Integer partition

Last updated
Young diagrams associated to the partitions of the positive integers 1 through 8. They are arranged so that images under the reflection about the main diagonal of the square are conjugate partitions. Ferrer partitioning diagrams.svg
Young diagrams associated to the partitions of the positive integers 1 through 8. They are arranged so that images under the reflection about the main diagonal of the square are conjugate partitions.
Partitions of n with largest part k Partitions of n with biggest addend k.svg
Partitions of n with largest part k

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. (If order matters, the sum becomes a composition.) For example, 4 can be partitioned in five distinct ways:

Contents

4
3 + 1
2 + 2
2 + 1 + 1
1 + 1 + 1 + 1

The only partition of zero is the empty sum, having no parts.

The order-dependent composition 1 + 3 is the same partition as 3 + 1, and the two distinct compositions 1 + 2 + 1 and 1 + 1 + 2 represent the same partition as 2 + 1 + 1.

An individual summand in a partition is called a part. The number of partitions of n is given by the partition function p(n). So p(4) = 5. The notation λn means that λ is a partition of n.

Partitions can be graphically visualized with Young diagrams or Ferrers diagrams. They occur in a number of branches of mathematics and physics, including the study of symmetric polynomials and of the symmetric group and in group representation theory in general.

Examples

The seven partitions of 5 are

Some authors treat a partition as a decreasing sequence of summands, rather than an expression with plus signs. For example, the partition 2 + 2 + 1 might instead be written as the tuple (2, 2, 1) or in the even more compact form (22, 1) where the superscript indicates the number of repetitions of a part.

This multiplicity notation for a partition can be written alternatively as , where m1 is the number of 1's, m2 is the number of 2's, etc. (Components with mi = 0 may be omitted.) For example, in this notation, the partitions of 5 are written , and .

Diagrammatic representations of partitions

There are two common diagrammatic methods to represent partitions: as Ferrers diagrams, named after Norman Macleod Ferrers, and as Young diagrams, named after Alfred Young. Both have several possible conventions; here, we use English notation, with diagrams aligned in the upper-left corner.

Ferrers diagram

The partition 6 + 4 + 3 + 1 of the number 14 can be represented by the following diagram:

GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg

The 14 circles are lined up in 4 rows, each having the size of a part of the partition. The diagrams for the 5 partitions of the number 4 are shown below:

GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg
GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg
GrayDot.svg
GrayDot.svg
GrayDot.svg
GrayDot.svg
GrayDot.svg
GrayDot.svg
4=3 + 1=2 + 2=2 + 1 + 1=1 + 1 + 1 + 1

Young diagram

An alternative visual representation of an integer partition is its Young diagram (often also called a Ferrers diagram). Rather than representing a partition with dots, as in the Ferrers diagram, the Young diagram uses boxes or squares. Thus, the Young diagram for the partition 5 + 4 + 1 is

Young diagram for 541 partition.svg

while the Ferrers diagram for the same partition is

GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg

While this seemingly trivial variation does not appear worthy of separate mention, Young diagrams turn out to be extremely useful in the study of symmetric functions and group representation theory: filling the boxes of Young diagrams with numbers (or sometimes more complicated objects) obeying various rules leads to a family of objects called Young tableaux, and these tableaux have combinatorial and representation-theoretic significance. [1] As a type of shape made by adjacent squares joined together, Young diagrams are a special kind of polyomino. [2]

Partition function

Using Euler's method to find p(40): A ruler with plus and minus signs (grey box) is slid downwards, the relevant parts 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 parts 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 partition function counts the partitions of a non-negative integer . For instance, because the integer has the five partitions , , , , and . The values of this function for 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 ).

The generating function of is

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., [3] as follows:

as

In 1937, Hans Rademacher found a way to represent the partition function by the convergent series

where

and is the Dedekind sum.

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

Restricted partitions

In both combinatorics and number theory, families of partitions subject to various restrictions are often studied. [5] This section surveys a few such restrictions.

Conjugate and self-conjugate partitions

If we flip the diagram of the partition 6 + 4 + 3 + 1 along its main diagonal, we obtain another partition of 14:

RedDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg RedDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg RedDot.svg
GrayDot.svg
RedDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg RedDot.svg GrayDot.svg
GrayDot.svg GrayDot.svg RedDot.svg
GrayDot.svg GrayDot.svg
GrayDot.svg
GrayDot.svg
6 + 4 + 3 + 1=4 + 3 + 3 + 2 + 1 + 1

By turning the rows into columns, we obtain the partition 4 + 3 + 3 + 2 + 1 + 1 of the number 14. Such partitions are said to be conjugate of one another. [6] In the case of the number 4, partitions 4 and 1 + 1 + 1 + 1 are conjugate pairs, and partitions 3 + 1 and 2 + 1 + 1 are conjugate of each other. Of particular interest are partitions, such as 2 + 2, which have themselves as conjugate. Such partitions are said to be self-conjugate. [7]

Claim: The number of self-conjugate partitions is the same as the number of partitions with distinct odd parts.

Proof (outline): The crucial observation is that every odd part can be "folded" in the middle to form a self-conjugate diagram:

GrayDot.svg GrayDot.svg RedDot.svg GrayDot.svg GrayDot.svg    RedDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg
GrayDot.svg

One can then obtain a bijection between the set of partitions with distinct odd parts and the set of self-conjugate partitions, as illustrated by the following example:

GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
RedDot.svg RedDot.svg RedDot.svg RedDot.svg RedDot.svg RedDot.svg RedDot.svg
BlackDot.svg BlackDot.svg BlackDot.svg

GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg GrayDot.svg
GrayDot.svg RedDot.svg RedDot.svg RedDot.svg RedDot.svg
GrayDot.svg RedDot.svg BlackDot.svg BlackDot.svg
GrayDot.svg RedDot.svg BlackDot.svg
GrayDot.svg RedDot.svg
9 + 7 + 3=5 + 5 + 4 + 3 + 2
Dist. oddself-conjugate

Odd parts and distinct parts

Among the 22 partitions of the number 8, there are 6 that contain only odd parts:

Alternatively, we could count partitions in which no number occurs more than once. Such a partition is called a partition with distinct parts. If we count the partitions of 8 with distinct parts, we also obtain 6:

This is a general property. For each positive number, the number of partitions with odd parts equals the number of partitions with distinct parts, denoted by q(n). [8] [9] This result was proved by Leonhard Euler in 1748 [10] and later was generalized as Glaisher's theorem.

For every type of restricted partition there is a corresponding function for the number of partitions satisfying the given restriction. An important example is q(n) (partitions into distinct parts). The first few values of q(n) are (starting with q(0)=1):

1, 1, 1, 2, 2, 3, 4, 5, 6, 8, 10, ... (sequence A000009 in the OEIS ).

The generating function for q(n) is given by [11]

The pentagonal number theorem gives a recurrence for q: [12]

q(k) = ak + q(k 1) + q(k 2) q(k 5) q(k 7) + q(k 12) + q(k 15) q(k 22) ...

where ak is (1)m if k = 3m2m for some integer m and is 0 otherwise.

Restricted part size or number of parts

By taking conjugates, the number pk(n) of partitions of n into exactly k parts is equal to the number of partitions of n in which the largest part has size k. The function pk(n) satisfies the recurrence

pk(n) = pk(nk) + pk−1(n 1)

with initial values p0(0) = 1 and pk(n) = 0 if n 0 or k 0 and n and k are not both zero. [13]

One recovers the function p(n) by

One possible generating function for such partitions, taking k fixed and n variable, is

More generally, if T is a set of positive integers then the number of partitions of n, all of whose parts belong to T, has generating function

This can be used to solve change-making problems (where the set T specifies the available coins). As two particular cases, one has that the number of partitions of n in which all parts are 1 or 2 (or, equivalently, the number of partitions of n into 1 or 2 parts) is

and the number of partitions of n in which all parts are 1, 2 or 3 (or, equivalently, the number of partitions of n into at most three parts) is the nearest integer to (n + 3)2 / 12. [14]

Partitions in a rectangle and Gaussian binomial coefficients

One may also simultaneously limit the number and size of the parts. Let p(N, M; n) denote the number of partitions of n with at most M parts, each of size at most N. Equivalently, these are the partitions whose Young diagram fits inside an M × N rectangle. There is a recurrence relation

obtained by observing that counts the partitions of n into exactly M parts of size at most N, and subtracting 1 from each part of such a partition yields a partition of nM into at most M parts. [15]

The Gaussian binomial coefficient is defined as:

The Gaussian binomial coefficient is related to the generating function of p(N, M; n) by the equality

Rank and Durfee square

The rank of a partition is the largest number k such that the partition contains at least k parts of size at least k. For example, the partition 4 + 3 + 3 + 2 + 1 + 1 has rank 3 because it contains 3 parts that are ≥ 3, but does not contain 4 parts that are  4. In the Ferrers diagram or Young diagram of a partition of rank r, the r × r square of entries in the upper-left is known as the Durfee square:

RedDot.svg RedDot.svg RedDot.svg GrayDot.svg
RedDot.svg RedDot.svg RedDot.svg
RedDot.svg RedDot.svg RedDot.svg
GrayDot.svg GrayDot.svg
GrayDot.svg
GrayDot.svg

The Durfee square has applications within combinatorics in the proofs of various partition identities. [16] It also has some practical significance in the form of the h-index.

A different statistic is also sometimes called the rank of a partition (or Dyson rank), namely, the difference for a partition of k parts with largest part . This statistic (which is unrelated to the one described above) appears in the study of Ramanujan congruences.

Young's lattice

There is a natural partial order on partitions given by inclusion of Young diagrams. This partially ordered set is known as Young's lattice . The lattice was originally defined in the context of representation theory, where it is used to describe the irreducible representations of symmetric groups Sn for all n, together with their branching properties, in characteristic zero. It also has received significant study for its purely combinatorial properties; notably, it is the motivating example of a differential poset.

Random partitions

There is a deep theory of random partitions chosen according to the uniform probability distribution on the symmetric group via the Robinson–Schensted correspondence. In 1977, Logan and Shepp, as well as Vershik and Kerov, showed that the Young diagram of a typical large partition becomes asympototically close to the graph of a certain analytic function minimizing a certain functional. In 1988, Baik, Deift and Johansson extended these results to determine the distribution of the longest increasing subsequence of a random permutation in terms of the Tracy–Widom distribution. [17] Okounkov related these results to the combinatorics of Riemann surfaces and representation theory. [18] [19]

See also

Notes

  1. Andrews 1976, p. 199.
  2. Josuat-Vergès, Matthieu (2010), "Bijections between pattern-avoiding fillings of Young diagrams", Journal of Combinatorial Theory , Series A, 117 (8): 1218–1230, arXiv: 0801.4928 , doi:10.1016/j.jcta.2010.03.006, MR   2677686, S2CID   15392503 .
  3. Andrews 1976, p. 69.
  4. Hardy & Wright 2008, p. 380.
  5. Alder, Henry L. (1969). "Partition identities - from Euler to the present". American Mathematical Monthly. 76 (7): 733–746. doi:10.2307/2317861. JSTOR   2317861.
  6. Hardy & Wright 2008, p. 362.
  7. Hardy & Wright 2008, p. 368.
  8. Hardy & Wright 2008, p. 365.
  9. Notation follows Abramowitz & Stegun 1964 , p. 825
  10. Andrews, George E. (1971). Number Theory. Philadelphia: W. B. Saunders Company. pp. 149–50.
  11. Abramowitz & Stegun 1964 , p. 825, 24.2.2 eq. I(B)
  12. Abramowitz & Stegun 1964 , p. 826, 24.2.2 eq. II(A)
  13. Richard Stanley, Enumerative Combinatorics, volume 1, second edition. Cambridge University Press, 2012. Chapter 1, section 1.7.
  14. Hardy, G.H. (1920). Some Famous Problems of the Theory of Numbers. Clarendon Press.
  15. Andrews 1976, pp. 33–34.
  16. see, e.g., Stanley 1999 , p. 58
  17. Romik, Dan (2015). The surprising mathematics of longest increasing subsequences. Institute of Mathematical Statistics Textbooks. New York: Cambridge University Press. ISBN   978-1-107-42882-9.
  18. Okounkov, Andrei (2000). "Random matrices and random permutations". International Mathematics Research Notices. 2000 (20): 1043. doi: 10.1155/S1073792800000532 . S2CID   14308256.
  19. Okounkov, A. (2001-04-01). "Infinite wedge and random partitions". Selecta Mathematica. 7 (1): 57–81. arXiv: math/9907127 . doi:10.1007/PL00001398. ISSN   1420-9020. S2CID   119176413.

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.

In complex analysis, an entire function, also called an integral function, is a complex-valued function that is holomorphic on the whole complex plane. Typical examples of entire functions are polynomials and the exponential function, and any finite sums, products and compositions of these, such as the trigonometric functions sine and cosine and their hyperbolic counterparts sinh and cosh, as well as derivatives and integrals of entire functions such as the error function. If an entire function has a root at , then , taking the limit value at , is an entire function. On the other hand, the natural logarithm, the reciprocal function, and the square root are all not entire functions, nor can they be continued analytically to an entire function.

The Möbius function μ(n) is a multiplicative function in number theory introduced by the German mathematician August Ferdinand Möbius (also transliterated Moebius) in 1832. It is ubiquitous in elementary and analytic number theory and most often appears as part of its namesake the Möbius inversion formula. Following work of Gian-Carlo Rota in the 1960s, generalizations of the Möbius function were introduced into combinatorics, and are similarly denoted μ(x).

<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">Square-free integer</span> Number without repeated prime factors

In mathematics, a square-free integer (or squarefree integer) is an integer which is divisible by no square number other than 1. That is, its prime factorization has exactly one factor for each prime that appears in it. For example, 10 = 2 ⋅ 5 is square-free, but 18 = 2 ⋅ 3 ⋅ 3 is not, because 18 is divisible by 9 = 32. The smallest positive square-free numbers are

<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:

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.

In combinatorial mathematics, the Bell numbers count the possible partitions of a set. These numbers have been studied by mathematicians since the 19th century, and their roots go back to medieval Japan. In an example of Stigler's law of eponymy, they are named after Eric Temple Bell, who wrote about them in the 1930s.

In mathematics, the falling factorial is defined as the polynomial

<span class="mw-page-title-main">Partition function (number theory)</span> The number of partitions of an integer

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.

In mathematics, Euler's pentagonal number theorem relates the product and series representations of the Euler function. It states that

Faà di Bruno's formula is an identity in mathematics generalizing the chain rule to higher derivatives. It is named after Francesco Faà di Bruno, although he was not the first to state or prove the formula. In 1800, more than 50 years before Faà di Bruno, the French mathematician Louis François Antoine Arbogast had stated the formula in a calculus textbook, which is considered to be the first published reference on the subject.

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">Euler function</span> Mathematical function

In mathematics, the Euler function is given by

<span class="mw-page-title-main">Plane partition</span>

In mathematics and especially in combinatorics, a plane partition is a two-dimensional array of nonnegative integers that is nonincreasing in both indices. This means that

In the mathematical field of combinatorics, the q-Pochhammer symbol, also called the q-shifted factorial, is the product

In discrete mathematics, dominance order is a partial order on the set of partitions of a positive integer n that plays an important role in algebraic combinatorics and representation theory, especially in the context of symmetric functions and representation theory of the symmetric group.

The spt function is a function in number theory that counts the sum of the number of smallest parts in each integer partition of a positive integer. It is related to the partition function.

<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 an integer partition is a certain number 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.

In mathematics, solid partitions are natural generalizations of integer partitions and plane partitions defined by Percy Alexander MacMahon. A solid partition of is a three-dimensional array of non-negative integers such that

References