
  57. ^ a b Miller, Steven J.; Takloo-Bighash, Ramin (2006). An Invitation to Modern Number Theory. Princeton University Press. pp. 43–44. ISBN 9780691120607.
  58. ^ Giblin, P. J. (1993). Primes and Programming. Cambridge University Press. p. 39. ISBN 9780521409889. {{cite book}}: Invalid |ref=harv (help)
  59. ^ Giblin 1993, p. 54
  60. ^ Riesel 1994, p. 220.
  61. ^ Bullynck, Maarten (2010). "A history of factor tables with notes on the birth of number theory 1657–1817". Revue d'Histoire des Mathématiques. 16 (2): 133–216.
  62. ^ Wagstaff, Samuel S. Jr. (2013). The Joy of Factoring. Student mathematical library. Vol. 68. American Mathematical Society. p. 191. ISBN 9781470410483.
  63. ^ Crandall, Richard; Pomerance, Carl (2005). Prime Numbers: A Computational Perspective (2nd ed.). Springer. p. 121. ISBN 9780387252827. {{cite book}}: Invalid |ref=harv (help)
  64. ^ Farach-Colton, Martín; Tsai, Meng-Tsung (2015). "On the complexity of computing prime tables". In Elbassioni, Khaled; Makino, Kazuhisa (eds.). Algorithms and Computation: 26th International Symposium, ISAAC 2015, Nagoya, Japan, December 9-11, 2015, Proceedings. Lecture Notes in Computer Science. Vol. 9472. Springer. pp. 677–688. arXiv:1504.05240. doi:10.1007/978-3-662-48971-0_57.
  65. ^ Greaves, George (2013). Sieves in Number Theory. Ergebnisse der Mathematik und ihrer Grenzgebiete (3. Folge). Vol. 43. Springer. p. 1. ISBN 9783662046586.
  66. ^ a b Hromkovič, Juraj (2001). "5.5 Bibliographic Remarks". Algorithmics for Hard Problems. Texts in Theoretical Computer Science. An EATCS Series. Springer-Verlag, Berlin. pp. 383–385. doi:10.1007/978-3-662-04616-6. ISBN 3-540-66860-8. MR 1843669.
  67. ^ a b Koblitz, Neal (1987). "Chapter V. Primality and Factoring". A Course in Number Theory and Cryptography. Graduate Texts in Mathematics. Vol. 114. Springer-Verlag, New York. pp. 112–149. doi:10.1007/978-1-4684-0310-7_5. ISBN 0-387-96576-9. MR 0910297.
  68. ^ Pieprzyk, Josef; Hardjono, Thomas; Seberry, Jennifer (2013). "2.3.9 Probabilistic Computations". Fundamentals of Computer Security. Springer. pp. 51–52. ISBN 9783662073247.
  69. ^ a b Tao, Terence (2010). "1.11 The AKS primality test". An epsilon of room, II: Pages from year three of a mathematical blog. Providence, RI: American Mathematical Society. pp. 82–86. doi:10.1090/gsm/117. ISBN 978-0-8218-5280-4. MR 2780010.
  70. ^ a b Morain, F. (2007). "Implementing the asymptotically fast version of the elliptic curve primality proving algorithm". Mathematics of Computation. 76 (257): 493–505. doi:10.1090/S0025-5718-06-01890-4. MR 2261033.
  71. ^ Indeed, much of the analysis of elliptic curve primality proving is based on the assumption that the input to the algorithm has already passed a probabilistic test. See, e.g., Atkin, A. O. L.; Morain, F. (1993). "Elliptic curves and primality proving". Mathematics of Computation. 61 (203): 29–68. doi:10.2307/2152935. MR 1199989.
  72. ^ Lenstra, Jr., H. W.; Pomerance, Carl (December 11, 2016). "Primality testing with Gaussian periods" (PDF). {{cite journal}}: Cite journal requires |journal= (help)CS1 maint: multiple names: authors list (link)
  73. ^ a b Monier, Louis (1980). "Evaluation and comparison of two efficient probabilistic primality testing algorithms". Theoretical Computer Science. 12 (1): 97–108. doi:10.1016/0304-3975(80)90007-9. MR 0582244.
  74. ^ Tao, Terence (2009). "1.7 The Lucas–Lehmer test for Mersenne primes". Poincaré's legacies, pages from year two of a mathematical blog. Part I. Providence, RI: American Mathematical Society. pp. 36–41. ISBN 978-0-8218-4883-8. MR 2523047. {{cite book}}: Invalid |ref=harv (help)
  75. ^ Kraft & Washington 2014, p. 41.
  76. ^ "Record 12-Million-Digit Prime Number Nets $100,000 Prize". Electronic Frontier Foundation. October 14, 2009. Retrieved 2010-01-04.
  77. ^ "EFF Cooperative Computing Awards". Electronic Frontier Foundation. Retrieved 2010-01-04.
  78. ^ Priday, Richard (January 5, 2018). "GIMPS crack whip on plucky processor to find largest prime number: 23 million-digit figure revealed after six days of non-stop calculations". The Register.
  79. ^ "PrimeGrid's Seventeen or Bust Subproject" (PDF). Retrieved 2017-01-03.
  80. ^ Chris K. Caldwell. "The Top Twenty: Largest Known Primes". The Prime Pages. Retrieved 2017-01-03.
  81. ^ Chris K. Caldwell. "The Top Twenty: Factorial". The Prime Pages. Retrieved 2017-01-03.
  82. ^ Chris K. Caldwell. "The Top Twenty: Primorial". The Prime Pages. Retrieved 2017-01-03.
  83. ^ Chris K. Caldwell. "The Top Twenty: Twin Primes". The Prime Pages. Retrieved 2017-01-03.
  84. ^ Kraft & Washington 2014, p. 275.
  85. ^ Riesel 1994, p. 220.
  86. ^ Hoffstein, Jeffrey; Pipher, Jill; Silverman, Joseph H. (2014). An Introduction to Mathematical Cryptography. Undergraduate Texts in Mathematics (2nd ed.). Springer. p. 329. ISBN 9781493917112.
  87. ^ Pomerance, Carl (1996). "A tale of two sieves". Notices of the American Mathematical Society. 43 (12): 1473–1485. MR 1416721.
  88. ^ T. Kleinjung; K. Aoki; J. Franke; A. K. Lenstra; E. Thomé; J. W. Bos; P. Gaudry; A. Kruppa; P. L. Montgomery; D. A. Osvik; H. te Riele; A. Timofeev; P. Zimmermann. "Factorization of a 768-bit RSA modulus" (PDF). Retrieved 2013-04-11.
  89. ^ Rieffel, Eleanor G.; Polak, Wolfgang H. (2011). "Chapter 8. Shor's Algorithm". Quantum Computing: A Gentle Introduction. MIT Press. pp. 163–176. ISBN 9780262015066.
  90. ^ Martín-López, Enrique; Enrique Martín-López; Anthony Laing; Thomas Lawson; Roberto Alvarez; Xiao-Qi Zhou; Jeremy L. O'Brien (12 October 2012). "Experimental realization of Shor's quantum factoring algorithm using qubit recycling". Nature Photonics. 6 (11): 773. arXiv:1111.4147. Bibcode:2012NaPho...6..773M. doi:10.1038/nphoton.2012.259.
  91. ^ Havil, Julian (2003). Gamma: Exploring Euler's Constant. Princeton University Press. p. 171. ISBN 978-0-691-09983-5. MR 1968276.
  92. ^ a b c Matiyasevich, Yuri V. (1999). "Formulas for prime numbers". In Tabachnikov, Serge (ed.). Kvant Selecta: Algebra and Analysis,. Vol. II. American Mathematical Society. pp. 13–24. ISBN 978-0-8218-1915-9. {{cite book}}: Unknown parameter |editorlink1= ignored (|editor-link1= suggested) (help)
  93. ^ Mackinnon, Nick (June 1987). "Prime number formulae". The Mathematical Gazette. 71 (456): 113. doi:10.2307/3616496.
  94. ^ Wright, E. M. (1951). "A prime-representing function". American Mathematical Monthly. 58 (9): 616–618. doi:10.2307/2306356. JSTOR 2306356.
  95. ^ Crandall & Pomerance 2005, p. 6.
  96. ^ Crandall & Pomerance 2005, Section 3.7, Counting primes, pp. 152–162.
  97. ^ a b Crandall & Pomerance 2005, p. 10.
  98. ^ du Sautoy, Marcus (2011). "What are the odds that your telephone number is prime?". The Number Mysteries: A Mathematical Odyssey through Everyday Life. St. Martin's Press. pp. 50–52. ISBN 9780230120280.
  99. ^ Apostol 1976, Section 4.6, Theorem 4.7
  100. ^ a b Riesel 1994, "Large gaps between consecutive primes", pp. 78–79.
  101. ^ Koshy 2002, Theorem 2.14, p. 109. Riesel 1994 gives a similar argument using the primorial in place of the factorial.
  102. ^ Sloane, N. J. A. (ed.). "Sequence A100964 (Smallest prime number that begins a prime gap of at least 2n)". The On-Line Encyclopedia of Integer Sequences. OEIS Foundation.
  103. ^ Gelfand, I. M.; Shen, Alexander (2003). Algebra. Springer. p. 37. ISBN 9780817636777.
  104. ^ Mollin, Richard A. (1997). Fundamental Number Theory with Applications. Discrete Mathematics and Its Applications. CRC Press. p. 76. ISBN 9780849339875.
  105. ^ Crandall & Pomerance 2005, Theorem 1.1.5, p. 12.
  106. ^ Green, Ben; Tao, Terence (2008). "The primes contain arbitrarily long arithmetic progressions". Annals of Mathematics. 167 (2): 481–547. arXiv:math.NT/0404188. doi:10.4007/annals.2008.167.481.
  107. ^ Kraft & Washington 2014, Section 12.1, Sums of two squares, pp. 297–301.
  108. ^ Hua, L. K. (1965). Additive Theory of Prime Numbers. Translations of Mathematical Monographs. Vol. 13. Providence, RI: American Mathematical Society. pp. 176–177. ISBN 978-0-8218-4942-2. MR 0194404.
  109. ^ The sequence of these primes, starting at rather than , is listed by Lava, Paolo Pietro; Balzarotti, Giorgio (2010). "Chapter 33. Formule fortunate". 103 curiosità matematiche: Teoria dei numeri, delle cifre e delle relazioni nella matematica contemporanea (in Italian). Ulrico Hoepli Editore S.p.A. p. 133. ISBN 9788820358044.
  110. ^ Chamberland, Marc (2015). "The Heegner numbers". Single Digits: In Praise of Small Numbers. Princeton University Press. pp. 213–215. ISBN 9781400865697.
  111. ^ a b Guy, Richard (2013). "A1 Prime values of quadratic functions". Unsolved Problems in Number Theory. Problem Books in Mathematics (3rd ed.). Springer. pp. 7–10. ISBN 9780387266770. {{cite book}}: Invalid |ref=harv (help)
  112. ^ Patterson, S. J. (1988). An introduction to the theory of the Riemann zeta-function. Cambridge Studies in Advanced Mathematics. Vol. 14. Cambridge University Press, Cambridge. p. 1. doi:10.1017/CBO9780511623707. ISBN 0-521-33535-3. MR 0933558. {{cite book}}: Invalid |ref=harv (help)
  113. ^ Borwein, Peter; Choi, Stephen; Rooney, Brendan; Weirathmueller, Andrea (2008). The Riemann hypothesis: A resource for the afficionado and virtuoso alike. CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC. New York: Springer. pp. 10–11. doi:10.1007/978-0-387-72126-2. ISBN 978-0-387-72125-5. MR 2463715. {{cite book}}: Invalid |ref=harv (help)
  114. ^ Sandifer 2007, pp. 191–193.
  115. ^ Borwein et al. 2008, Theorem 2.4, p. 14.
  116. ^ Tao, Terence (2013). "3.7 The Euler-Maclaurin formula, Bernoulli numbers, the zeta function, and real-variable analytic continuation". Compactness and Contradiction. Providence, RI: American Mathematical Society. pp. 88–104. doi:10.1090/mbk/081. ISBN 978-0-8218-9492-7. MR 3026767.
  117. ^ Sandifer 2007, Chapter 35, Estimating the Basel problem, pp. 205–208.
  118. ^ Ogilvy, C. S.; Anderson, J. T. (1988). Excursions in Number Theory. Dover Publications Inc. pp. 29–35. ISBN 0-486-25778-9.
  119. ^ Borwein et al. 2008, Conjecture 2.7 (the Riemann hypothesis), p. 15.
  120. ^ Patterson 1988, p. 7.
  121. ^ a b Borwein et al. 2008, [https://books.google.com/books?id=Qm1aZA-UwX4C&pg=PA18 p. 18.
  122. ^ Nathanson 2000, Chapter 9, The prime number theorem, pp. 289–324.
  123. ^ Zagier, Don (1977). "The first 50 million prime numbers". The Mathematical Intelligencer. 1 (S2): 7–19. doi:10.1007/bf03351556. See especially pp. 14–16.
  124. ^ Guy 2013, p. vii.
  125. ^ Guy 2013, C1 Goldbach's conjecture, pp. 105–107.
  126. ^ Oliveira e Silva, Tomás; Herzog, Siegfried; Pardi, Silvio (2014). "Empirical verification of the even Goldbach conjecture and computation of prime gaps up to ". Mathematics of Computation. 83 (288): 2033–2060. doi:10.1090/S0025-5718-2013-02787-1. MR 3194140.
  127. ^ Tao 2009, 3.1 Structure and randomness in the prime numbers, pp. 239–247. See especially p. 239.
  128. ^ Guy 2013, p. 159.
  129. ^ Ramaré, Olivier (1995). "On Šnirel'man's constant". Annali della Scuola Normale Superiore di Pisa. 22 (4): 645–706. MR 1375315.
  130. ^ Rassias, Michael Th. (2017). Goldbach's Problem: Selected Topics. Cham: Springer. p. vii. doi:10.1007/978-3-319-57914-6. ISBN 978-3-319-57912-2. MR 3674356.
  131. ^ Graham, R. L. (1964). "A Fibonacci-like sequence of composite nNumbers". Mathematics Magazine. 37 (5): 322–324. JSTOR 2689243. MR 1571455.
  132. ^ E.g., see Guy 2013, A3 Mersenne primes. Repunits. Fermat numbers. Primes of shape . pp. 13–21.
  133. ^ Crandall & Pomerance 2005, pp. 31–32.
  134. ^ Caldwell, Chris K.; Gallot, Yves (2002). "On the primality of and ". Mathematics of Computation. 71 (237): 441–448. doi:10.1090/S0025-5718-01-01315-1. MR 1863013.
  135. ^ a b c Ribenboim 2004, Gaps between primes, pp. 186–192.
  136. ^ a b Ribenboim 2004, p. 183.
  137. ^ Chan, Joel (February 1996). "Prime time!". Math Horizons. 3 (3): 23–25. JSTOR 25678057. Note that Chan lists Legendre's conjecture as "Sierpinski's Postulate".
  138. ^ Hardy, Godfrey Harold (1940). A Mathematician's Apology. Cambridge University Press. p. 140. ISBN 978-0-521-42706-7. No one has yet discovered any warlike purpose to be served by the theory of numbers or relativity, and it seems unlikely that anyone will do so for many years.
  139. ^ Kraft & Washington (2014), Proposition 5.3, p. 96.
  140. ^ Shahriari, Shahriar (2017). Algebra in Action: A Course in Groups, Rings, and Fields. Pure and Applied Undergraduate Texts. Vol. 27. American Mathematical Society. pp. 20–21. ISBN 9781470428495. {{cite book}}: Invalid |ref=harv (help)
  141. ^ Dudley 1978, Theorem 3, p. 28.
  142. ^ Shahriari 2017, pp. 27–28.
  143. ^ Ribenboim 2004, Fermat's little theorem and primitive roots modulo a prime, pp. 17–21.
  144. ^ Ribenboim 2004, The property of Giuga, pp. 21–22.
  145. ^ Ribenboim 2004, The theorem of Wilson, p. 21.
  146. ^ a b Křížek, Michal; Luca, Florian; Somer, Lawrence (2001). 17 Lectures on Fermat Numbers: From Number Theory to Geometry. CMS Books in Mathematics. Vol. 9. New York: Springer-Verlag. pp. 1–2. doi:10.1007/978-0-387-21850-2. ISBN 0-387-95332-9. MR 1866957.
  147. ^ Boklan, Kent D.; Conway, John H. (January 2017). "Expect at most one billionth of a new Fermat prime!". The Mathematical Intelligencer. 39 (1): 3–5. arXiv:1605.01371. doi:10.1007/s00283-016-9644-3.
  148. ^ Goles, E.; Schulz, O.; Markus, M. (2001). "Prime number selection of cycles in a predator-prey model". Complexity. 6 (4): 33–38. doi:10.1002/cplx.1040.
  149. ^ Paulo R. A. Campos; Viviane M. de Oliveira; Ronaldo Giro; Douglas S. Galvão. (2004). "Emergence of Prime Numbers as the Result of Evolutionary Strategy". Physical Review Letters. 93 (9): 098107. arXiv:q-bio/0406017. Bibcode:2004PhRvL..93i8107C. doi:10.1103/PhysRevLett.93.098107. {{cite journal}}: Unknown parameter |last-author-amp= ignored (|name-list-style= suggested) (help)CS1 maint: postscript (link)
  150. ^ "Invasion of the Brood". The Economist. May 6, 2004. Retrieved 2006-11-26.
  151. ^ Ivars Peterson (June 28, 1999). "The Return of Zeta". MAA Online. Archived from the original on October 20, 2007. Retrieved 2008-03-14. {{cite web}}: Unknown parameter |deadurl= ignored (|url-status= suggested) (help)
  152. ^ Lang, Serge (2002). Algebra. Graduate Texts in Mathematics. Vol. 211. Berlin, New York: Springer-Verlag. ISBN 978-0-387-95385-4. MR 1878556., Section II.1, p. 90
  153. ^ Schubert, H. "Die eindeutige Zerlegbarkeit eines Knotens in Primknoten". S.-B Heidelberger Akad. Wiss. Math.-Nat. Kl. 1949 (1949), 57–104.
  154. ^ Eisenbud, David (1995). Commutative Algebra. Graduate Texts in Mathematics. Vol. 150. Berlin, New York: Springer-Verlag. Section 3.3. ISBN 978-0-387-94268-1. MR 1322960.
  155. ^ Shafarevich, Basic Algebraic Geometry volume 2 (Schemes and Complex Manifolds), p. 5, section V.1
  156. ^ Neukirch, Jürgen (1999). Algebraic Number Theory. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Vol. 322. Berlin: Springer-Verlag. Section I.8, p. 50. doi:10.1007/978-3-662-03983-0. ISBN 3-540-65399-6. MR 1697859. {{cite book}}: Invalid |ref=harv (help)
  157. ^ Neukirch 1999, Section I.7, p. 38
  158. ^ Endler, Valuation Theory, p. 1
  159. ^ Some sources also put .
  160. ^ Gouvea: p-adic numbers: an introduction, Chapter 3, p. 43
  161. ^ Hill, Peter Jensen, ed. (1995). The Messiaen companion. Portland, Or: Amadeus Press. Ex. 13.2 Messe de la Pentecôte 1 'Entrée'. ISBN 978-0-931340-95-6.
  162. ^ Carl Pomerance, Prime Numbers and the Search for Extraterrestrial Intelligence, Retrieved on December 22, 2007
  163. ^ Mark Sarvas, Book Review: The Curious Incident of the Dog in the Night-Time Archived 2013-04-02 at the Wayback Machine, at The Modern Word Archived 2012-01-31 at the Wayback Machine, Retrieved on March 30, 2012
  164. ^ "Introducing Paolo Giordano". Books Quarterly.[dead link]

External links

Template:Wikinewspar2

Prime number generators and calculators