Rademacher’s Formula for the Partition Function
Abstract.
For a positive integer , let be the number of ways to express as a sum of positive integers. In this note, we revisit the derivation of the Rademacher’s convergent series for in a pedagogical way, with all the details given. We also derive the leading asymptotic behavior of when approaches infinity. Some numerical results are tabled.
1. Introduction
Given a positive integer , let be the number of ways to express as a sum of positive integer. For example, when , the followings are the different ways to express 7 as a sum of positive integers.
There are 15 of them. Hence, .
Let . Then the generating function of the sequence is
(1.1) |
The calculation of using the generating function is inefficient. A more efficient method is to make use of the Euler pentagonal number theorem, which states that
where
are the pentagonal numbers, and
From this, one can deduce the following recursive formula to calculate :
(1.2) |
Notice that for all ,
This implies that
The sum on the right hand side of (1.2) only involves approximately terms. Hence, it is an effective formula to calculate . By using and up to , one can calculate all for . Selected values of for in this range are tabled in Table 4.
In 1918, Hardy and Ramanujan [HR18] showed that satisfies the asymptotic formula:
More precisely, if we let
(1.3) |
then
This was discovered independently by Uspensky [Usp20] in 1920.
In fact, in [HR18], Hardy and Ramanujan obtained the following aymptotic formula
(1.4) |
where is a positive constant, and is the leading term given by (1.3). The terms , , , have similar forms, but with a constant smaller than in the exponential term.
A drawback of the asymptotic formula (1.4) is that the infinite series
is divergent, as was proved by Lehmer [Leh37].
In 1937, when Rademacher prepared lecture notes on the work of Hardy and Ramanujan, he made an improvement on the asymptotic formula (1.4). He obtained the following remarkable formula.
Theorem 1.1.
[Rademacher’s Formula [Rad37]]
If , the partition function is represented by the convergent series
where is defined in Theorem 3.1.
In contrast to the Hardy-Ramanujan formula, Rademacher’s formula is a convergent series. In [Rad37], Rademacher also showed that if is of order , then the remainder after terms is of order .
The Rademacher’s formula is a manifestation of the achievement of the circle method of Hardy, Ramanujan and Littlewood. This method has been very successful in problems of additive number theory [Rad40, Rad43, Vau81].
Partition function and its generalizations have been under active studies [Gro58, Gro60a, Gro60b, Gro62, Gro63, Gro84, Hag62, Hag63, Hag64a, Hag64b, Hag63, Hag65a, Hag65b, Hag65c, Hag66, Hag70, Hag71a, Hag71b, Hua42, Ise59, Ise60, Liv45, Niv40, RS76, Rob76, Rob77, Sel89, Spe73, Sub72, VSS82, Van82], especially in recent years [AW95, BL13, BM20, Bri06, Cra22, DM13, IJT20, JS15, Joh12, KMT19, KY99, MLP12, O’S20, PW19, Sel89, Sil10b, Sil10a, Sil10c, SZ12, dAP00, Pri09, PW19, BTB11, Zag21]. Modern approach to proving Rademacher’s formula usually involves advanced mathematics such as modular forms or Poincar series [DM13, dAP00, Pri09, PW19].
The purpose if this note is to present the proof of the Rademacher’s formula of partition function with all the necessary details. We follow closely the approach in [Apo90].
2. Preliminaries
2.1. Dedekind Eta Function
The Dedekind eta function is introduced by Dedekind in 1877 and is defined in the upper half plane by the equation
(2.1) |
The generating function (1.1) for the sequence is related to by
(2.2) |
The eta function is closely related to the theory of modular forms [Apo90]. Its 24th power, , is a modular form of weight 12.
To prove Rademacher’s formula, one needs a transformation formula for the Dedekind eta function under a fractional linear transformation
defined by an element of the modular group .
Before stating the transformation formula, let us define the Dedekind sum.
Definition 2.1.
If is an integer, is a positive integer larger than 1, the Dedekind sum is defined as
(2.3) |
When , is defined as 0 for any integer .
Theorem 2.2 (The Dedekind’s Functional Equation).
Let be integers with and . Under the fractional linear transformation
the Dedekind eta function satisfies the transformation formula.
(2.4) |
A proof of this theorem is given in [Apo90] using a more general formula proved by Iseki [Ise57]. In [KT23], we use a simpler approach to derive that formula.
2.2. Modified Bessel Functions
Let be a complex number. The modified Bessel function is defined as
(2.5) |
has a contour integral representation given by
where is a positive number. To prove this, we need the Hankel’s formula.
Let be the contour which is a loop around the negative real axis. As shown in Figure 1, it consists of three parts , and , where and are the lower and upper edges of a cut in the -plane along the negative real-axis, and is a positively oriented circle of radius about the origin.

Lemma 2.3 (Hankel’s Formula).
When , we have
Proof.
The integral defines an analytic function for all . It is sufficient to prove the formula for with . Then the result follows from analytic continuation. When , the integral over goes to as . In this limit, we have
∎
Using this lemma, we can prove the contour integral formula for the modified Bessel function .
Proposition 2.4.
Let be a complex number and let be a positive number. If is real, then
(2.6) |
Proof.
The integrand
is an analytic function of on the domain . It decays exponentially fast in the region when . Using Cauchy residue theorem, we can take the limit , and change the contour to . Namely,
Now, using the Taylor series of the exponential function, we have
where we have used Lemma 2.3. It follows from (2.5) that
∎
For our applications, we are interested in the case where .
Proposition 2.5.
has an explicit formula given by
Proof.
2.3. Farey Series and Ford Circles
Let be a positive integer. The set of Farey fractions of order , denoted by , is the set of all reduced fractions in the closed interval whose denominator is not larger than , listed in increasing order of magnitude.
Example 2.6.
The sets , , are given by
Obviously, for , is a subset of . Moreover,
which is the number of positive integers less than or equal to that are relatively prime to .
To be more precise, one adds in fractions of the form with and into the set to obtain . To determine the precise places to insert these fractions, one need the following lemmas.
Lemma 2.7.
If , , and are positive integers such that
then
Proof.
We have
Since and are positive, we have
Hence,
This proves the assertion. ∎
Lemma 2.8.
If , , , , and are positive integers such that
(2.7) |
and , then
Proof.
From this, we can deduce the following theorem.
Theorem 2.9.
Let , , , be positive integers such that
and . If
then and are consecutive terms in .
Proof.
Since , and are less than or equal to . Hence, and are in .
By Lemma 2.8, no fractions with can satisfy
Since only contains reduced fractions with denominator at most and , this shows that and are consecutive terms in . ∎
Now we can give the precise algorithm to construct from . First notice that for the two terms
in , , and thus
Now to construct , we insert the term
in between them. It is easy to check that if , and , then
In general, assume that has been constructed and for any two consecutive terms and in , . To construct , we need to insert the fractions with and into . To do this, we look for consecutive pairs of reduced fractions and in for which and . Let . Since and , we have
Lemma 2.7 implies that
By induction hypothesis, . As we have shown, this implies that
It remains to show that we can find exactly pairs of consecutive terms and in such that .
For , there are exactly positive integers with and . In fact, we must also have . For each of these integers , is also relatively prime to and . Since , and must also be relatively prime. There exists and such that
Any pairs of integers with can be written as
for some integer . Let be the smallest one so that . Then . If , then . Since
we find that . Since , we have and thus . In other words, we have shown that for each of the positive integers so that and , there are unique integers and such that
and . This shows that we can find exactly consecutive pairs in for us to insert the fractions with and .
Given a reduced fraction , the Ford circle (Figure 2) is the circle in the complex plane with center at
and radius
Obviously, each Ford circle touches the -axis at the point .

Theorem 2.10.
Two distinct Ford circles and are either tangent to each other or disjoint. They are tangent if and only if . In particular, if and are two consecutive Farey fractions, then the two Ford circles and are tangent to each other.
Proof.

Let be the distance between the centers of and . Then
The sum of the radii of the two circles is
Notice that the two circles are disjoint if and only if , and the two circles are tangent to each other if and only if . A straightforward computation gives
The two circles are distinct, so . Therefore, . This shows that
and so . Moreover, if and only if , if and only if . ∎
Now we study the points of tangency of two Ford circles that are tangent to each other.
Theorem 2.11.
Let
be three consecutive Farey fractions. The points of tangency of with and are given by
(2.8a) | ||||
(2.8b) |
Proof.
Notice that the condition of tangency implies that

For each positive integer , we construct a path joining the points and in the complex plane as follows. Let
be the Farey fractions in listed in increasing order. Let
For each , let be the point of tangency between and ; and let be the point of tangency between and . Then for . and divide the circle into two arcs, the upper arc and the lower arc. Let be the upper arc. Finally, let be the arc of the circle that joins the point to that is in the right half plane , and let be the arc of the circle that joins to the point that is in the left half-plane . The path is defined as
(2.9) |
following the orientation from the point to the point .



Let us study the image of under the mapping .
Lemma 2.12.
Let
be the unit disc on the complex plane and let
be an infinite vertical strip on the upper half plane. The transformation
is a conformal mapping that maps the vertical strip in the -plane onto the punctured unit disc in the -plane. The mapping is one-to-one on the interior of . The image of under this mapping is a simple closed contour that enclosed the origin.
Proof.

The mapping
is clearly analytic, and so it is a conformal mapping. Let and . Then we find that
As , we find that
On the other hand, as varies from 0 to 1, varies from 0 to . Hence, we see that the mapping transforms onto , and it is one-to-one except on the two vertical lines and .
Let be the image of under the map . Since the initial and end points of are and , and they are mapped to the same point in the unit disc , and so is a closed curve. Except for the initial and end points, the rest of lies in the interior of . Since is a simple curve and the mapping is one-to-one in the interior of , is a simple closed curve. ∎
To derive the Rademacher’s formula, we are extracting the coefficient from the generating function . As a function of , is periodic modulo 1. Hence, we can extend the path periodically modulo 1.
The Farey fractions in are given by
where
Define
By (2.8), is an arc joining the point to
and is an arc joining the point
to . Translating by 1 and combine with , we obtain a single arc of the circle joining to .
Henceforth, we regard the path as a path consists of the union of arcs , , , corresponding to the fractions with .
(2.10) |
For , ; whereas is the union of and that we have just described. If we take , and as consecutive Farey fractions, is the upper arc of the circle that joins the point to given by (2.8).
Before ending this section, we study further properties of the arc in the Ford circle under the fractional linear transformation
Theorem 2.13.
The mapping
(2.11) |
transforms the Ford circle in the -plane onto a circle in the -plane. The center of is the point and the radius is . If are three consecutive Farey fractions in , is the point of tangency between and , is the point of tangency between and , the images of and under the mapping (2.11) are and given by
(2.12) |
The upper arc on joining to is mapped to the arc on joining to that does not touch the imaginary axis.

Proof.
The mapping (2.11) is a fractional linear transformation. So it maps a circle to a circle. It is easy to verify the assertions in the theorem. In particular, since the mapping maps the real axis to the imaginary axis, the image of the upper arc on joining to , is the arc on joining to that is away from the imaginary axis. ∎
Proposition 2.14.
Let be the circle on the complex plane with center at and radius . For any , and
Proof.
If is a point on ,
This implies that
and hence
On the other hand, can be written as
for some . It follows that
∎
Theorem 2.15.
Let be three consecutive Farey fractions in , and let and be the points given in (2.12) in Theorem 2.13. The moduli of these points are given by
(2.13) |
For any point on the chord joining and ,
Hence, the length the chord is not larger than . On the other hand, the real part of is bounded below by , and hence,
Proof.
Eq. (2.13) is easily verified. To prove the second assertion, notice that if , then is a Farey fraction in and it lies between and . This contradicts to and are consecutive in . Therefore, we must have . Similarly, . Hence,
Similarly,
If is on the chord joining and , then there is a real number in the interval such that
It follows that
By triangle inequality, the length of the chord is not larger than , and hence, it is not larger than .
Now, for the assertion about the real part, we use the fact that are all not larger than . Then
It follows that
∎
Finally, we have the following elementary proposition.
Proposition 2.16.
Let be the circle with center at and radius . Given a point on , let be the minor arc joining the point 0 to the point . Then , the arc-length of , is bounded above by For any on , .
Proof.

Any point on can be written as
(2.14) |
for some . Let
By symmetry of , we can assume that . Then
The arc-length of is
where is the central angle of . It is elementary to show that if ,
Therefore,
This proves the first assertion.
Now, if is a point on , then is given by (2.14) for some . If follows that
∎
3. Rademacher’s Convergent Series
In this section, we will prove the Rademacher’s Theorem 1.1, which we state again here.
Theorem 3.1.
[Rademacher’s Formula [Rad37]]
If , the partition function is represented by the convergent series
(3.1) |
where
and is the Dedekind sum defined by
(3.2) |
We break the proof into a few lemmas.
Lemma 3.2.
Let
(3.3) |
and let be any positively oriented simple closed curve in the unit disc surrounding the point . Given a positive integer integer ,
Proof.
The infinite product in (3.3) converges absolutely when . Hence, is an analytic function on the unit disc . Since the Taylor series of is given by
By Cauchy integral formula,
∎
As we have mentioned in the proof of Lemma 3.2, the function is analytic in the unit disc . By the product expansion, we see that it has singularities at the points , where and are positive integers. In the following theorem, we present a transformation formula for near a singularity point .
Theorem 3.3.
Let
(3.4) |
and let and be integers with and . Let be a positive integer not larger than such that . Finally, let be a complex number with . Define
Then
(3.5) |
where is given by (3.2).
When is small, is a point that is close to the point , whereas is a point that is close to the origin. Since , this theorem says that when is close to the singularity point ,
Proof.
Proof of Theorem 3.1.
Fixed a positive integer . We let be the image of the path defined by (2.10) under the map
By Lemma 3.2,
Making the change of variables , we have
Each is the upper arc of the circle with initial point and end point , where
For the integral over the arc , we make a change of variables
so that
Then
(3.6) |
Here is the arc of the circle with center at and radius , which do not touch the imaginary axis, and joining the points and given by
(3.7) |
Now we use the transformation formula given in Theorem 3.3,
(3.8) |
where is a positive integer not larger than and such that mod .
Let
(3.9) |
Then (3.6) and (3.8) imply that
where
As runs from to , runs through all pairs of positive integers with and . Therefore,
We estimate first. Given in , let be the chord joining the point to the point . Then is a closed curve that does not enclose the origin. Since the integrand
(3.10) |
is analytic in an open set containing the region enclosed by the closed curve , by residue theorem, we have
Hence,
For any , Theorem 2.15 says that
(3.11) |
and the arclength of , , is not larger than . If we can find an such that
(3.12) |
then
Let us now find an satisfying (3.12). By definition (3.10),
The estimate of is given by (3.11). Now,
For , Proposition 2.14 implies that . Hence,
On the other hand,
Therefore,
Next, we estimate
Using definition (3.9) and triangle inequality, we have
Therefore,
For a point , Theorem 2.15 says that
Using the fact that when , we find that
Since all the coefficients of are positive, we find that if . Hence,
It follows that
Hence,
where
is a fixed constant that only depends on . It is easy to see that
since the number of positive integer such that and is not larger than .
Hence, we have shown that
which is of order . It follows that
Here we write
to mean
To compute , we notice that when traverses the circle from to along the arc that is away from the imaginary axis, it traverses clockwise. We can write this arc as the whole circle (the minus sign is for the clockwise orientation) minus the two arcs and , where is the arc from 0 to which is clockwise, and is the arc 0 to which is anticlockwise. This gives,
For any , Proposition 2.14 and Proposition 2.16 imply that
Theorem 2.15 and Proposition 2.16 say that and are bounded above by
It follows that
where
This shows that
(3.13) |
Taking limit, we find that
where
In particular, . Making a change of variables
the circle is mapped to the line , with goes from to . Hence,
Making a change of variables
we find that
where
By (2.6), we find that
where
By Proposition 2.5,
Notice that (3.13) shows that the series indeed converges to . This completes the proof of the theorem.
∎
4. Leading Asymptotic of the Partition Function
In the following, we use the exact formula for we obtained in the previous section to find the leading asymptotic of as a function of when .
Theorem 4.1.
The sequence satisfies the following asymptotic formula when .
(4.1) |
Given a positive integer , let
Eq. (4.1) asserts that
To prove this, let
Then by (3.1),
To prove (4.1), it is sufficient to show that
(4.2) |
and
(4.3) |
By a straightforward computation, we find that
(4.4) |
where
We first prove (4.2).
Lemma 4.2.
Let be a positive integer. Then
Proof.
To prove (4.3), we need the following.
Lemma 4.3.
For ,
(4.5) |
Proof.
Let
By Cauchy mean value theorem, there exists such that
Now,
Thus,
By mean value theorem, there exists such that
Since is a strictly increasing function, . Thus,
The assertion follows. ∎
Proof of Theorem 4.1.
By (4.5) and (4.4), and using the fact that , we have
Using the fact that when , we have
Hence,
where
It follows that
It is easy to verify that
Hence,
This completes the proof of (4.1).
∎
Theorem 4.1 states that the ratio of to approaches 1 when approaches infinity. This does not mean when we use to approximate , the error in the approximation, defined as , is small. In fact, this error also tends to infinity when approaches infinity. What we assert is that the relative error
goes to 0. The percentage relative error is defined as
Table 1 shows some values of as well as the percentage relative errors in the approximation of by . Figure 11 depicts the percentage relative errors graphically.

10 | 42 |
---|---|
50 | 204226 |
100 | 190569292 |
200 | 3972999029388 |
500 | 2300165032574323995027 |
1000 | 24061467864032622473692149727991 |
2000 | 4720819175619413888601432406799959512200344166 |
3000 | 496025142797537184410324879054927095334462742231683423624 |
4000 | 1024150064776551375119256307915896842122498030313150910234889093895 |
5000 |
1698201688254421218519751016893064313617576830498292333222038246523
29144349 |
6000 |
4671727531970209092971024643973690643364629153270037033856605528925
072405349246129 |
7000 |
3285693080344061578628092563592416686195015157453224065969903215743
2236394 374450791229199 |
8000 |
7836026435156834949059314501336459971901076935298586433111860020941
7827764524450990388402844164 |
9000 |
7713363811780888490732079142740313496163979832207203426264771369460
5367979684296948790335590435626459 |
10000 |
3616725132563629398882047189095369549501603033931565042208186860588
7952568754066420592310556052906916435144 |
12000 |
1294107667757322067493842620367467386268131006205640080126511905905
0170600581269291250270699 01623662251809128853180610 |
15000 |
2626337936403790841371023191659066988029320559654372494065885879713
75120081791056718639088570913175942816125969709246029351672130266 |
Selected values of . |
percentage relative error | ||
---|---|---|
10 | ||
50 | ||
100 | ||
200 | ||
500 | ||
1000 | ||
2000 | ||
3000 | ||
4000 | ||
5000 | ||
6000 | ||
7000 | ||
8000 | ||
9000 | ||
10000 | ||
12000 | ||
References
- [Apo90] Tom M. Apostol, Modular functions and Dirichlet series in number theory, second ed., Graduate Texts in Mathematics, vol. 41, Springer-Verlag, New York, 1990. MR 1027834
- [AW95] Gert Almkvist and Herbert S. Wilf, On the coefficients in the Hardy-Ramanujan-Rademacher formula for , J. Number Theory 50 (1995), no. 2, 329–334. MR 1316828
- [BL13] Ben Barber and Imre Leader, Partition regularity with congruence conditions, J. Comb. 4 (2013), no. 3, 293–297. MR 3096137
- [BM20] Stella Brassesco and Arnaud Meyroneinc, An expansion for the number of partitions of an integer, Ramanujan J. 51 (2020), no. 3, 563–592. MR 4076172
- [Bri06] Kathrin Bringmann, Asymptotic formulas for some restricted partition functions, Ramanujan J. 12 (2006), no. 2, 257–266. MR 2286249
- [BTB11] Nesrine Benyahia Tani and Sadek Bouroubi, Enumeration of the partitions of an integer into parts of a specified number of different sizes and especially two sizes, J. Integer Seq. 14 (2011), no. 3, Article 11.3.6, 12. MR 2783952
- [Cra22] William Craig, On the number of parts in congruence classes for partitions into distinct parts, Res. Number Theory 8 (2022), no. 3, Paper No. 52, 24. MR 4456287
- [dAP00] Wladimir de Azevedo Pribitkin, Revisiting Rademacher’s formula for the partition function , Ramanujan J. 4 (2000), no. 4, 455–467. MR 1811909
- [DM13] Michael Dewar and M. Ram Murty, A derivation of the Hardy-Ramanujan formula from an arithmetic formula, Proc. Amer. Math. Soc. 141 (2013), no. 6, 1903–1911. MR 3034417
- [Gro58] Emil Grosswald, Some theorems concerning partitions, Trans. Amer. Math. Soc. 89 (1958), 113–128. MR 97371
- [Gro60a] E. Grosswald, Partitions into prime powers, Michigan Math. J. 7 (1960), 97–122. MR 121348
- [Gro60b] Emil Grosswald, Correction and addition to “Some theorems concerning partitions”, Trans. Amer. Math. Soc. 95 (1960), 190. MR 120212
- [Gro62] E. Grosswald, Results, new and old, in the theory of partitions, Rev. Un. Mat. Argentina 20 (1962), 40–57. MR 150119
- [Gro63] by same author, Elementary proofs in the theory of partitions, Math. Z. 81 (1963), 52–61. MR 148623
- [Gro84] Emil Grosswald, Partitions into squares, Enseign. Math. (2) 30 (1984), no. 3-4, 223–245. MR 767902
- [Hag62] Peter Hagis, Jr., A problem on partitions with a prime modulus , Trans. Amer. Math. Soc. 102 (1962), 30–62. MR 146166
- [Hag63] by same author, Partitions into odd summands, Amer. J. Math. 85 (1963), 213–222. MR 153651
- [Hag64a] by same author, On a class of partitions with distinct summands, Trans. Amer. Math. Soc. 112 (1964), 401–415. MR 166177
- [Hag64b] by same author, Partitions into odd and unequal parts, Amer. J. Math. 86 (1964), 317–324. MR 166178
- [Hag65a] by same author, A correction of some theorems on partitions, Trans. Amer. Math. Soc. 118 (1965), 550. MR 176970
- [Hag65b] by same author, On the partitions of an integer into distinct odd summands, Amer. J. Math. 87 (1965), 867–873. MR 188182
- [Hag65c] by same author, Partitions with odd summands—some comments and corrections, Amer. J. Math. 87 (1965), 218–220. MR 176969
- [Hag66] by same author, Some theorems concerning partitions into odd summands, Amer. J. Math. 88 (1966), 664–681. MR 200260
- [Hag70] by same author, A root of unity occurring in partition theory, Proc. Amer. Math. Soc. 26 (1970), 579–582. MR 265308
- [Hag71a] by same author, Partitions into unequal parts satisfying certain congruence conditions, J. Number Theory 3 (1971), 115–123. MR 268145
- [Hag71b] by same author, Partitions with a restriction on the multiplicity of the summands, Trans. Amer. Math. Soc. 155 (1971), 375–384. MR 272735
- [HR18] G. H. Hardy and S. Ramanujan, Asymptotic Formulaae in Combinatory Analysis, Proc. London Math. Soc. (2) 17 (1918), 75–115. MR 1575586
- [Hua42] Loo-keng Hua, On the number of partitions of a number into unequal parts, Trans. Amer. Math. Soc. 51 (1942), 194–201. MR 6195
- [IJT20] Jonas Iskander, Vanshika Jain, and Victoria Talvola, Exact formulae for the fractional partition functions, Res. Number Theory 6 (2020), no. 2, Paper No. 20, 17. MR 4092503
- [Ise57] Shô Iseki, The transformation formula for the Dedekind modular function and related functional equations, Duke Math. J. 24 (1957), 653–662. MR 91301
- [Ise59] by same author, A partition function with some congruence condition, Amer. J. Math. 81 (1959), 939–961. MR 108473
- [Ise60] by same author, On some partition functions, J. Math. Soc. Japan 12 (1960), 81–88. MR 113861
- [Joh12] Fredrik Johansson, Efficient implementation of the Hardy-Ramanujan-Rademacher formula, LMS J. Comput. Math. 15 (2012), 341–359. MR 2988821
- [JS15] Chris Jennings-Shaffer, Congruences for partition pairs with conditions, Q. J. Math. 66 (2015), no. 3, 837–860. MR 3396094
- [KMT19] Narissara Khaochim, Riad Masri, and Wei-Lun Tsai, An effective bound for the partition function, Res. Number Theory 5 (2019), no. 1, Paper No. 14, 25. MR 3903861
- [KT23] Z.Y. Kong and L.P. Teo, An elementary proof of the transformation formula for the Dedekind eta function, Preprint arXiv:2302.03280 (2023).
- [KY99] Dongsu Kim and Ae Ja Yee, A note on partitions into distinct parts and odd parts, Ramanujan J. 3 (1999), no. 2, 227–231. MR 1703261
- [Leh37] D. H. Lehmer, On the Hardy-Ramanujan Series for the Partition Function, J. London Math. Soc. 12 (1937), no. 3, 171–176. MR 1575064
- [Liv45] John Livingood, A partition function with the prime modulus , Amer. J. Math. 67 (1945), 194–208. MR 12101
- [MLP12] James Mc Laughlin and Scott Parsell, A Hardy-Ramanujan-Rademacher-type formula for -regular partitions, Ramanujan J. 28 (2012), no. 2, 253–271. MR 2925178
- [Niv40] Ivan Niven, On a certain partition function, Amer. J. Math. 62 (1940), 353–364. MR 1235
- [O’S20] Cormac O’Sullivan, Rademacher’s conjecture and expansions at roots of unity of products generating restricted partitions, J. Number Theory 216 (2020), 335–379. MR 4130086
- [Pri09] Wladimir de Azevedo Pribitkin, Simple upper bounds for partition functions, Ramanujan J. 18 (2009), no. 1, 113–119. MR 2471621
- [PW19] Wladimir de Azevedo Pribitkin and Brandon Williams, Short proof of Rademacher’s formula for partitions, Res. Number Theory 5 (2019), no. 2, Paper No. 17, 6. MR 3916267
- [Rad37] Hans Rademacher, On the Partition Function p(n), Proc. London Math. Soc. (2) 43 (1937), no. 4, 241–254. MR 1575213
- [Rad40] by same author, Fourier expansions of modular forms and problems of partition, Bull. Amer. Math. Soc. 46 (1940), 59–73. MR 842
- [Rad43] by same author, On the expansion of the partition function in a series, Ann. of Math. (2) 44 (1943), 416–422. MR 8618
- [Rob76] M. M. Robertson, Partitions with congruence conditions, Proc. Amer. Math. Soc. 57 (1976), no. 1, 45–49. MR 404184
- [Rob77] by same author, Partitions of large multipartites with congruence conditions. II, J. Number Theory 9 (1977), no. 2, 258–270. MR 444596
- [RS76] M. M. Robertson and D. Spencer, Partitions of large multipartites with congruence conditions. I, Trans. Amer. Math. Soc. 219 (1976), 299–322. MR 401688
- [Sel89] Atle Selberg, The history of Rademacher’s formula for the partition function , Normat 37 (1989), no. 4, 141–146, 176. MR 1042708
- [Sil10a] Andrew V. Sills, A Rademacher type formula for partitions and overpartitions, Int. J. Math. Math. Sci. (2010), Art. ID 630458, 21. MR 2606900
- [Sil10b] by same author, Rademacher-type formulas for restricted partition and overpartition functions, Ramanujan J. 23 (2010), no. 1-3, 253–264. MR 2739216
- [Sil10c] by same author, Towards an automation of the circle method, Gems in experimental mathematics, Contemp. Math., vol. 517, Amer. Math. Soc., Providence, RI, 2010, pp. 321–338. MR 2731079
- [Spe73] D. Spencer, Partitions of large multipartites with congruence conditions, ProQuest LLC, Ann Arbor, MI, 1973, Thesis (Ph.D.)–University of Surrey (United Kingdom). MR 3809911
- [Sub72] V. V. Subrahmanyasastri, Partitions with congruence conditions, J. Indian Math. Soc. (N.S.) 36 (1972), 177–194 (1973). MR 325560
- [SZ12] Andrew V. Sills and Doron Zeilberger, Formulæfor the number of partitions of into at most parts (using the quasi-polynomial ansatz), Adv. in Appl. Math. 48 (2012), no. 5, 640–645. MR 2920836
- [Usp20] J.V. Uspensky, Asymptotic Formulae for Numerical Functions which Occur in the Theory of Partitions [Russian], Bull. Acad. Sci. URSS 14 (1920), 199–218.
- [Van82] S. Vangipuram, Partitions with congruence conditions and color restrictions, Number theory (Mysore, 1981), Lecture Notes in Math., vol. 938, Springer, Berlin-New York, 1982, pp. 157–169. MR 665446
- [Vau81] R. C. Vaughan, The Hardy-Littlewood method, Cambridge Tracts in Mathematics, vol. 80, Cambridge University Press, Cambridge-New York, 1981. MR 628618
- [VSS82] S. Vangipuram and V. V. Subrahmanya Sastri, A problem on partitions with congruence conditions, Indian J. Math. 24 (1982), no. 1-3, 165–174. MR 724333
- [Zag21] Don Zagier, Power partitions and a generalized eta transformation property, Hardy-Ramanujan J. 44 (2021), 1–18. MR 4400882