Divisor function

Last updated
Divisor function s0(n) up to n = 250 Divisor.svg
Divisor function σ0(n) up to n = 250
Sigma function s1(n) up to n = 250 Sigma function.svg
Sigma function σ1(n) up to n = 250
Sum of the squares of divisors, s2(n), up to n = 250 Divisor square.svg
Sum of the squares of divisors, σ2(n), up to n = 250
Sum of cubes of divisors, s3(n) up to n = 250 Divisor cube.svg
Sum of cubes of divisors, σ3(n) up to n = 250

In mathematics, and specifically in number theory, a divisor function is an arithmetic function related to the divisors of an integer. When referred to as the divisor function, it counts the number of divisors of an integer (including 1 and the number itself). It appears in a number of remarkable identities, including relationships on the Riemann zeta function and the Eisenstein series of modular forms. Divisor functions were studied by Ramanujan, who gave a number of important congruences and identities; these are treated separately in the article Ramanujan's sum.

Contents

A related function is the divisor summatory function, which, as the name implies, is a sum over the divisor function.

Definition

The sum of positive divisors function σz(n), for a real or complex number z, is defined as the sum of the zth powers of the positive divisors of n. It can be expressed in sigma notation as

where is shorthand for "d divides n". The notations d(n), ν(n) and τ(n) (for the German Teiler = divisors) are also used to denote σ0(n), or the number-of-divisors function [1] [2] ( OEIS:  A000005 ). When z is 1, the function is called the sigma function or sum-of-divisors function, [1] [3] and the subscript is often omitted, so σ(n) is the same as σ1(n) ( OEIS:  A000203 ).

The aliquot sum s(n) of n is the sum of the proper divisors (that is, the divisors excluding n itself, OEIS:  A001065 ), and equals σ1(n)  n; the aliquot sequence of n is formed by repeatedly applying the aliquot sum function.

Example

For example, σ0(12) is the number of the divisors of 12:

while σ1(12) is the sum of all the divisors:

and the aliquot sum s(12) of proper divisors is:

σ-1(n) is sometimes called the abundancy index of n, and we have:

Table of values

The cases x = 2 to 5 are listed in OEIS:  A001157 through OEIS:  A001160 , x = 6 to 24 are listed in OEIS:  A013954 through OEIS:  A013972 .

nfactorization𝜎0(n)𝜎1(n)𝜎2(n)𝜎3(n)𝜎4(n)
1111111
22235917
3324102882
422372173273
552626126626
62×3412502521394
7728503442402
823415855854369
932313917576643
102×5418130113410642
1111212122133214642
1222×3628210204422386
1313214170219828562
142×7424250309640834
153×5424260352851332
1624531341468169905
1717218290491483522
182×326394556813112931
19192203626860130322
2022×56425469198170898
213×74325009632196964
222×1143661011988248914
232322453012168279842
2423×386085016380358258
255233165115751391251
262×1344285019782485554
273344082020440538084
2822×7656105025112655746
292923084224390707282
302×3×5872130031752872644
313123296229792923522
32256631365374491118481
333×114481220372961200644
342×174541450442261419874
355×74481300433441503652
3622×329911911552611813539
37372381370506541874162
382×194601810617402215474
393×134561700615442342084
4023×58902210737102734994
41412421682689222825762
422×3×78962500866883348388
43432441850795083418802
4422×116842562972363997266
4532×56782366953824158518
462×2347226501095124757314
474724822101038244879682
4824×31012434101310685732210
497235724511179935767203
502×5269332551417596651267

Properties

Formulas at prime powers

For a prime number p,

because by definition, the factors of a prime number are 1 and itself. Also, where pn# denotes the primorial,

since n prime factors allow a sequence of binary selection ( or 1) from n terms for each proper divisor formed. However, these are not in general the smallest numbers whose number of divisors is a power of two; instead, the smallest such number may be obtained by multiplying together the first n Fermi–Dirac primes, prime powers whose exponent is a power of two. [4]

Clearly, for all , and for all , .

The divisor function is multiplicative (since each divisor c of the product mn with distinctively correspond to a divisor a of m and a divisor b of n), but not completely multiplicative:

The consequence of this is that, if we write

where r = ω(n) is the number of distinct prime factors of n, pi is the ith prime factor, and ai is the maximum power of pi by which n is divisible, then we have: [5]

which, when x  0, is equivalent to the useful formula: [5]

When x = 0, is: [5]

This result can be directly deduced from the fact that all divisors of are uniquely determined by the distinct tuples of integers with (i.e. independent choices for each ).

For example, if n is 24, there are two prime factors (p1 is 2; p2 is 3); noting that 24 is the product of 23×31, a1 is 3 and a2 is 1. Thus we can calculate as so:

The eight divisors counted by this formula are 1, 2, 4, 8, 3, 6, 12, and 24.

Other properties and identities

Euler proved the remarkable recurrence: [6] [7] [8]

where if it occurs and for , and are consecutive pairs of generalized pentagonal numbers ( OEIS:  A001318 , starting at offset 1). Indeed, Euler proved this by logarithmic differentiation of the identity in his pentagonal number theorem.

For a non-square integer, n, every divisor, d, of n is paired with divisor n/d of n and is even; for a square integer, one divisor (namely ) is not paired with a distinct divisor and is odd. Similarly, the number is odd if and only if n is a square or twice a square. [9]

We also note s(n) = σ(n)  n. Here s(n) denotes the sum of the proper divisors of n, that is, the divisors of n excluding n itself. This function is used to recognize perfect numbers, which are the n such that s(n) = n. If s(n) > n, then n is an abundant number, and if s(n) < n, then n is a deficient number.

If n is a power of 2, , then and , which makes n almost-perfect.

As an example, for two primes , let

.

Then

and

where is Euler's totient function.

Then, the roots of

express p and q in terms of σ(n) and φ(n) only, requiring no knowledge of n or , as

Also, knowing n and either or , or, alternatively, and either or allows an easy recovery of p and q.

In 1984, Roger Heath-Brown proved that the equality

is true for infinitely many values of n, see OEIS:  A005237 .

Series relations

Two Dirichlet series involving the divisor function are: [10]

where is the Riemann zeta function. The series for d(n) = σ0(n) gives: [10]

and a Ramanujan identity [11]

which is a special case of the Rankin–Selberg convolution.

A Lambert series involving the divisor function is: [12]

for arbitrary complex |q|  1 and a. This summation also appears as the Fourier series of the Eisenstein series and the invariants of the Weierstrass elliptic functions.

For , there is an explicit series representation with Ramanujan sums as : [13]

The computation of the first terms of shows its oscillations around the "average value" :

Growth rate

In little-o notation, the divisor function satisfies the inequality: [14] [15]

More precisely, Severin Wigert showed that: [15]

On the other hand, since there are infinitely many prime numbers, [15]

In Big-O notation, Peter Gustav Lejeune Dirichlet showed that the average order of the divisor function satisfies the following inequality: [16] [17]

where is Euler's gamma constant. Improving the bound in this formula is known as Dirichlet's divisor problem.

The behaviour of the sigma function is irregular. The asymptotic growth rate of the sigma function can be expressed by: [18]

where lim sup is the limit superior. This result is Grönwall's theorem, published in 1913 ( Grönwall 1913 ). His proof uses Mertens' third theorem, which says that:

where p denotes a prime.

In 1915, Ramanujan proved that under the assumption of the Riemann hypothesis, Robin's inequality

(where γ is the Euler–Mascheroni constant)

holds for all sufficiently large n( Ramanujan 1997 ). The largest known value that violates the inequality is n=5040. In 1984, Guy Robin proved that the inequality is true for all n > 5040 if and only if the Riemann hypothesis is true ( Robin 1984 ). This is Robin's theorem and the inequality became known after him. Robin furthermore showed that if the Riemann hypothesis is false then there are an infinite number of values of n that violate the inequality, and it is known that the smallest such n > 5040 must be superabundant ( Akbary & Friggstad 2009 ). It has been shown that the inequality holds for large odd and square-free integers, and that the Riemann hypothesis is equivalent to the inequality just for n divisible by the fifth power of a prime ( Choie et al. 2007 ).

Robin also proved, unconditionally, that the inequality:

holds for all n ≥ 3.

A related bound was given by Jeffrey Lagarias in 2002, who proved that the Riemann hypothesis is equivalent to the statement that:

for every natural number n> 1, where is the nth harmonic number, ( Lagarias 2002 ).

See also

Notes

  1. 1 2 Long (1972 , p. 46)
  2. Pettofrezzo & Byrkit (1970 , p. 63)
  3. Pettofrezzo & Byrkit (1970 , p. 58)
  4. Ramanujan, S. (1915), "Highly Composite Numbers", Proceedings of the London Mathematical Society, s2-14 (1): 347–409, doi:10.1112/plms/s2_14.1.347 ; see section 47, pp. 405–406, reproduced in Collected Papers of Srinivasa Ramanujan, Cambridge Univ. Press, 2015, pp. 124–125
  5. 1 2 3 Hardy & Wright (2008), pp. 310 f, §16.7.
  6. Euler, Leonhard; Bell, Jordan (2004). "An observation on the sums of divisors". arXiv: math/0411587 .
  7. https://scholarlycommons.pacific.edu/euler-works/175/, Découverte d'une loi tout extraordinaire des nombres par rapport à la somme de leurs diviseurs
  8. https://scholarlycommons.pacific.edu/euler-works/542/, De mirabilis proprietatibus numerorum pentagonalium
  9. Gioia & Vaidya (1967).
  10. 1 2 Hardy & Wright (2008), pp. 326–328, §17.5.
  11. Hardy & Wright (2008), pp. 334–337, §17.8.
  12. Hardy & Wright (2008), pp. 338–341, §17.10.
  13. E. Krätzel (1981). Zahlentheorie. Berlin: VEB Deutscher Verlag der Wissenschaften. p. 130. (German)
  14. Apostol (1976), p. 296.
  15. 1 2 3 Hardy & Wright (2008), pp. 342–347, §18.1.
  16. Apostol (1976), Theorem 3.3.
  17. Hardy & Wright (2008), pp. 347–350, §18.2.
  18. Hardy & Wright (2008), pp. 469–471, §22.9.

Related Research Articles

In number theory, an arithmetic, arithmetical, or number-theoretic function is for most authors any function f(n) whose domain is the positive integers and whose range is a subset of the complex numbers. Hardy & Wright include in their definition the requirement that an arithmetical function "expresses some arithmetical property of n".

In number theory, a multiplicative function is an arithmetic function f(n) of a positive integer n with the property that f(1) = 1 and

<span class="mw-page-title-main">Riemann zeta function</span> Analytic function in mathematics

The Riemann zeta function or Euler–Riemann zeta function, denoted by the Greek letter ζ (zeta), is a mathematical function of a complex variable defined as

<span class="mw-page-title-main">Euler's totient function</span> Number of integers coprime to and not exceeding n

In number theory, Euler's totient function counts the positive integers up to a given integer n that are relatively prime to n. It is written using the Greek letter phi as or , and may also be called Euler's phi function. In other words, it is the number of integers k in the range 1 ≤ kn for which the greatest common divisor gcd(n, k) is equal to 1. The integers k of this form are sometimes referred to as totatives of n.

The Liouville lambda function, denoted by λ(n) and named after Joseph Liouville, is an important arithmetic function. Its value is +1 if n is the product of an even number of prime numbers, and −1 if it is the product of an odd number of primes.

<span class="mw-page-title-main">Harmonic number</span> Sum of the first n whole number reciprocals; 1/1 + 1/2 + 1/3 + ... + 1/n

In mathematics, the n-th harmonic number is the sum of the reciprocals of the first n natural numbers:

In mathematics, a Dirichlet series is any series of the form

<span class="mw-page-title-main">Clausen function</span> Transcendental single-variable function

In mathematics, the Clausen function, introduced by Thomas Clausen (1832), is a transcendental, special function of a single variable. It can variously be expressed in the form of a definite integral, a trigonometric series, and various other forms. It is intimately connected with the polylogarithm, inverse tangent integral, polygamma function, Riemann zeta function, Dirichlet eta function, and Dirichlet beta function.

Eisenstein series, named after German mathematician Gotthold Eisenstein, are particular modular forms with infinite series expansions that may be written down directly. Originally defined for the modular group, Eisenstein series can be generalized in the theory of automorphic forms.

In mathematics, the Lerch zeta function, sometimes called the Hurwitz–Lerch zeta function, is a special function that generalizes the Hurwitz zeta function and the polylogarithm. It is named after Czech mathematician Mathias Lerch, who published a paper about the function in 1887.

In mathematics, the explicit formulae for L-functions are relations between sums over the complex number zeroes of an L-function and sums over prime powers, introduced by Riemann (1859) for the Riemann zeta function. Such explicit formulae have been applied also to questions on bounding the discriminant of an algebraic number field, and the conductor of a number field.

In number theory, the Dedekind psi function is the multiplicative function on the positive integers defined by

In mathematics, a natural number a is a unitary divisor of a number b if a is a divisor of b and if a and are coprime, having no common factor other than 1. Thus, 5 is a unitary divisor of 60, because 5 and have only 1 as a common factor, while 6 is a divisor but not a unitary divisor of 60, as 6 and have a common factor other than 1, namely 2. 1 is a unitary divisor of every natural number.

In number theory, Ramanujan's sum, usually denoted cq(n), is a function of two positive integer variables q and n defined by the formula

In mathematics, the multiple zeta functions are generalizations of the Riemann zeta function, defined by

In number theory, an average order of an arithmetic function is some simpler or better-understood function which takes the same values "on average".

<span class="mw-page-title-main">Riemann hypothesis</span> Conjecture on zeros of the zeta function

In mathematics, the Riemann hypothesis is the conjecture that the Riemann zeta function has its zeros only at the negative even integers and complex numbers with real part 1/2. Many consider it to be the most important unsolved problem in pure mathematics. It is of great interest in number theory because it implies results about the distribution of prime numbers. It was proposed by Bernhard Riemann (1859), after whom it is named.

<span class="mw-page-title-main">Anatoly Karatsuba</span> Russian mathematician

Anatoly Alexeyevich Karatsuba was a Russian mathematician working in the field of analytic number theory, p-adic numbers and Dirichlet series.

In number theory, the prime omega functions and count the number of prime factors of a natural number Thereby counts each distinct prime factor, whereas the related function counts the total number of prime factors of honoring their multiplicity. That is, if we have a prime factorization of of the form for distinct primes , then the respective prime omega functions are given by and . These prime factor counting functions have many important number theoretic relations.

In number theory, the totient summatory function is a summatory function of Euler's totient function defined by:

References