HOME

TheInfoList



OR:

Euclid's theorem is a fundamental statement in number theory that asserts that there are infinitely many prime numbers. It was first proved by Euclid in his work ''
Elements Element or elements may refer to: Science * Chemical element, a pure substance of one type of atom * Heating element, a device that generates heat by electrical resistance * Orbital elements, parameters required to identify a specific orbit of ...
''. There are several proofs of the theorem.


Euclid's proof

Euclid offered a proof published in his work ''Elements'' (Book IX, Proposition 20), which is paraphrased here. Consider any finite list of prime numbers ''p''1, ''p''2, ..., ''p''''n''. It will be shown that at least one additional prime number not in this list exists. Let ''P'' be the product of all the prime numbers in the list: ''P'' = ''p''1''p''2...''p''''n''. Let ''q'' = ''P'' + 1. Then ''q'' is either prime or not: *If ''q'' is prime, then there is at least one more prime that is not in the list, namely, ''q'' itself. *If ''q'' is not prime, then some
prime factor A prime number (or a prime) is a natural number greater than 1 that is not a product of 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 ...
''p'' divides ''q''. If this factor ''p'' were in our list, then it would divide ''P'' (since ''P'' is the product of every number in the list); but ''p'' also divides ''P'' + 1 = ''q'', as just stated. If ''p'' divides ''P'' and also ''q,'' then ''p'' must also divide the difference of the two numbers, which is (''P'' + 1) − ''P'' or just 1. Since no prime number divides 1, ''p'' cannot be in the list. This means that at least one more prime number exists beyond those in the list. This proves that for every finite list of prime numbers there is a prime number not in the list. In the original work, as Euclid had no way of writing an arbitrary list of primes, he used a method that he frequently applied, that is, the method of generalizable example. Namely, he picks just three primes and using the general method outlined above, proves that he can always find an additional prime. Euclid presumably assumes that his readers are convinced that a similar proof will work, no matter how many primes are originally picked. Euclid is often erroneously reported to have proved this result by contradiction beginning with the assumption that the finite set initially considered contains all prime numbers, though it is actually a
proof by cases Proof by exhaustion, also known as proof by cases, proof by case analysis, complete induction or the brute force method, is a method of mathematical proof in which the statement to be proved is split into a finite number of cases or sets of equiv ...
, a
direct proof In mathematics and logic, a direct proof is a way of showing the truth or falsehood of a given statement by a straightforward combination of established facts, usually axioms, existing lemmas and theorems, without making any further assumptions. ...
method. The philosopher Torkel Franzén, in a book on logic, states, "Euclid's proof that there are infinitely many primes is not an indirect proof ..The argument is sometimes formulated as an indirect proof by replacing it with the assumption 'Suppose are all the primes'. However, since this assumption isn't even used in the proof, the reformulation is pointless."


Variations

Several variations on Euclid's proof exist, including the following: The
factorial In mathematics, the factorial of a non-negative denoted is the product of all positive integers less than or equal The factorial also equals the product of n with the next smaller factorial: \begin n! &= n \times (n-1) \times (n-2) \t ...
''n''! of a positive integer ''n'' is divisible by every integer from 2 to ''n'', as it is the product of all of them. Hence, is not divisible by any of the integers from 2 to ''n'', inclusive (it gives a remainder of 1 when divided by each). Hence is either prime or divisible by a prime larger than ''n''. In either case, for every positive integer ''n'', there is at least one prime bigger than ''n''. The conclusion is that the number of primes is infinite.


Euler's proof

Another proof, by the Swiss mathematician Leonhard Euler, relies on the fundamental theorem of arithmetic: that every integer has a unique prime factorization. What Euler wrote (not with this modern notation and, unlike modern standards, not restricting the arguments in sums and products to any finite sets of integers) is equivalent to the statement that we have : \prod_ \frac=\sum_\frac, where P_k denotes the set of the first prime numbers, and N_k is the set of the positive integers whose prime factors are all in P_k. In order to show this, one expands each factor in the product as a geometric series, and distributes the product over the sum (this is a special case of the Euler product formula for the Riemann zeta function). : \begin \prod_ \frac & =\prod_ \sum_ \frac\\ & = \left(\sum_ \frac\right) \cdot \left(\sum_ \frac\right) \cdot \left(\sum_ \frac\right) \cdot \left(\sum_ \frac\right)\cdots \\ & =\sum_ \frac \\ & =\sum_\frac. \end In the penultimate sum every product of primes appears exactly once, and so the last equality is true by the fundamental theorem of arithmetic. In his first corollary to this result Euler denotes by a symbol similar to \infty the « absolute infinity » and writes that the infinite sum in the statement equals the « value » \log\infty, to which the infinite product is thus also equal (in modern terminology this is equivalent to say that the partial sum up to x of the harmonic series diverges asymptotically like \log x). Then in his second corollary Euler notes that the product : \prod_ \frac converges to the finite value 2, and that there are consequently more primes than squares (« sequitur infinities plures esse numeros primos »). This proves Euclid Theorem. In the same paper (Theorem 19) Euler in fact used the above equality to prove a much stronger theorem that was unknown before him, namely that the series :\sum_\frac 1p is divergent, where denotes the set of all prime numbers (Euler writes that the infinite sum =\log\log\infty, which in modern terminology is equivalent to say that the partial sum up to x of this series behaves asymptotically like \log\log x).


Erdős's proof

Paul Erdős Paul Erdős ( hu, Erdős Pál ; 26 March 1913 – 20 September 1996) was a Hungarian mathematician. He was one of the most prolific mathematicians and producers of mathematical conjectures of the 20th century. pursued and proposed problems in ...
gave a proof that also relies on the fundamental theorem of arithmetic. Every positive integer has a unique factorization into a square-free number and a square number . For example, . Let be a positive integer, and let be the number of primes less than or equal to . Call those primes . Any positive integer which is less than or equal to can then be written in the form :\left( p_1^ p_2^ \cdots p_k^ \right) s^2, where each is either or . There are ways of forming the square-free part of . And can be at most , so . Thus, at most numbers can be written in this form. In other words, :N \leq 2^k \sqrt. Or, rearranging, , the number of primes less than or equal to , is greater than or equal to . Since was arbitrary, can be as large as desired by choosing appropriately.


Furstenberg's proof

In the 1950s, Hillel Furstenberg introduced a proof by contradiction using point-set topology. Define a topology on the integers Z, called the evenly spaced integer topology, by declaring a subset ''U'' ⊆ Z to be an open set if and only if it is either the
empty set In mathematics, the empty set is the unique set having no elements; its size or cardinality (count of elements in a set) is zero. Some axiomatic set theories ensure that the empty set exists by including an axiom of empty set, while in other ...
, ∅, or it is a union of arithmetic sequences ''S''(''a'', ''b'') (for ''a'' ≠ 0), where :S(a, b) = \ = a \mathbb + b. Then a contradiction follows from the property that a finite set of integers cannot be open and the property that the basis sets ''S''(''a'', ''b'') are both open and closed, since :\mathbb \setminus \ = \bigcup_ S(p, 0) cannot be closed because its complement is finite, but is closed since it is a finite union of closed sets.


Recent proofs


Proof using the inclusion-exclusion principle

Juan Pablo Pinasco has written the following proof. Let ''p''1, ..., ''p''''N'' be the smallest ''N'' primes. Then by the inclusion–exclusion principle, the number of positive integers less than or equal to ''x'' that are divisible by one of those primes is : \begin 1 + \sum_ \left\lfloor \frac \right\rfloor - \sum_ \left\lfloor \frac \right\rfloor & + \sum_ \left\lfloor \frac \right\rfloor - \cdots \\ & \cdots \pm (-1)^ \left\lfloor \frac \right\rfloor. \qquad (1) \end Dividing by ''x'' and letting ''x'' → ∞ gives : \sum_ \frac - \sum_ \frac + \sum_ \frac - \cdots \pm (-1)^ \frac. \qquad (2) This can be written as : 1 - \prod_^N \left( 1 - \frac \right). \qquad (3) If no other primes than ''p''1, ..., ''p''''N'' exist, then the expression in (1) is equal to \lfloor x \rfloor and the expression in (2) is equal to 1, but clearly the expression in (3) is not equal to 1. Therefore, there must be more primes than  ''p''1, ..., ''p''''N''.


Proof using de Polignac's formula

In 2010, Junho Peter Whang published the following proof by contradiction. Let ''k'' be any positive integer. Then according to de Polignac's formula (actually due to Legendre) : k! = \prod_ p^ where : f(p,k) = \left\lfloor \frac \right\rfloor + \left\lfloor \frac \right\rfloor + \cdots. : f(p,k) < \frac + \frac + \cdots = \frac \le k. But if only finitely many primes exist, then : \lim_ \frac = 0, (the numerator of the fraction would grow singly exponentially while by Stirling's approximation the denominator grows more quickly than singly exponentially), contradicting the fact that for each ''k'' the numerator is greater than or equal to the denominator.


Proof by construction

Filip Saidak gave the following proof by construction, which does not use reductio ad absurdum or Euclid's lemma (that if a prime ''p'' divides ''ab'' then it must divide ''a'' or ''b''). Since each natural number (> 1) has at least one prime factor, and two successive numbers ''n'' and (''n'' + 1) have no factor in common, the product ''n''(''n'' + 1) has more different prime factors than the number ''n'' itself.  So the chain of pronic numbers:
1×2 = 2 ,    2×3 = 6 ,    6×7 = 42 ,    42×43 = 1806 ,    1806×1807 = 3263442 , · · ·
provides a sequence of unlimited growing sets of primes.


Proof using the

incompressibility method In mathematics, the incompressibility method is a proof method like the probabilistic method, the counting method or the pigeonhole principle. To prove that an object in a certain class (on average) satisfies a certain property, select an object o ...

Suppose there were only ''k'' primes (''p''1, ..., ''p''''k''). By the fundamental theorem of arithmetic, any positive integer ''n'' could then be represented as n = ^ ^ \cdots ^, where the non-negative integer exponents ''e''''i'' together with the finite-sized list of primes are enough to reconstruct the number. Since p_i \geq 2 for all ''i'', it follows that e_i \leq \lg n for all ''i'' (where \lg denotes the base-2 logarithm). This yields an encoding for ''n'' of the following size (using
big O notation Big ''O'' notation is a mathematical notation that describes the limiting behavior of a function when the argument tends towards a particular value or infinity. Big O is a member of a family of notations invented by Paul Bachmann, Edmund Lan ...
): :O(\text + k \lg \lg n) = O(\lg \lg n) bits. This is a much more efficient encoding than representing ''n'' directly in binary, which takes N = O(\lg n) bits. An established result in lossless data compression states that one cannot generally compress ''N'' bits of information into fewer than ''N'' bits. The representation above violates this by far when ''n'' is large enough since \lg \lg n = o(\lg n). Therefore, the number of primes must not be finite.


Stronger results

The theorems in this section simultaneously imply Euclid's theorem and other results.


Dirichlet's theorem on arithmetic progressions

Dirichlet's theorem states that for any two positive coprime integers ''a'' and ''d'', there are infinitely many primes of the form ''a'' + ''nd'', where ''n'' is also a positive integer. In other words, there are infinitely many primes that are congruent to ''a''
modulo In computing, the modulo operation returns the remainder or signed remainder of a division, after one number is divided by another (called the '' modulus'' of the operation). Given two positive numbers and , modulo (often abbreviated as ) is t ...
''d''.


Prime number theorem

Let be the
prime-counting function In mathematics, the prime-counting function is the function counting the number of prime numbers less than or equal to some real number ''x''. It is denoted by (''x'') (unrelated to the number ). History Of great interest in number theory is t ...
that gives the number of primes less than or equal to , for any real number . The prime number theorem then states that is a good approximation to , in the sense that the
limit Limit or Limits may refer to: Arts and media * ''Limit'' (manga), a manga by Keiko Suenobu * ''Limit'' (film), a South Korean film * Limit (music), a way to characterize harmony * "Limit" (song), a 2016 single by Luna Sea * "Limits", a 2019 ...
of the ''quotient'' of the two functions and as increases without bound is 1: :\lim_ \frac=1. Using asymptotic notation this result can be restated as :\pi(x)\sim \frac. This yields Euclid's theorem, since \lim_ \frac=\infty.


Bertrand–Chebyshev theorem

In number theory,
Bertrand's postulate In number theory, Bertrand's postulate is a theorem stating that for any integer n > 3, there always exists at least one prime number p with :n < p < 2n - 2. A less restrictive formulation is: for every n > 1, there is always ...
is a theorem stating that for any integer n > 1, there always exists at least one prime number such that :n < p < 2n. Bertrand–Chebyshev theorem can also be stated as a relationship with \pi(x), where \pi(x) is the
prime-counting function In mathematics, the prime-counting function is the function counting the number of prime numbers less than or equal to some real number ''x''. It is denoted by (''x'') (unrelated to the number ). History Of great interest in number theory is t ...
(number of primes less than or equal to x \,): :\pi(x) - \pi(\tfrac) \ge 1, for all x \ge 2. This statement was first conjectured in 1845 by Joseph Bertrand (1822–1900). Bertrand himself verified his statement for all numbers in the interval His conjecture was completely proved by
Chebyshev Pafnuty Lvovich Chebyshev ( rus, Пафну́тий Льво́вич Чебышёв, p=pɐfˈnutʲɪj ˈlʲvovʲɪtɕ tɕɪbɨˈʂof) ( – ) was a Russian mathematician and considered to be the founding father of Russian mathematics. Chebyshe ...
(1821–1894) in 1852. (Proof of the postulate: 371–382). Also see Mémoires de l'Académie Impériale des Sciences de St. Pétersbourg, vol. 7, pp. 15–33, 1854 and so the postulate is also called the Bertrand–Chebyshev theorem or Chebyshev's theorem.


Notes and references


External links

*
Euclid's Elements, Book IX, Prop. 20
(Euclid's proof, on David Joyce's website at Clark University) {{Ancient Greek mathematics Articles containing proofs Theorems about prime numbers