Partition function (number theory)

Last updated
The values
p
(
1
)
,
...
,
p
(
8
)
{\displaystyle p(1),\dots ,p(8)}
of the partition function (1, 2, 3, 5, 7, 11, 15, and 22) can be determined by counting the Young diagrams for the partitions of the numbers from 1 to 8. Ferrer partitioning diagrams.svg
The values of the partition function (1, 2, 3, 5, 7, 11, 15, and 22) can be determined by counting the Young diagrams for the partitions of the numbers from 1 to 8.

In number theory, the partition functionp(n) represents the number of possible partitions of a non-negative integer n. For instance, p(4) = 5 because the integer 4 has the five partitions 1 + 1 + 1 + 1, 1 + 1 + 2, 1 + 3, 2 + 2, and 4.

Contents

No closed-form expression for the partition function is known, but it has both asymptotic expansions that accurately approximate it and recurrence relations by which it can be calculated exactly. It grows as an exponential function of the square root of its argument. The multiplicative inverse of its generating function is the Euler function; by Euler's pentagonal number theorem this function is an alternating sum of pentagonal number powers of its argument.

Srinivasa Ramanujan first discovered that the partition function has nontrivial patterns in modular arithmetic, now known as Ramanujan's congruences. For instance, whenever the decimal representation of n ends in the digit 4 or 9, the number of partitions of n will be divisible by 5.

Definition and examples

For a positive integer n, p(n) is the number of distinct ways of representing n as a sum of positive integers. For the purposes of this definition, the order of the terms in the sum is irrelevant: two sums with the same terms in a different order are not considered to be distinct.

By convention p(0) = 1, as there is one way (the empty sum) of representing zero as a sum of positive integers. Furthermore p(n) = 0 when n is negative.

The first few values of the partition function, starting with p(0) = 1, are:

1, 1, 2, 3, 5, 7, 11, 15, 22, 30, 42, 56, 77, 101, 135, 176, 231, 297, 385, 490, 627, 792, 1002, 1255, 1575, 1958, 2436, 3010, 3718, 4565, 5604, ... (sequence A000041 in the OEIS).

Some exact values of p(n) for larger values of n include: [1]

As of June 2022, the largest known prime number among the values of p(n) is p(1289844341), with 40,000 decimal digits. [2] [3] Until March 2022, this was also the largest prime that has been proved using elliptic curve primality proving. [4]

Generating function

Using Euler's method to find p(40): A ruler with plus and minus signs (grey box) is slid downwards, the relevant terms added or subtracted. The positions of the signs are given by differences of alternating natural (blue) and odd (orange) numbers. In the SVG file, hover over the image to move the ruler. Euler partition function.svg
Using Euler's method to find p(40): A ruler with plus and minus signs (grey box) is slid downwards, the relevant terms added or subtracted. The positions of the signs are given by differences of alternating natural (blue) and odd (orange) numbers. In the SVG file, hover over the image to move the ruler.

The generating function for p(n) is given by [5]

The equality between the products on the first and second lines of this formula is obtained by expanding each factor into the geometric series To see that the expanded product equals the sum on the first line, apply the distributive law to the product. This expands the product into a sum of monomials of the form for some sequence of coefficients , only finitely many of which can be non-zero. The exponent of the term is , and this sum can be interpreted as a representation of as a partition into copies of each number . Therefore, the number of terms of the product that have exponent is exactly , the same as the coefficient of in the sum on the left. Therefore, the sum equals the product.

The function that appears in the denominator in the third and fourth lines of the formula is the Euler function. The equality between the product on the first line and the formulas in the third and fourth lines is Euler's pentagonal number theorem. The exponents of in these lines are the pentagonal numbers for (generalized somewhat from the usual pentagonal numbers, which come from the same formula for the positive values of ). The pattern of positive and negative signs in the third line comes from the term in the fourth line: even choices of produce positive terms, and odd choices produce negative terms.

More generally, the generating function for the partitions of into numbers selected from a set of positive integers can be found by taking only those terms in the first product for which . This result is due to Leonhard Euler. [6] The formulation of Euler's generating function is a special case of a -Pochhammer symbol and is similar to the product formulation of many modular forms, and specifically the Dedekind eta function.

Recurrence relations

The same sequence of pentagonal numbers appears in a recurrence relation for the partition function: [7]

As base cases, is taken to equal , and is taken to be zero for negative . Although the sum on the right side appears infinite, it has only finitely many nonzero terms, coming from the nonzero values of in the range

Another recurrence relation for can be given in terms of the sum of divisors function σ: [8]

If denotes the number of partitions of with no repeated parts then it follows by splitting each partition into its even parts and odd parts, and dividing the even parts by two, that [9]

Congruences

Srinivasa Ramanujan is credited with discovering that the partition function has nontrivial patterns in modular arithmetic. For instance the number of partitions is divisible by five whenever the decimal representation of ends in the digit 4 or 9, as expressed by the congruence [10]

For instance, the number of partitions for the integer 4 is 5. For the integer 9, the number of partitions is 30; for 14 there are 135 partitions. This congruence is implied by the more general identity

also by Ramanujan, [11] [12] where the notation denotes the product defined by

A short proof of this result can be obtained from the partition function generating function.

Ramanujan also discovered congruences modulo 7 and 11: [10]

The first one comes from Ramanujan's identity [12]

Since 5, 7, and 11 are consecutive primes, one might think that there would be an analogous congruence for the next prime 13, for some a. However, there is no congruence of the form for any prime b other than 5, 7, or 11. [13] Instead, to obtain a congruence, the argument of should take the form for some . In the 1960s, A. O. L. Atkin of the University of Illinois at Chicago discovered additional congruences of this form for small prime moduli. For example:

KenOno  ( 2000 ) proved that there are such congruences for every prime modulus greater than 3. Later, Ahlgren & Ono (2001) showed there are partition congruences modulo every integer coprime to 6. [14] [15]

Approximation formulas

Approximation formulas exist that are faster to calculate than the exact formula given above.

An asymptotic expression for p(n) is given by

as .

This asymptotic formula was first obtained by G. H. Hardy and Ramanujan in 1918 and independently by J. V. Uspensky in 1920. Considering , the asymptotic formula gives about , reasonably close to the exact answer given above (1.415% larger than the true value).

Hardy and Ramanujan obtained an asymptotic expansion with this approximation as the first term: [16]

where

Here, the notation means that the sum is taken only over the values of that are relatively prime to . The function is a Dedekind sum.

The error after terms is of the order of the next term, and may be taken to be of the order of . As an example, Hardy and Ramanujan showed that is the nearest integer to the sum of the first terms of the series. [16]

In 1937, Hans Rademacher was able to improve on Hardy and Ramanujan's results by providing a convergent series expression for . It is [17] [18]

The proof of Rademacher's formula involves Ford circles, Farey sequences, modular symmetry and the Dedekind eta function.

It may be shown that the th term of Rademacher's series is of the order

so that the first term gives the Hardy–Ramanujan asymptotic approximation. PaulErdős  ( 1942 ) published an elementary proof of the asymptotic formula for . [19] [20]

Techniques for implementing the Hardy–Ramanujan–Rademacher formula efficiently on a computer are discussed by Johansson (2012), who shows that can be computed in time for any . This is near-optimal in that it matches the number of digits of the result. [21] The largest value of the partition function computed exactly is , which has slightly more than 11 billion digits. [22]

Strict partition function

Definition and properties

A partition in which no part occurs more than once is called strict, or is said to be a partition into distinct parts. The function q(n) gives the number of these strict partitions of the given sum n. For example, q(3) = 2 because the partitions 3 and 1 + 2 are strict, while the third partition 1 + 1 + 1 of 3 has repeated parts. The number q(n) is also equal to the number of partitions of n in which only odd summands are permitted. [23]

Example values of q(n) and associated partitions
nq(n)Strict partitionsPartitions with only odd parts
01() empty partition() empty partition
1111
2121+1
321+2, 31+1+1, 3
421+3, 41+1+1+1, 1+3
532+3, 1+4, 51+1+1+1+1, 1+1+3, 5
641+2+3, 2+4, 1+5, 61+1+1+1+1+1, 1+1+1+3, 3+3, 1+5
751+2+4, 3+4, 2+5, 1+6, 71+1+1+1+1+1+1, 1+1+1+1+3, 1+3+3, 1+1+5, 7
861+3+4, 1+2+5, 3+5, 2+6, 1+7, 81+1+1+1+1+1+1+1, 1+1+1+1+1+3, 1+1+3+3, 1+1+1+5, 3+5, 1+7
982+3+4, 1+3+5, 4+5, 1+2+6, 3+6, 2+7, 1+8, 91+1+1+1+1+1+1+1+1, 1+1+1+1+1+1+3, 1+1+1+3+3, 3+3+3, 1+1+1+1+5, 1+3+5, 1+1+7, 9

Generating function

The generating function for the numbers q(n) is given by a simple infinite product: [24]

where the notation represents the Pochhammer symbol From this formula, one may easily obtain the first few terms (sequence A000009 in the OEIS ):

This series may also be written in terms of theta functions as

where

and

In comparison, the generating function of the regular partition numbers p(n) has this identity with respect to the theta function:

Identities about strict partition numbers

Following identity is valid for the Pochhammer products:

From this identity follows that formula:

Therefore those two formulas are valid for the synthesis of the number sequence p(n):

In the following, two examples are accurately executed:

Restricted partition function

More generally, it is possible to consider partitions restricted to only elements of a subset A of the natural numbers (for example a restriction on the maximum value of the parts), or with a restriction on the number of parts or the maximum difference between parts. Each particular restriction gives rise to an associated partition function with specific properties. Some common examples are given below.

Euler and Glaisher's theorem

Two important examples are the partitions restricted to only odd integer parts or only even integer parts, with the corresponding partition functions often denoted and .

A theorem from Euler shows that the number of strict partitions is equal to the number of partitions with only odd parts: for all n, . This is generalized as Glaisher's theorem, which states that the number of partitions with no more than d-1 repetitions of any part is equal to the number of partitions with no part divisible by d.

Gaussian binomial coefficient

If we denote the number of partitions of n in at most M parts, with each part smaller or equal to N, then the generating function of is the following Gaussian binomial coefficient:

Asymptotics

Some general results on the asymptotic properties of restricted partition functions are known. If pA(n) is the partition function of partitions restricted to only elements of a subset A of the natural numbers, then:

If A possesses positive natural density α then , with

and conversely if this asymptotic property holds for pA(n) then A has natural density α. [25] This result was stated, with a sketch of proof, by Erdős in 1942. [19] [26]

If A is a finite set, this analysis does not apply (the density of a finite set is zero). If A has k elements whose greatest common divisor is 1, then [27]

Related Research Articles

In number theory, an arithmetic, arithmetical, or number-theoretic function is generally 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". There is a larger class of number-theoretic functions that do not fit this definition, for example, the prime-counting functions. This article provides links to functions of both classes.

<span class="mw-page-title-main">Bessel function</span> Families of solutions to related differential equations

Bessel functions, first defined by the mathematician Daniel Bernoulli and then generalized by Friedrich Bessel, are canonical solutions y(x) of Bessel's differential equation

In mathematics, the prime number theorem (PNT) describes the asymptotic distribution of the prime numbers among the positive integers. It formalizes the intuitive idea that primes become less common as they become larger by precisely quantifying the rate at which this occurs. The theorem was proved independently by Jacques Hadamard and Charles Jean de la Vallée Poussin in 1896 using ideas introduced by Bernhard Riemann.

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">Integer partition</span> Decomposition of an integer as a sum of positive integers

In number theory and combinatorics, a partition of a non-negative integer n, also called an integer partition, is a way of writing n as a sum of positive integers. Two sums that differ only in the order of their summands are considered the same partition. For example, 4 can be partitioned in five distinct ways:

<span class="mw-page-title-main">Euler's constant</span> Constant value used in mathematics

Euler's constant is a mathematical constant, usually denoted by the lowercase Greek letter gamma, defined as the limiting difference between the harmonic series and the natural logarithm, denoted here by log:

<span class="mw-page-title-main">Stirling's approximation</span> Approximation for factorials

In mathematics, Stirling's approximation is an asymptotic approximation for factorials. It is a good approximation, leading to accurate results even for small values of . It is named after James Stirling, though a related but less precise result was first stated by Abraham de Moivre.

In mathematics, a generating function is a representation of an infinite sequence of numbers as the coefficients of a formal power series. Unlike an ordinary series, the formal power series is not required to converge: in fact, the generating function is not actually regarded as a function, and the "variable" remains an indeterminate. Generating functions were first introduced by Abraham de Moivre in 1730, in order to solve the general linear recurrence problem. One can generalize to formal power series in more than one indeterminate, to encode information about infinite multi-dimensional arrays of numbers.

<span class="mw-page-title-main">Divisor function</span> Arithmetic function related to the divisors of an integer

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. 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.

In mathematics, the Dedekind eta function, named after Richard Dedekind, is a modular form of weight 1/2 and is a function defined on the upper half-plane of complex numbers, where the imaginary part is positive. It also occurs in bosonic string theory.

<span class="mw-page-title-main">Theta function</span> Special functions of several complex variables

In mathematics, theta functions are special functions of several complex variables. They show up in many topics, including Abelian varieties, moduli spaces, quadratic forms, and solitons. As Grassmann algebras, they appear in quantum field theory.

In mathematics, the Jacobi elliptic functions are a set of basic elliptic functions. They are found in the description of the motion of a pendulum, as well as in the design of electronic elliptic filters. While trigonometric functions are defined with reference to a circle, the Jacobi elliptic functions are a generalization which refer to other conic sections, the ellipse in particular. The relation to trigonometric functions is contained in the notation, for example, by the matching notation for . The Jacobi elliptic functions are used more often in practical problems than the Weierstrass elliptic functions as they do not require notions of complex analysis to be defined and/or understood. They were introduced by Carl Gustav Jakob Jacobi. Carl Friedrich Gauss had already studied special Jacobi elliptic functions in 1797, the lemniscate elliptic functions in particular, but his work was published much later.

The Basel problem is a problem in mathematical analysis with relevance to number theory, concerning an infinite sum of inverse squares. It was first posed by Pietro Mengoli in 1650 and solved by Leonhard Euler in 1734, and read on 5 December 1735 in The Saint Petersburg Academy of Sciences. Since the problem had withstood the attacks of the leading mathematicians of the day, Euler's solution brought him immediate fame when he was twenty-eight. Euler generalised the problem considerably, and his ideas were taken up more than a century later by Bernhard Riemann in his seminal 1859 paper "On the Number of Primes Less Than a Given Magnitude", in which he defined his zeta function and proved its basic properties. The problem is named after Basel, hometown of Euler as well as of the Bernoulli family who unsuccessfully attacked the problem.

In mathematical analysis, asymptotic analysis, also known as asymptotics, is a method of describing limiting behavior.

In mathematics, specifically the theory of elliptic functions, the nome is a special function that belongs to the non-elementary functions. This function is of great importance in the description of the elliptic functions, especially in the description of the modular identity of the Jacobi theta function, the Hermite elliptic transcendents and the Weber modular functions, that are used for solving equations of higher degrees.

In mathematics, the Rogers–Ramanujan identities are two identities related to basic hypergeometric series and integer partitions. The identities were first discovered and proved by Leonard James Rogers, and were subsequently rediscovered by Srinivasa Ramanujan some time before 1913. Ramanujan had no proof, but rediscovered Rogers's paper in 1917, and they then published a joint new proof. Issai Schur independently rediscovered and proved the identities.

In mathematics, Ramanujan's congruences are some remarkable congruences for the partition function p(n). The mathematician Srinivasa Ramanujan discovered the congruences

<span class="mw-page-title-main">Anatoly Karatsuba</span> Russian mathematician (1937–2008)

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

<span class="mw-page-title-main">Rank of a partition</span>

In mathematics, particularly in the fields of number theory and combinatorics, the rank of a partition of a positive integer is a certain integer associated with the partition. In fact at least two different definitions of rank appear in the literature. The first definition, with which most of this article is concerned, is that the rank of a partition is the number obtained by subtracting the number of parts in the partition from the largest part in the partition. The concept was introduced by Freeman Dyson in a paper published in the journal Eureka. It was presented in the context of a study of certain congruence properties of the partition function discovered by the Indian mathematical genius Srinivasa Ramanujan. A different concept, sharing the same name, is used in combinatorics, where the rank is taken to be the size of the Durfee square of the partition.

References

  1. Sloane, N. J. A. (ed.), "SequenceA070177", The On-Line Encyclopedia of Integer Sequences , OEIS Foundation
  2. Caldwell, Chris K. (2017), The Top Twenty
  3. "PrimePage Primes: p(1289844341)", primes.utm.edu, retrieved 30 June 2022
  4. "The Top Twenty: Elliptic Curve Primality Proof", primes.utm.edu, retrieved 30 June 2022
  5. Abramowitz, Milton; Stegun, Irene (1964), Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables , United States Department of Commerce, National Bureau of Standards, p.  825, ISBN   0-486-61272-4
  6. Euler, Leonhard (1753), "De partitione numerorum", Novi Commentarii Academiae Scientiarum Petropolitanae (in Latin), 3: 125–169
  7. Ewell, John A. (2004), "Recurrences for the partition function and its relatives", The Rocky Mountain Journal of Mathematics, 34 (2): 619–627, doi: 10.1216/rmjm/1181069871 , JSTOR   44238988, MR   2072798
  8. Wilf, Herbert S. (1982), "What is an answer?", American Mathematical Monthly , 89 (5): 289–292, doi:10.2307/2321713, JSTOR   2321713, MR   0653502
  9. Al, Busra; Alkan, Mustafa (2018), "A note on relations among partitions", Proc. Mediterranean Int. Conf. Pure & Applied Math. and Related Areas (MICOPAM 2018), pp. 35–39
  10. 1 2 Hardy, G. H.; Wright, E. M. (2008) [1938], An Introduction to the Theory of Numbers (6th ed.), Oxford University Press, p. 380, ISBN   978-0-19-921986-5, MR   2445243, Zbl   1159.11001
  11. Berndt, Bruce C.; Ono, Ken (1999), "Ramanujan's unpublished manuscript on the partition and tau functions with proofs and commentary" (PDF), The Andrews Festschrift (Maratea, 1998), Séminaire Lotharingien de Combinatoire, vol. 42, Art. B42c, 63, MR   1701582
  12. 1 2 Ono, Ken (2004), The web of modularity: arithmetic of the coefficients of modular forms and -series, CBMS Regional Conference Series in Mathematics, vol. 102, Providence, Rhode Island: American Mathematical Society, p. 87, ISBN   0-8218-3368-5, Zbl   1119.11026
  13. Ahlgren, Scott; Boylan, Matthew (2003), "Arithmetic properties of the partition function" (PDF), Inventiones Mathematicae , 153 (3): 487–502, Bibcode:2003InMat.153..487A, doi:10.1007/s00222-003-0295-6, MR   2000466, S2CID   123104639
  14. Ono, Ken (2000), "Distribution of the partition function modulo ", Annals of Mathematics , 151 (1): 293–307, arXiv: math/0008140 , Bibcode:2000math......8140O, doi:10.2307/121118, JSTOR   121118, MR   1745012, S2CID   119750203, Zbl   0984.11050
  15. Ahlgren, Scott; Ono, Ken (2001), "Congruence properties for the partition function" (PDF), Proceedings of the National Academy of Sciences , 98 (23): 12882–12884, Bibcode:2001PNAS...9812882A, doi: 10.1073/pnas.191488598 , MR   1862931, PMC   60793 , PMID   11606715
  16. 1 2 Hardy, G. H.; Ramanujan, S. (1918), "Asymptotic formulae in combinatory analysis", Proceedings of the London Mathematical Society , Second Series, 17 (75–115). Reprinted in Collected papers of Srinivasa Ramanujan, Amer. Math. Soc. (2000), pp. 276–309.
  17. Andrews, George E. (1976), The Theory of Partitions, Cambridge University Press, p. 69, ISBN   0-521-63766-X, MR   0557013
  18. Rademacher, Hans (1937), "On the partition function ", Proceedings of the London Mathematical Society , Second Series, 43 (4): 241–254, doi:10.1112/plms/s2-43.4.241, MR   1575213
  19. 1 2 Erdős, P. (1942), "On an elementary proof of some asymptotic formulas in the theory of partitions" (PDF), Annals of Mathematics , Second Series, 43 (3): 437–450, doi:10.2307/1968802, JSTOR   1968802, MR   0006749, Zbl   0061.07905
  20. Nathanson, M. B. (2000), Elementary Methods in Number Theory, Graduate Texts in Mathematics, vol. 195, Springer-Verlag, p. 456, ISBN   0-387-98912-9, Zbl   0953.11002
  21. Johansson, Fredrik (2012), "Efficient implementation of the Hardy–Ramanujan–Rademacher formula", LMS Journal of Computation and Mathematics, 15: 341–59, arXiv: 1205.5991 , doi:10.1112/S1461157012001088, MR   2988821, S2CID   16580723
  22. Johansson, Fredrik (March 2, 2014), New partition function record: p(1020) computed
  23. Stanley, Richard P. (1997). Enumerative Combinatorics 1. Cambridge Studies in Advanced Mathematics. Vol. 49. Cambridge University Press. Proposition 1.8.5. ISBN   0-521-66351-2.
  24. Stanley, Richard P. (1997). Enumerative Combinatorics 1. Cambridge Studies in Advanced Mathematics. Vol. 49. Cambridge University Press. Proof of Proposition 1.8.5. ISBN   0-521-66351-2.
  25. Nathanson 2000, pp. 475–85.
  26. Nathanson 2000, p. 495.
  27. Nathanson 2000, pp. 458–64.