Separable integer partition classes

By George E. Andrews

Abstract

A classical method for partition generating function is developed into a tool with wide applications. New expansions of well-known theorems are derived, and new results for partitions with copies of are presented.

1. Introduction

The object of this paper is to systematize a process in the theory of integer partition that really dates back to Euler. It is epitomized in the partition-theoretic interpretation of three classical identities:

and

where

Equations Equation 1.1 and Equation 1.2 are Euler’s Reference 10, p. 19 while Equation 1.3 is the first of the celebrated Rogers-Ramanujan identities Reference 10, Ch. 7.

In section 2, we will analyze Equation 1.1, Equation 1.2, and Equation 1.3 from the point of view of separable integer partition classes.

A separable integer partition class (SIP), , with modulus , is a subset of all the integer partitions satisfying the following:

There is a subset ( is called the basis of ) such that for each integer , the number of elements of with parts is finite and every element of with parts is uniquely of the form

where is a partition in and is a partition into nonnegative parts, whose only restriction is that each part is divisible by . Furthermore, all partitions of the form Equation 1.5 are in .

As we will see in section 2, each of Equation 1.1, Equation 1.2, and Equation 1.3 can be developed from this point of view with modulus . However. this setting allows a similar examination of the first Göllnitz-Gordon identity Reference 16 in section 3:

More surprising is an analysis of Schur’s 1926 partition theorem Reference 24 in section 4. It is interesting to note that this analysis leads naturally to full proofs of both Equation 1.6 and Schur’s theorem.

In section 5, we examine a new application. In section 6, we extend the concept of SIP classes to partitions with copies of Reference 1. Section 7 extends SIP classes to overpartitions. In the final section, we discuss open questions and note that the idea first arose in dynamical systems concerning billiard trajectories Reference 13.

2. General theory and classical identities

The infinite series in each of Equation 1.1, Equation 1.2, and Equation 1.3 fit neatly into the SIP program with modulus . We begin with Equation 1.1. In this case, we let be the set of all integer partitions. Now for each , there is only one element of with parts, namely

and every element of with parts, say can be written

The generating function for the elements of with parts is therefore

Summing over all yields

the left-hand side of Equation 1.1.

Next we consider , the integer partitions that have distinct parts. Here for each there is again exactly one partition in with parts, namely

and every element of , say

can be written

Furthermore

constitutes an ordinary partition into nonnegative parts.

The generating function is therefore

Finally, if ( for Rogers and Ramanujan) is the set of integer partitions where the difference between parts is , the only element of with parts is

and, as with Equation 1.2, we obtain the generating function for the partitions in as

The above analysis of the series in Equation 1.1, Equation 1.2, and Equation 1.3 is far from new. Indeed these are proofs whose ideas date back to Euler and were discussed fully in centuries old number theory and combinatorics books (cf. Reference 23, Sec. 7, Ch. III, Reference 21, Ch. 19) for this way of looking at Equation 1.1, Equation 1.2, and Equation 1.3.

Perhaps the reason that this type of study has not gone farther is the fact that in each of the classical cases there was only one element of with parts.

As we will see in the remaining sections, there are many SIP classes with a number of elements of with parts. The real challenge in each instance will be to determine the generating function for . Obviously if we denote by the generating function for these elements of , then the generating function for all the partitions in is given by

where is the modulus associated with . In the three cases just considered, we hardly need to think about since in each case, there is only one element of with parts.

So how does one determine ? The idea is to refine one’s consideration of where is the generating function for those elements of with parts and largest part . Clearly

In practice we shall obtain recurrences for the . The recurrences will arise by noting the parts in the partitions in can’t get too far apart. Namely if is too far from the next part then can be subtracted from yielding another partition in and contradicting the uniqueness of the decomposition Equation 1.5.

The previous paragraph is vague because each individual SIP class provides different meaning for “too far from.”

Theorem 1 provides a large number of SIP classes and will facilitate the subsequent theorems in sections 3-5.

Theorem 1.

Let be a set of positive integers with (mod ) and be a set of nonnegative integers. Let be the set of all integer partitions

where , and for and each if (mod ), then , and if , .

Then is an SIP class with modulus , and consists of all those partitions

where if (mod ) then , and for , if (mod ), then

Proof.

We proceed by induction on the number of parts in the partition of .

Clearly, if , then we see that the single part partitions in are . Furthermore if is a one part partition in with (mod ), then

with .

Now suppose that our theorem holds for all partitions with fewer than parts. Let us consider an arbitrary partition in with parts

From the definition of , we see that

where is in and is a partition whose parts are multiples of and .

Now we know from the definition of that if (mod ), then

Now define to be the unique integer congruent to modulo in the interval

Clearly is in . It remains to show that exists so that

Since (mod ), we need only show that in Equation 2.2 is .

Now

thus

or

i.e.

Thus we have completed the induction step and the theorem follows.

Corollary 2.

As before, denotes the generating function for partitions in with exactly parts, and denotes the generating function for the partition in , where is an SIP class of modulus . Then

Proof.

It is clear from Theorem 1 that

is the generating function for all partitions in with exactly parts. Summing over all proves the corollary.

3. Göllnitz-Gordon

Identity Equation 1.6 is the perfect prototype to reveal how SIPs truly generalize the classical series that appear in Equation 1.1, Equation 1.2, and Equation 1.3.

First let us give the well-known partition-theoretic interpretation in Reference 16, Reference 17, Reference 19, Reference 20:

First Göllnitz-Gordon Theorem.

The number of partitions of in which the difference between parts is at least 2 and at least 4 between even parts equals the number of partitions of into parts congruent to 1, 4 or 7 modulo 8.

Let denote the set of all partitions in which the difference between parts is at least 2 and at least 4 between multiples of 2.

Lemma 3.

is an SIP class of modulus 2.

Proof.

is an instance of Theorem 1 with , , , .

Lemma 4.

Let be the generating function for the partitions in with parts and largest part equal to . Then

and for , ,

and

where

and

Proof.

We refer to Theorem 1 to determine recurrences for . Clearly Equation 3.1 is immediate.

By the conditions requiring closeness of parts as stated in Theorem 1, we see that for ,

This clearly implies

which establishes Equation 3.2.

As for Equation 3.3, we see by Equation 3.6 that

Now the standard recurrence for the -binomial coefficients (defined in Equation 3.4), namely Reference 10, p. 35

establishes that the right-hand side of Equation 3.3 also satisfies the recurrence Equation 3.8. In addition the right-hand side of Equation 3.3 fulfills Equation 3.1 when . Thus Equation 3.3 follows by a straightforward mathematical induction on .

Theorem 5.

First Göllnitz-Gordon identity

Remark 6.

This celebrated theorem and its history are presented in Reference 3 and Reference 10, sec. 7.4. Indeed the proof of Equation 3.11 given below follows essentially that given in Reference 6, pp. 40-41. We include it here to reveal that it is naturally suggested by the work here.

Proof.

Let us first treat Equation 3.10. By Corollary 2,

Now sum the -sum using the -binomial theorem Reference 10, p. 36. Hence

and Equation 3.10 is proved.

Suppose in Equation 3.12, we summed on rather than . Thus

Therefore

Hence

as desired.

4. Schur’s 1926 theorem

Here is the theorem in question Reference 24.

Schur’s Theorem.

The number of partitions of in which the parts are equals the number of partitions of in which the parts differ by at least 3 and at least 4 if one of the parts in question is divisible by 3.

In light of the fact that

we see that the first class of partitions in Schur’s theorem may be replaced by partitions into distinct nonmultiples of 3.

Indeed, this revision of Schur may be refined as follows (an idea first effectively considered in Reference 2 and amplified in Reference 11):

Refinement of Schur’s Theorem.

The generating function for partitions in which there are parts and parts and the difference conditions in Schur’s original theorem hold is the coefficient of in

Let denote the class of all partitions satisfying the difference conditions in Schur’s theorem.

Theorem 7.

is an SIP class.

Proof.

is the instance of Theorem 1 with , and .

In the remainder of this section, we shall first determine an explicit formula for the generating functions associated with . Then we will apply Corollary 2 to prove the Refinement of Schur’s Theorem.

Theorem 8.

Let be the generating function for the partitions in (with marking parts and marking parts ). Then for and ,

and for ,

where

Proof.

We let

and

By the definition of , we see that

and for ,

Now Equation 4.4 follows directly by comparing Equation 4.9 with , and Equation 4.5 is a restatement of Equation 4.8.

To prove Equation 4.2 and Equation 4.3, we need only show that the coefficient of on both sides of Equation 4.9 is identical, when the are replaced by the corresponding right hand sides of Equation 4.2 and Equation 4.3.

We begin with Equation 4.9 when . To make clear what we are doing, we write the right hand side of Equation 4.2 as

and the right hand side of Equation 4.3 as

Subtracting these expressions into the right hand side of Equation 4.9 with replaced by , we have

the coefficient of in Equation 4.12 is

and this last expression simplifies to through three applications of one or the other of the standard -binomial recurrences Reference 10, p. 35 reiterated here:

and

Thus Equation 4.9 is established for .

Finally we consider Equation 4.9 with in which case the assertion becomes

Cancelling the from both sides of Equation 4.16, we see that we need to establish that the coefficients of are identical on each side.

On the right hand side the coefficient is

by Equation 4.14.

Thus we have fulfilled the defining recurrence and initial conditions to establish the formulas Equation 4.2 and Equation 4.3. This concludes the proof of Theorem 8.

We shall conclude this section by proving the Refinement of Schur’s Theorem combining Corollary 2 with Theorem 8.

For the sake of brevity, we write

and

We note that by Equation 4.4,

To make the final theorem of this section readable, we first prove Lemmas 9, 10, 11.

Lemma 9.
Proof.

The coefficient of in

is

which is the coefficient asserted in Equation 4.20.

Lemma 10.
Proof.

This is an instance of the -Chu-Vandermonde summation Reference 4, p. 37, eq. (3,(Equation 3.10), , and .

Lemma 11.
Proof.

The left-hand side may be transformed into -Pochhammer symbols as is done in detail in the proof of Theorem 3.3 from Reference 9, p. 36.

Theorem 12.
Remark 13.

This is the restatement of the refinement of Schur’s theorem. It will appear in the following proof that the are infinite sums, but this is only a convenience of notation. Thus

Proof.

First we note that

Hence the coefficient of on the right hand of Equation 4.23 is

To complete the proof we must evaluate the coefficient of on the left side of Equation 4.23.

Now

by Lemma 9, with

So to get the coefficient of , we need . Thus the coefficient of on the left side of Equation 4.23 is

Now the sum on turns out to be the sum on the left side of Equation 4.21. Hence by Lemma 10, the above sum reduces to

by Lemma 11 which is exactly the expression in Equation 4.24. Thus Theorem 12 is proved.

5. Glasgow Mod 8

H. Göllnitz Reference 16, Reference 17 provided four partition identities related to partitions whose parts are restricted to certain residue classes modulo 8. Two of these theorems were independently discovered by B. Gordon Reference 19, Reference 20 and have been given the name Göllnitz-Gordon, as mentioned previously.

Lesser known is the following theorem which first appeared in the Glasgow Mathematics Journal in 1967 Reference 4, p. 127:

Glasgow Mod 8 Theorem.

Let denote the number of partitions of into parts congruent to 0, 2, 3, 4, or . Let denote the number of partitions of in which all parts are and each odd part is at least 3 larger than any part not exceeding it. Then for ,

For example, enumerating

and enumerating

A natural bijective proof appears in Reference 5.

We have chosen to consider this theorem owing to the fact that it has never appeared as a direct consequence of a series-product identity. Indeed, the relevant identity turns out to be

We shall first prove that Equation 5.1 is valid. We shall then prove that the left side of Equation 5.1 is an instance of Corollary 2.

Theorem 14.

Equation Equation 5.1 is valid.

Proof.

For ,

This follows by mathematical induction. For ,

Generally,

which is the term of the left-hand side. The result then follows by induction.

Now let in Equation 5.2. The left side converges to the left side of Equation 5.1, and

Lemma 15.

Let denote the class of partitions related to . Then is an SIP class of modulus 2.

Proof.

This follows immediately from Theorem 1 with .

Lemma 16.

Let be the generating function for the partitions in with parts and largest part equal to . Then

and for ,

Proof.

First we see that Equation 5.3 is immediate by inspection. Next we note that the two part partitions in are . Thus

and inspection reveals that Equation 5.4, Equation 5.5, Equation 5.6, and Equation 5.7 are valid for .

Now as in the previous sections, we see that

All that remains is to show that the right hand sides of Equation 5.4, Equation 5.5, Equation 5.6, and Equation 5.7 satisfy the defining recurrence Equation 5.9. Each is proved using instances of Equation 4.14 or Equation 4.15. We shall do one case which is typical. When , equation Equation 5.9 asserts

Do the asserted formulas for the , _) satisfy the recurrence Equation 5.10? Yes, as is clear from the following:

and this is exactly the recurrence Equation 5.10.

Lemma 17.
Proof.

Each of these four assertions is an instance of the -binomial theorem Reference 10, p. 36 applied to the corresponding equation in Lemma 16. We prove Equation 5.10 as typical.

By Equation 5.4 and the -binomial theorem,

Theorem 18.
Proof.

Corollary 19.

The Glasgow Mod 8 Theorem is true.

Proof.

This follows by comparing Equation 5.1 with Equation 5.14 and invoking Corollary 2 with .

6. Partitions with copies of

The basic idea epitomized by Theorem 1 is actually applicable in a broader context. In this section we shall describe its application to partitions with copies of Reference 1.

This subject considers partitions taken from the set of ordered pairs of positive integers with the second entry not exceeding the first entry. A partition with copies of of the positive integer is a finite collection of elements of wherein the first elements of the ordered pairs sum to . For example, there are six partitions of with copies of :

As was noted in Reference 1, there is a bijection between partitions with copies of and plane partitions.

Most important for our current considerations is the weighted difference between two elements of . Namely, we define , the weighted difference of and , as follows:

The main point of Reference 1 was to prove the following two results.

Theorem 20 (Reference 1, p. 41).

The partitions of with copies of wherein each pair of parts has positive weighted difference are equinumerous with the ordinary partitions of into parts .

Theorem 21 (Reference 1, p. 41).

The partitions of with copies of wherein each pair of parts has nonnegative weighted difference are equinumerous with the ordinary partitions of into parts .

These two theorems are special cases of a general theorem proved in Reference 1, Th. 3, p. 42. The proofs relied on bijection between the partitions in question and the results which provide several Rogers-Ramanujan type theorems concerning partitions with specified hook differences.

Our object here is to reveal a completely different path to proof by using an adaptation of Theorem 1.

Let denote the generating function for partitions with copies of where the weighted difference between successive parts (written in lexicographic ascending order, i.e. if or and ) is exactly , there are exactly parts, and the smallest part is of the form .

Theorem 22.

As we will see, Theorems 20 and 21 follow from Theorem 22 plus Lemma 24 via two identities given in L. J. Slater’s compendium Reference 25, eqs (46) and (61):

and

In addition, the case is related to the Slater identity Reference 25, p. 160, eq. (81)

The right hand side of Equation 6.4 is easily seen to be the generating function for , the number of partitions in which multiples of 7 are not repeated, all other parts are and parts appear in two colors. This observation together with Equation 6.4 establishes the following result.

Theorem 23.

The number of partitions of with copies of wherein successive parts have weighted difference equals .

We shall not require the full generality of Theorem 1 for our application of the SIP idea to partitions with copies of . Indeed we only need something analogous to the three classical examples provided initially in section 2.

Lemma 24.

Let . Suppose is a partition with copies of with the parts written in ascending lexicographic order (i.e. if or and ). Assume that the weighted difference between successive parts is . Then if has parts

there is a unique ordinary partition with nonnegative parts in nondecreasing order

and a unique partition with copies of

where and the successive weighted differences are all equal to , and

Remark 25.

Note that the subscripts for the original are identical with the subscript set for .

Proof.

We begin by noting that the subscript tuple uniquely defines the as follows:

where is the number of parts of the partition. Note that the weighted difference between successive terms in this sequence is always .

Now we uniquely construct the as follows. We begin with :

and since the first subscript is , we know that must be , so . Next we define

Clearly is unique. Is ? Yes, because

Next we define

and again

This continues for all the parts of , and this concludes the proof of the lemma.

Thus we have established the analogous paradigm for -copies of partitions that we considered for ordinary partitions.

The next step is to consider the generating function for the partitions

where and all weighted differences between successive parts in ascending order equal . Call this generating function for such partitions where the number of parts is and the largest part is .

Lemma 26.
Proof.

The first three lines of Equation 6.5 are immediate because the smallest part must be of the form .

For the last line, we see that the part must have directly below it a part that produces a weighted difference of . Thus if the subscript is , the part must be

because

Thus summing over we obtain the fourth line of Equation 6.5.

Lemma 27.

For ,

All instances of apart from those listed in Equation 6.5, Equation 6.6, Equation 6.7, Equation 6.8, and Equation 6.9 are identically zero.

Proof.

Note that Equation 6.6, Equation 6.7, Equation 6.8, and Equation 6.9 are asserting 8 identities between the and a power of times a -binomial coefficient. We denote by that asserted expression that is to be shown equal to . For example, from Equation 6.6 we are asserting

It is easy to check directly that the two top lines of Equation 6.5 hold for . It is also a simple matter to verify that all of the instances of that should be identically zero are indeed that via the given recurrence.

The heart of the proof is to show that each of 8 instances for required by Equation 6.6, Equation 6.7, Equation 6.8, and Equation 6.9 actually fulfills the recurrence. Each one is very similar to the others, so we will do Equation 6.6 for . First the case , which asserts

This is true because if there are just two parts where the larger lexicographically is and the smaller is some with the requirement that

then . So , and , as required.

Next we must treat the recurrence step. We evaluate the last line in Equation 6.5 for the ’s,

The other seven recurrences and initial conditions are proved in exactly this way.

We are now in a position to prove Theorem 22.

Proof.

By the definition of we see that .

There are four cases to treat: even or odd and even or odd. The cases are entirely similar, so we consider only odd and odd.

as desired. The other three cases, as noted previously, are perfectly analogous to this case.

Corollary 28.

For , the generating function for partitions with copies of in which the weighted difference between parts is at least is given by

Proof.

By Lemma 24 and Theorem 22, we see that the generating function for all partitions with copies of and having exactly parts is given by

and summing over all , we obtain the result.

Proof of Theorem 20.

This follows directly from Corollary 28 with and the identity Equation 6.2.

Proof of Theorem 21.

This follows directly from Corollary 28 with and the identity Equation 6.3.

Proof of Theorem 22.

This follows directly from Corollary 28 with and the identity Equation 6.4.

In addition, we can now interpret a couple of Ramanujan’s mock theta functions with partitions with copies of .

Theorem 29.

The tenth order mock theta function Reference 12, p. 149, eq. (8.1.2)

is the generating function for partitions with copies of where the weighted difference between parts is , and the smallest part is of the form

Proof.

We note by Theorem 21 that

and is the generating function for partitions with copies of where the weighted difference between parts is and the smallest part is of the form . Summing over all , we obtain the result.

Theorem 30.

The third order mock theta function Reference 12, p. 5, eq. (2.1.3)

is the generating function for partitions with copies of where the weighted difference between parts is 0 and the smallest part is of the form .

Proof.

The argument here is exactly that of the proof of Theorem 29 with the only change being that

Corollary 31.

Let denote the number of ordinary partitions of in which the largest part is unique and every other part occurs exactly twice. Let denote the number of partitions of with copies of where the weighted difference between successive parts is 0 and the smallest part is of the form . Then

Remark 32.

We shall show that is the generating function for both and . As an example, consider . , the partitions in question being , the partitions in question being .

Proof.

N. Fine Reference 15, p 57 has observed that is the generating function for partitions into odd parts without gaps. The conjugates of these partitions are the partitions enumerated by .

The result now follows from Corollary 31 which shows that is also the generating function for .

7. Overpartitions

Overpartitions were introduced by Corteel and Lovejoy Reference 14 as the natural combinatorial object counted by the coefficients of

Namely these are the partitions of wherein each part size can have (or not) one summand overlined. Thus the 8 overpartitions of 3 are , and .

Among the most appealing theorems on overpartitions is Lovejoy’s extension of Schur’s theorem to overpartitions Reference 22.

Theorem 33.

The number of overpartitions of in which no part is divisible by 3 equals the number of overpartitions of wherein adjacent parts differ by at least 3 if the smaller is overlined or divisible by 3 and by at least 6 if the smaller is overlined and divisible by 3.

Now the generating function for overpartitions in which no part is divisible by 3 is:

and this function appears in the sixth identity in L. J. Slater’s list Reference 25, p. 152, eq. (6) corrected.

It is completely unclear how exactly the right-hand side of Equation 7.1 fits in with Lovejoy’s theorem. It turns out that Equation 7.1 naturally fits into overpartitions with copies of . Here the same principle as before applies where now only one instance of may be overlined in any partition. The generating function is:

Thus the 16 overpartitions of 3 using copies of are .

Theorem 34.

Let denote the number of overpartitions of whose parts are not divisible by 3. Let denote the number of overpartitions with copies of in which (i) the weighted differences between adjacent parts is (ii). If the weighted difference of two or more successive parts is zero, then only the smallest part in the sequence (ignoring the subscript) can be overlined. Then for ,

As an example, when , and the overpartitions in question are , and the overpartitions with copies of are .

Proof of Theorem 34.

This result relies heavily on the discoveries chronicled in section 6. First let us consider

The first term on the right in Equation 7.2 generates overpartitions with exactly positive parts. The second term generates overpartitions with exactly nonnegative parts (including exactly one zero).

This dissection of Equation 7.2 then leads directly to the desired conclusion. Namely

Now we recall from Theorem 22 with that

is the generating function for partitions with copies of having parts with smallest part of the form .

Now instead of attaching an ordinary partition to this basic partition (as is done in Corollary 28), we attach overpartitions as generated in Equation 7.2.

First note that these are overpartitions that are being attached. Consequently this means that if there are several identical parts being attached the result will be a sequence of parts with successive differences still 0 and with only the smallest part in the chain possibly being overlined.

Second, the first term (as noted after Equation 7.2) produces those partitions where the smallest summand is not of the form , and the second term accounts for those partitions where the smallest summand is of the form .

8. Partitions with copies of and even subscripts

It may at first appear rather artificial to restrict ourselves to only those with even . However, this restriction leads to a new interpretation of one of the more striking results in L. J. Slater’s compendium Reference 25, p. 161, eq. (86)

The right hand side of Equation 8.1 is clearly the generating function for , the number of partitions of into parts . On the other hand, noting that

we may use the argument used to produce Equation 1.3 to see that the left hand side produces partitions

into nondecreasing parts with (cf. Reference 19).

The reason that we renew our study of Equation 8.1 is that it fits perfectly into the theme of the last two sections.

Theorem 35.

Let denote the number of partitions of using copies of but (i) restricted to even subscripts, (ii) the weighted differences between successive parts is , and (iii) excluding adjacent pairs with and both odd and . Then for ,

As an example, where the relevant partitions are where the relevant partitions are . Note that is disallowed because with both 7 and 3 odd.

Proof of Theorem 35.

We rewrite the left side of Equation 8.1 as

Now by conjugation of the 2 modular representations of partitions without repeated odd parts, we see that

is the generating function for partitions with exactly nonnegative parts with no repeated odd parts. On the other hand,

is merely the dilation of from Theorem 22. Thus is the generating function for partitions into copies of where summands are of the form .

Now we proceed here exactly as before in Lemma 27 and Corollary 28. The attachment of the ordinary partitions generated by Equation 8.2 to the partitions with copies of as generated by Equation 8.3 yields partitions into copies of with parts subject to the requirement that the weighted difference between parts is nonnegative, and that two successive parts and cannot have with both and odd.

9. Conclusion

There are several points to be made in summary.

First, we have chosen a sampling of possible applications of this method to make clear its widespread utility. Thus there are many instances of Theorem 1 that have yet to be considered. We have treated only a few of these.

Second, there are other examples of SIP classes. Indeed, the inspiration for this paper arose from Reference 13. The SIP class in Reference 13 is the set of integer partitions in which the parts are distinct, the smallest is even, and there are no consecutive odd parts. This SIP class is not an instance of Theorem 1. Consequently, it is surely valuable to explore SIPs not included in Theorem 1.

Finally, there are other theorems that cry out for an analogous theory. For example, the mod 7 instance of the generalization of the Rogers-Ramanujan identities given in Reference 7 may be stated:

There are several interpretations of the left hand side of Equation 6.1 Reference 8, Reference 9, Reference 18; however none seems to lend itself to an SIP-style interpretation. If such an interpretation could be found, this would open many further possibilities.

Acknowledgment

I wish to thank the referee whose care and insights greatly increased the readability of this work.

Mathematical Fragments

Equations (1.1), (1.2)
Equation (1.3)
Equation (1.5)
Equation (1.6)
Theorem 1.

Let be a set of positive integers with (mod ) and be a set of nonnegative integers. Let be the set of all integer partitions

where , and for and each if (mod ), then , and if , .

Then is an SIP class with modulus , and consists of all those partitions

where if (mod ) then , and for , if (mod ), then

Equation (2.2)
Corollary 2.

As before, denotes the generating function for partitions in with exactly parts, and denotes the generating function for the partition in , where is an SIP class of modulus . Then

Lemma 4.

Let be the generating function for the partitions in with parts and largest part equal to . Then

and for , ,

and

where

and

Equation (3.6)
Equation (3.7)
Equation (3.8)
Theorem 5.

First Göllnitz-Gordon identity

Equation (3.12)
Theorem 8.

Let be the generating function for the partitions in (with marking parts and marking parts ). Then for and ,

and for ,

where

Equation (4.8)
Equation (4.9)
Equation (4.12)
Equation (4.14)
Equation (4.15)
Equation (4.16)
Lemma 9.
Lemma 10.
Lemma 11.
Theorem 12.
Equation (4.24)
Equation (5.1)
Equation (5.2)
Lemma 16.

Let be the generating function for the partitions in with parts and largest part equal to . Then

and for ,

Equation (5.9)
Equation (5.10)
Lemma 17.
Theorem 20 (Reference 1, p. 41).

The partitions of with copies of wherein each pair of parts has positive weighted difference are equinumerous with the ordinary partitions of into parts .

Theorem 21 (Reference 1, p. 41).

The partitions of with copies of wherein each pair of parts has nonnegative weighted difference are equinumerous with the ordinary partitions of into parts .

Theorem 22.
Equation (6.2)
Equation (6.3)
Equation (6.4)
Lemma 24.

Let . Suppose is a partition with copies of with the parts written in ascending lexicographic order (i.e. if or and ). Assume that the weighted difference between successive parts is . Then if has parts

there is a unique ordinary partition with nonnegative parts in nondecreasing order

and a unique partition with copies of

where and the successive weighted differences are all equal to , and

Lemma 26.
Lemma 27.

For ,

All instances of apart from those listed in Equation 6.5, 6.6, 6.7, 6.8, and 6.9 are identically zero.

Corollary 28.

For , the generating function for partitions with copies of in which the weighted difference between parts is at least is given by

Theorem 29.

The tenth order mock theta function Reference 12, p. 149, eq. (8.1.2)

is the generating function for partitions with copies of where the weighted difference between parts is , and the smallest part is of the form

Corollary 31.

Let denote the number of ordinary partitions of in which the largest part is unique and every other part occurs exactly twice. Let denote the number of partitions of with copies of where the weighted difference between successive parts is 0 and the smallest part is of the form . Then

Equation (7.1)
Theorem 34.

Let denote the number of overpartitions of whose parts are not divisible by 3. Let denote the number of overpartitions with copies of in which (i) the weighted differences between adjacent parts is (ii). If the weighted difference of two or more successive parts is zero, then only the smallest part in the sequence (ignoring the subscript) can be overlined. Then for ,

Equation (7.2)
Equation (8.1)
Theorem 35.

Let denote the number of partitions of using copies of but (i) restricted to even subscripts, (ii) the weighted differences between successive parts is , and (iii) excluding adjacent pairs with and both odd and . Then for ,

Equation (8.2)
Equation (8.3)

References

Reference [1]
A. K. Agarwal and George E. Andrews, Rogers-Ramanujan identities for partitions with copies of , J. Combin. Theory Ser. A 45 (1987), no. 1, 40–49, DOI 10.1016/0097-3165(87)90045-8. MR883892,
Show rawAMSref \bib{1}{article}{ author={Agarwal, A. K.}, author={Andrews, George E.}, title={Rogers-Ramanujan identities for partitions with ``$n$ copies of $n$''}, journal={J. Combin. Theory Ser. A}, volume={45}, date={1987}, number={1}, pages={40--49}, issn={0097-3165}, review={\MR {883892}}, doi={10.1016/0097-3165(87)90045-8}, }
Reference [2]
Krishnaswami Alladi and Basil Gordon, Schur’s partition theorem, companions, refinements and generalizations, Trans. Amer. Math. Soc. 347 (1995), no. 5, 1591–1608, DOI 10.2307/2154961. MR1297520,
Show rawAMSref \bib{2}{article}{ author={Alladi, Krishnaswami}, author={Gordon, Basil}, title={Schur's partition theorem, companions, refinements and generalizations}, journal={Trans. Amer. Math. Soc.}, volume={347}, date={1995}, number={5}, pages={1591--1608}, issn={0002-9947}, review={\MR {1297520}}, doi={10.2307/2154961}, }
Reference [3]
George E. Andrews, A generalization of the Göllnitz-Gordon partition theorems, Proc. Amer. Math. Soc. 18 (1967), 945–952, DOI 10.2307/2035143. MR219497,
Show rawAMSref \bib{4}{article}{ author={Andrews, George E.}, title={A generalization of the G\"{o}llnitz-Gordon partition theorems}, journal={Proc. Amer. Math. Soc.}, volume={18}, date={1967}, pages={945--952}, issn={0002-9939}, review={\MR {219497}}, doi={10.2307/2035143}, }
Reference [4]
George E. Andrews, On Schur’s second partition theorem, Glasgow Math. J. 8 (1967), 127–132, DOI 10.1017/S0017089500000197. MR220692,
Show rawAMSref \bib{3}{article}{ author={Andrews, George E.}, title={On Schur's second partition theorem}, journal={Glasgow Math. J.}, volume={8}, date={1967}, pages={127--132}, issn={0017-0895}, review={\MR {220692}}, doi={10.1017/S0017089500000197}, }
Reference [5]
George E. Andrews, Note on a partition theorem, Glasgow Math. J. 11 (1970), 108–109, DOI 10.1017/S0017089500000938. MR269623,
Show rawAMSref \bib{5}{article}{ author={Andrews, George E.}, title={Note on a partition theorem}, journal={Glasgow Math. J.}, volume={11}, date={1970}, pages={108--109}, issn={0017-0895}, review={\MR {269623}}, doi={10.1017/S0017089500000938}, }
Reference [6]
George E. Andrews, Partition identities, Advances in Math. 9 (1972), 10–51, DOI 10.1016/0001-8708(72)90028-X. MR306105,
Show rawAMSref \bib{6}{article}{ author={Andrews, George E.}, title={Partition identities}, journal={Advances in Math.}, volume={9}, date={1972}, pages={10--51}, issn={0001-8708}, review={\MR {306105}}, doi={10.1016/0001-8708(72)90028-X}, }
Reference [7]
George E. Andrews, An analytic generalization of the Rogers-Ramanujan identities for odd moduli, Proc. Nat. Acad. Sci. U.S.A. 71 (1974), 4082–4085, DOI 10.1073/pnas.71.10.4082. MR351985,
Show rawAMSref \bib{7}{article}{ author={Andrews, George E.}, title={An analytic generalization of the Rogers-Ramanujan identities for odd moduli}, journal={Proc. Nat. Acad. Sci. U.S.A.}, volume={71}, date={1974}, pages={4082--4085}, issn={0027-8424}, review={\MR {351985}}, doi={10.1073/pnas.71.10.4082}, }
Reference [8]
George E. Andrews, On the Alder polynomials and a new generalization of the Rogers-Ramanujan identities, Trans. Amer. Math. Soc. 204 (1975), 40–64, DOI 10.2307/1997348. MR364083,
Show rawAMSref \bib{8}{article}{ author={Andrews, George E.}, title={On the Alder polynomials and a new generalization of the Rogers-Ramanujan identities}, journal={Trans. Amer. Math. Soc.}, volume={204}, date={1975}, pages={40--64}, issn={0002-9947}, review={\MR {364083}}, doi={10.2307/1997348}, }
Reference [9]
George E. Andrews, Partitions and Durfee dissection, Amer. J. Math. 101 (1979), no. 3, 735–742, DOI 10.2307/2373804. MR533197,
Show rawAMSref \bib{10}{article}{ author={Andrews, George E.}, title={Partitions and Durfee dissection}, journal={Amer. J. Math.}, volume={101}, date={1979}, number={3}, pages={735--742}, issn={0002-9327}, review={\MR {533197}}, doi={10.2307/2373804}, }
Reference [10]
George E. Andrews, The theory of partitions, Cambridge Mathematical Library, Cambridge University Press, Cambridge, 1998. Reprint of the 1976 original. MR1634067,
Show rawAMSref \bib{9}{book}{ author={Andrews, George E.}, title={The theory of partitions}, series={Cambridge Mathematical Library}, note={Reprint of the 1976 original}, publisher={Cambridge University Press, Cambridge}, date={1998}, pages={xvi+255}, isbn={0-521-63766-X}, review={\MR {1634067}}, }
Reference [11]
George E. Andrews, A refinement of the Alladi-Schur theorem, Lattice path combinatorics and applications, Dev. Math., vol. 58, Springer, Cham, 2019, pp. 71–77. MR3930452,
Show rawAMSref \bib{11}{article}{ label={11}, author={Andrews, George E.}, title={A refinement of the Alladi-Schur theorem}, conference={ title={Lattice path combinatorics and applications}, }, book={ series={Dev. Math.}, volume={58}, publisher={Springer, Cham}, }, date={2019}, pages={71--77}, review={\MR {3930452}}, }
Reference [12]
George E. Andrews and Bruce C. Berndt, Ramanujan’s lost notebook. Part V, Springer, Cham, 2018, DOI 10.1007/978-3-319-77834-1. MR3838409,
Show rawAMSref \bib{13}{book}{ author={Andrews, George E.}, author={Berndt, Bruce C.}, title={Ramanujan's lost notebook. Part V}, publisher={Springer, Cham}, date={2018}, pages={xii+430}, isbn={978-3-319-77832-7}, isbn={978-3-319-77834-1}, review={\MR {3838409}}, doi={10.1007/978-3-319-77834-1}, }
Reference [13]
G. E. Andrews, V. Dragovich, and M. Radnovic, Combinatorics of periodic ellipsoidal billards, (to appear).
Reference [14]
Sylvie Corteel and Jeremy Lovejoy, Overpartitions, Trans. Amer. Math. Soc. 356 (2004), no. 4, 1623–1635, DOI 10.1090/S0002-9947-03-03328-2. MR2034322,
Show rawAMSref \bib{14}{article}{ author={Corteel, Sylvie}, author={Lovejoy, Jeremy}, title={Overpartitions}, journal={Trans. Amer. Math. Soc.}, volume={356}, date={2004}, number={4}, pages={1623--1635}, issn={0002-9947}, review={\MR {2034322}}, doi={10.1090/S0002-9947-03-03328-2}, }
Reference [15]
Nathan J. Fine, Basic hypergeometric series and applications, Mathematical Surveys and Monographs, vol. 27, American Mathematical Society, Providence, RI, 1988. With a foreword by George E. Andrews, DOI 10.1090/surv/027. MR956465,
Show rawAMSref \bib{15}{book}{ author={Fine, Nathan J.}, title={Basic hypergeometric series and applications}, series={Mathematical Surveys and Monographs}, volume={27}, note={With a foreword by George E. Andrews}, publisher={American Mathematical Society, Providence, RI}, date={1988}, pages={xvi+124}, isbn={0-8218-1524-5}, review={\MR {956465}}, doi={10.1090/surv/027}, }
Reference [16]
H. Göllnitz, Einfache partitionen, Diplomarbeit W.S., Göttingen, 1960.
Reference [17]
H. Göllnitz, Partitionen mit Differenzenbedingungen (German), J. Reine Angew. Math. 225 (1967), 154–190, DOI 10.1515/crll.1967.225.154. MR211973,
Show rawAMSref \bib{17}{article}{ author={G\"{o}llnitz, H.}, title={Partitionen mit Differenzenbedingungen}, language={German}, journal={J. Reine Angew. Math.}, volume={225}, date={1967}, pages={154--190}, issn={0075-4102}, review={\MR {211973}}, doi={10.1515/crll.1967.225.154}, }
Reference [18]
Basil Gordon, A combinatorial generalization of the Rogers-Ramanujan identities, Amer. J. Math. 83 (1961), 393–399, DOI 10.2307/2372962. MR123484,
Show rawAMSref \bib{18}{article}{ author={Gordon, Basil}, title={A combinatorial generalization of the Rogers-Ramanujan identities}, journal={Amer. J. Math.}, volume={83}, date={1961}, pages={393--399}, issn={0002-9327}, review={\MR {123484}}, doi={10.2307/2372962}, }
Reference [19]
B. Gordon, Some Ramanujan-like continued fractions, Abstract of Short Communications, Stockholm, 1962, pp. 29–30, Inter. Congress of Math.
Reference [20]
Basil Gordon, Some continued fractions of the Rogers-Ramanujan type, Duke Math. J. 32 (1965), 741–748. MR184001,
Show rawAMSref \bib{20}{article}{ author={Gordon, Basil}, title={Some continued fractions of the Rogers-Ramanujan type}, journal={Duke Math. J.}, volume={32}, date={1965}, pages={741--748}, issn={0012-7094}, review={\MR {184001}}, }
Reference [21]
G. H. Hardy and E. M. Wright, An introduction to the theory of numbers, 5th ed., The Clarendon Press, Oxford University Press, New York, 1979. MR568909,
Show rawAMSref \bib{21}{book}{ author={Hardy, G. H.}, author={Wright, E. M.}, title={An introduction to the theory of numbers}, edition={5}, publisher={The Clarendon Press, Oxford University Press, New York}, date={1979}, pages={xvi+426}, isbn={0-19-853170-2}, isbn={0-19-853171-0}, review={\MR {568909}}, }
Reference [22]
Jeremy Lovejoy, A theorem on seven-colored overpartitions and its applications, Int. J. Number Theory 1 (2005), no. 2, 215–224, DOI 10.1142/S1793042105000157. MR2173381,
Show rawAMSref \bib{22}{article}{ author={Lovejoy, Jeremy}, title={A theorem on seven-colored overpartitions and its applications}, journal={Int. J. Number Theory}, volume={1}, date={2005}, number={2}, pages={215--224}, issn={1793-0421}, review={\MR {2173381}}, doi={10.1142/S1793042105000157}, }
Reference [23]
Percy A. MacMahon, Combinatory analysis, Chelsea Publishing Co., New York, 1960. Two volumes (bound as one). MR0141605,
Show rawAMSref \bib{23}{book}{ author={MacMahon, Percy A.}, title={Combinatory analysis}, note={Two volumes (bound as one)}, publisher={Chelsea Publishing Co., New York}, date={1960}, pages={xix+302+xix+340}, review={\MR {0141605}}, }
Reference [24]
I. Schur. Zur additiven Zahlentheorie, Gesammelte Abhandlungen. Vol II, Springer-Verlag, Berlin-New York, 1973, Herausgegeben von Alfred Brauer und Hans Rohrbach.
Reference [25]
L. J. Slater, Further identities of the Rogers-Ramanujan type, Proc. London Math. Soc. (2) 54 (1952), 147–167, DOI 10.1112/plms/s2-54.2.147. MR49225,
Show rawAMSref \bib{25}{article}{ label={25}, author={Slater, L. J.}, title={Further identities of the Rogers-Ramanujan type}, journal={Proc. London Math. Soc. (2)}, volume={54}, date={1952}, pages={147--167}, issn={0024-6115}, review={\MR {49225}}, doi={10.1112/plms/s2-54.2.147}, }

Article Information

MSC 2020
Primary: 11P83 (Partitions; congruences and congruential restrictions)
Keywords
  • Partitions
  • separable integer partition classes (SIP)
  • Rogers-Ramanujan
Author Information
George E. Andrews
Department of Mathematics, The Pennsylvania State University, University Park, Pennsylvania 16802
gea1@psu.edu
MathSciNet
Journal Information
Transactions of the American Mathematical Society, Series B, Volume 9, Issue 21, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the author under Creative Commons Attribution 3.0 License (CC BY 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/87
  • MathSciNet Review: 4445718
  • Show rawAMSref \bib{4445718}{article}{ author={Andrews, George}, title={Separable integer partition classes}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={9}, number={21}, date={2022}, pages={619-647}, issn={2330-0000}, review={4445718}, doi={10.1090/btran/87}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.