Acessibilidade / Reportar erro

Proving two partition identities

Abstracts

In this paper we give combinatorial proofs for two partition identities. The first one solves a recent open question formulated by G. E. Andrews.

Partition Identities; Combinatorics; Ferrers diagram


Neste artigo fornecemos provas combinatórias para duas identidades de partições. A primeira resolve uma questão recentemente formulada por G. E. Andrews.

Identidades de partições; Combinatória; Diagrama de Ferrers


Proving two partition identities

R. da SilvaI, 1 1 robson.dasilva@ufabc.edu.br ; J.C. FilhoII, 2 2 jair@unifei.edu.br ; J.P.O. SantosIII, 3 3 josepli@ime.unicamp.br

ICentro de Matemática, Computação e Cognição, UFABC, 09210-170 Santo André, SP, Brasil

IIDepartamento de Matemática e Computação, ICE - UNIFEI, 37500-903 Itajubá, MG, Brasil

IIIDepartamento de Matemática Aplicada, IMECC - UNICAMP, 13083-859 Campinas, SP, Brasil

ABSTRACT

In this paper we give combinatorial proofs for two partition identities. The first one solves a recent open question formulated by G. E. Andrews.

Keywords: Partition Identities, Combinatorics, Ferrers diagram.

RESUMO

Neste artigo fornecemos provas combinatórias para duas identidades de partições. A primeira resolve uma questão recentemente formulada por G. E. Andrews.

Palavras-chave: Identidades de partições, Combinatória, Diagrama de Ferrers.

1. Introduction

In [2] Andrews considers a variety of parity questions related to partition identities. We mention two of them just to illustrate results of this nature. The first one is a well known identity given by Euler.

Euler's Partition Identity. The number of partitions of any positive integer n into distinct parts equals the number of partitions of n into odd parts.

In terms of generating functions: for |q| < 1,

The second one is due to B. Gordon ([6], [7]) and H. Göllnitz ([4], [5]) that independently introduced parity considerations as follows:

First Gollnitz-Gordon Identity. The number of partitions of n into distinct non-consecutive parts with no even parts differing by exactly 2 equals the number of partitions of n into parts ≡ 1, 4, or 7 (mod 8).

At the end of [2], he presents a list of problems. In Section 2 we have a few definitions and the description of the Problem 5. In Section 3 we present our solution to the problem. In the last section we have a bijective proof for an identity related to the Durfee square of partitions.

2. Solving an Andrews's Problem

It is important to mention that a solution for Problem 5 of Andrews's paper [2] was given in [8]. However, Yee's proof is based on some results related to generating functions for partitions and q-series. Our proof, although equivalent to Yee's one, deals with combinatorial aspects of the Ferrers diagram of a partition.

Before presenting the proof we recall a few definitions from [1] and [2].

Definition 2.1. A partition of a positive integer n is a colection of positive integers λ1, λ2, ..., λssuch that λ1> λ2> · · · > λs. Each λiis called a part of the partition. We denote a partition by λ1 + λ2 + · · · + λsor1, λ2, ..., λs).

Example 2.1.The five partitions of 4 are: 4, 3 + 1, 2 + 2, 2 + 1 + 1, 1 + 1 + 1 + 1.

Definition 2.2. A Ferrers diagram of a partition λ1 + λ2 +· · ·+ λs is an array of dots left justified having λ1dots in the first row, λ2dots in the second row and so on.

Definition 2.3. Given a partition, the largest possible square of dots, starting in the upper left-hand corner, contained in its Ferrers diagram is called the Durfee square of this partition.

Example 2.2.Below we have the Ferrers diagram of the partition 8 + 7 + 4 + 3 with the Durfee square indicated.

Definition 2.4.Let λ = λ1 + λ2 + · · · + λjbe a partition, where λ1> λ2> · · · > λj . The lower odd parity index of λ, Ilo(λ), is defined as the maximum length of nondecreasing subsequences of 1, λ2, ..., λj} whose terms alternate in parity starting with an odd λi.

Example 2.3.

Ilo(11 + 11 + 7 + 6 + 6 + 3 + 2 + 1 + 1) = 5

Ilo(14 + 11 + 8 + 6 + 6 + 3 + 3 + 2) = 4

In [2] Andrews asks for a combinatorial proof for

where po(r, m, N) is the number of partitions of N into m distinct parts with lower odd parity index equal to r.

Consider, for example, the tables below. In the first columns we have partitions into distinct parts. The lower odd parity indices are shown in the second columns.

From these tables it is easy to see that and

In order to present a combinatorial proof of (2.2), we identify each partition with its Ferrers diagram, the standard graphical representation of a partition.

The combinatorial proof we are going to show is based on the one-to-one correspondence found by F. Franklin in 1881 to prove Legendre's combinatorial version of Euler's pentagonal number theorem (see [1]).

3. The Combinatorial Proof

Let De(N) (resp. Do(N)) be the number of partitions of N into distinct parts having even (resp. odd) lower odd parity index. Then, we can rephrase (2.2) in the following theorem.

Theorem 3.1.

Proof. We shall establish a one-to-one correspondence between the partitions enumerated by De(N) and those enumerated by Do(N). This correspondence will not work for some partitions of the numbers of the form N = n(3n+1)/2 and N = n(3n+5)/2 + 1.

Each partition λ = λ1 + λ2 + · · ·+ λs of N into distinct parts has a smallest part λs and the largest part λ1 is the first element of a decreasing sequence of l consecutive integers that are parts of λ. For example, below we have the Ferrers diagrams of the partitions 8+7+4+3 and 8+7+6+3+2 with their parameter λs and l.

In order to establish the one-to-one correspondence we consider the following two cases.

Case 1: λs > l. In this case, we remove 1 from the parts λ1, λ2, ..., λl and form a new smallest part of size l. For example,

8+7+4+3 7+6+4+3+2,

or, in terms of the graphical representation,

Case 2: λs< l. In this case, we remove the smallest part and add 1 to the first λs of the l largest parts. For example,

8+7+6+3+2 9+8+6+3,

that is,

Note that exactly one case is applicable to a given partition into distinct parts. Then, it seems that the mapping establishes a one-to-one correspondence. This is true except for certain partitions like 5+4+3.

The procedure described in Case 1 is not applicable to those partitions having l parts and λs = l+1, in which case the number being partitioned is

While the procedure described in Case 2 is not applicable to those partitions having l parts and λs = l, in which case the number being partitioned is

where m = l - 1. For example, Case 1 and Case 2 are not applicable to

Note that partitions of the form λ = (2l - 1) + · · ·+ (l + 1) + l have an odd lower odd parity index: if l is odd, then Ilo(λ) = l; if l is even, then Ilo(λ) = l - 1. We also note that partitions of the form λ = 2l + · · ·+ (l + 2) + (l + 1) have an even lower odd parity index: if l is odd, then Ilo(λ) = l - 1; if l is even, then Ilo(λ) = l.

In order to finish the proof we shall show that the correspondence described above changes the parity of the index Ilo(λ), when λ = λ1 + λ2 + · · ·+ λs is neither a partition into l parts with λs = l nor a partition into l parts with λs = l + 1. Again, we consider the two cases above.

We will call m the image partition obtained from λ = λ1 + λ2 + · · ·+ λl + λl+1 + · · ·+ λs by the correspondence above, i.e.,

We shall show that Ilo(λ) and Ilo(µ) have different parities.

Case 1: λs > l. We split this case into two sub-cases.

Sub-case 1.1: s = l. In this sub-case, λ = λ1 + λ2 + · · ·+ λl and µ = (λ1 - 1) + (λ2 -1) + · · ·+ (λl - 1) + l. Hence, if Ilo(λ) = l, then λl is odd and, consequently,

  • if l is odd, then Ilo(µ) = l + 1.

  • if l is even, then Ilo(µ) = l - 1.

If Ilo(λ) = l - 1, then λl is even and, consequently,

  • if l is odd, then Ilo(µ) = l.

  • if l is even, then Ilo(µ) = l.

For example

Sub-case 1.2: s > l. The parts λ1, λ2, ..., λl of λ alternate in parity. The same is true for the parts (λ1 - 1), (λ2 - 1), ..., (λl - 1) of µ. Then in order to determine the index Ilo(µ) in terms of Ilo(λ) we have to look at the parity of λl and λl+1 as well as the parity of l and λs. While the parity of λl is changed by the correspondence, the same does not happen with λl+1. For example,

As there are two choices for the parities of l, λl, λl+1, and λs, then we have to verify, in this sub-case, all the 24 possibilities. In the table below we summarise the results. In this table e means even and o means odd.

Case 2: λs< l. Again we consider two sub-cases.

Sub-case 2.1: λs < l. In this case the operation of removing the smallest part of

and adding 1 to each of its first λs largest parts produces

Since and have opposite parities, we have to consider the parities of λs, λs-1, and . For example, by this operation we can obtain

The table below shows the 23 possibilities and the values of Ilo(µ) in terms of Ilo(λ).

Sub-case 2.2: λs = l. In this sub-case we have to look at the parity of λs, λs-1, λl, and λl+1. For instance, if λs and λs-1 are odd and both λl and λl+1 are even, then Ilo(µ) = Ilo(λ) + 1. We summarize all possibilities in the next table.

As one can see from the tables above the parity of the index

Ilo is always changed by the one-to-one correspondence described above. Therefore the theorem is proved.

4. The Second Bijection

In this section we present a bijective proof for the partition identity below, which is related to the Durfee square (the largest possible square contained within the Ferrers diagram of a partition starting in the upper left-hand corner). This identity comes from two different interpretations for the Fibonacci Numbers (see [3]).

Theorem 4.2. The number of partitions into at most n parts without gaps 4 4 all positive integer less than or equal to the largest part are parts of the partition and having each part appearing at least twice is equal to the number of partitions where the largest part is the side of its Durfee square and the largest part plus the number of parts is less than or equal to n.

Proof. Let n be a non-negative integer. Define as the set of partitions into at most n parts without gaps and having each part appearing at least twice. We also define as being the set of partitions where the largest part is the side of its Durfee square and the largest part plus the number of parts is less than or equal to n.

Note that the empty partition is in both sets and , consequently we associate these two partitions. Now we describe the bijection between and for non-empty partitions. In order to do this we associate each partition with its Ferrers diagrams.

Given a non-empty partition λ ∈ , with largest part p, we obtain a partition µ, with Durfee square of side p, by the following procedure:

  • delete p dots from the first column of λ

  • move dots from the columns to the columns , respectively.

For example, consider the partition λ = 7 + 7 + 6 + 6 + 6 + 6 + 5 + 5 + 4 + 4 + 4 + 3 + 3 + 3 + 3 + 2 + 2 + 2 + 1 + 1 + 1. The partition m obtained by the above procedure is µ = 7 + 7 + 7 + 7 + 7 + 7 + 7 + 6 + 6 + 4 + 3 + 3 + 2 + 1 or, in terms of Ferrers diagrams,

It is important to note that the resulting diagram still represents a partition. This is ensured by the condition that each part appears at least twice.

Given a partition σ we obtain a partition δ just by inverting the above procedure. As an example, consider σ = 6 + 6 + 6 + 6 + 6 + 6 + + 3 + 1; the partition associated to σ is δ = 6 + 6 + 5 + 5 + 4 + 4 + 3 + 3 + 3 + 2 + 2 + 1 + 1 + 1:

Recebido em 13 Setembro 2011; Aceito em 20 Maio 2012.

  • [1] G.E. Andrews, "The Theory of Partitions", Cambridge University Press, 1984.
  • [2] G.E. Andrews, Parity in partition identities, The Ramanujan Journal, 23 (2010), 45-90.
  • [3] J.C. Filho, "Variações do Diagrama de Ferrers, Partições Planas e Funções Geradoras", Tese de Doutorado, IMECC, UNICAMP, Campinas, SP, 2006.
  • [4] H. Göllnitz, Einfache Partionen, Diplomarbeit W.S. 1960, Göttingen, 65 pp.
  • [5] H. Göllnitz, Partitionen mit Differenzenbedingungen, J. Reine Angew. Math, 225 (1967), 154-190.
  • [6] B. Gordon, Some Ramanujan-like continued fractions, Abstracts of Short Com- munications, Int. Congr. of Math., 29-30, Stockholm, 1962.
  • [7] B. Gordon, Some continued fractions of the Rogers-Ramanujan type, Duke Math. J., 31 (1965), 741-748.
  • [8] A.J. Yee, Ramanujan's partial theta series and parity in partitions, The Ramanujan Journal, 23 (2010), 215-225.
  • 1
  • 2
  • 3
  • 4
    all positive integer less than or equal to the largest part are parts of the partition
  • Publication Dates

    • Publication in this collection
      15 Oct 2012
    • Date of issue
      2012

    History

    • Received
      13 Sept 2011
    • Accepted
      20 May 2012
    Sociedade Brasileira de Matemática Aplicada e Computacional Rua Maestro João Seppe, nº. 900, 16º. andar - Sala 163 , 13561-120 São Carlos - SP, Tel. / Fax: (55 16) 3412-9752 - São Carlos - SP - Brazil
    E-mail: sbmac@sbmac.org.br