This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Rademacher’s Formula for the Partition Function

Ze-Yong Kong and Lee-Peng Teo Department of Mathematics, Xiamen University Malaysia
Jalan Sunsuria, Bandar Sunsuria, 43900, Sepang, Selangor, Malaysia.
[email protected], [email protected]
Abstract.

For a positive integer nn, let p(n)p(n) be the number of ways to express nn as a sum of positive integers. In this note, we revisit the derivation of the Rademacher’s convergent series for p(n)p(n) in a pedagogical way, with all the details given. We also derive the leading asymptotic behavior of p(n)p(n) when nn approaches infinity. Some numerical results are tabled.

1. Introduction

Given a positive integer nn, let p(n)p(n) be the number of ways to express nn as a sum of positive integer. For example, when n=7n=7, the followings are the different ways to express 7 as a sum of positive integers.

7\displaystyle{\color[rgb]{1,0,0}7}
6+1, 5+2, 4+3\displaystyle{\color[rgb]{100,0,150}6+1},\;{\color[rgb]{0,120,20}5+2},\;{\color[rgb]{100,0,150}4+3}
5+1+1, 4+2+1, 3+3+1, 3+2+2\displaystyle{\color[rgb]{1,0,0}5+1+1},\;{\color[rgb]{0,0,1}4+2+1},\;{\color[rgb]{1,0,0}3+3+1},\;{\color[rgb]{0,0,1}3+2+2}
4+1+1+1, 3+2+1+1, 2+2+2+1\displaystyle{\color[rgb]{100,0,150}4+1+1+1},\;{\color[rgb]{0,120,20}3+2+1+1},\;{\color[rgb]{100,0,150}2+2+2+1}
3+1+1+1+1, 2+2+1+1+1\displaystyle{\color[rgb]{1,0,0}3+1+1+1+1},\;{\color[rgb]{0,0,1}2+2+1+1+1}
2+1+1+1+1+1\displaystyle{\color[rgb]{100,0,150}2+1+1+1+1+1}
1+1+1+1+1+1+1\displaystyle{\color[rgb]{0,120,20}1+1+1+1+1+1+1}

There are 15 of them. Hence, p(7)=15p(7)=15.

Let p(0)=1p(0)=1. Then the generating function of the sequence {p(n)}\{p(n)\} is

F(x)=n=0p(n)xn=m=1(1xm)1.\displaystyle F(x)=\sum_{n=0}^{\infty}p(n)x^{n}=\prod_{m=1}^{\infty}(1-x^{m})^{-1}. (1.1)

The calculation of p(n)p(n) using the generating function is inefficient. A more efficient method is to make use of the Euler pentagonal number theorem, which states that

m=1(1xm)=1+n=1(1)n(xω1(n)+xω2(n)),\prod_{m=1}^{\infty}(1-x^{m})=1+\sum_{n=1}^{\infty}(-1)^{n}\left(x^{\omega_{1}(n)}+x^{\omega_{2}(n)}\right),

where

ω1(n)=k=0n1(1+3k)=3n2n2\omega_{1}(n)=\sum_{k=0}^{n-1}(1+3k)=\frac{3n^{2}-n}{2}

are the pentagonal numbers, and

ω2(n)=ω1(n)=3n2+n2.\omega_{2}(n)=\omega_{1}(-n)=\frac{3n^{2}+n}{2}.

From this, one can deduce the following recursive formula to calculate p(n)p(n):

p(n)=k1w2(k)n(1)k+1{p(nω1(k)+p(nω2(k)}.p(n)=\sum_{\begin{subarray}{c}k\geq 1\\ w_{2}(k)\leq n\end{subarray}}(-1)^{k+1}\left\{p(n-\omega_{1}(k)+p(n-\omega_{2}(k)\right\}. (1.2)

Notice that for all n1n\geq 1,

ω2(n)ω1(n)=n,ω1(n+1)ω2(n)=2n+1.\omega_{2}(n)-\omega_{1}(n)=n,\hskip 28.45274pt\omega_{1}(n+1)-\omega_{2}(n)=2n+1.

This implies that

ω1(n)<ω2(n)<ω1(n+1).\omega_{1}(n)<\omega_{2}(n)<\omega_{1}(n+1).

The sum on the right hand side of (1.2) only involves approximately 22n/3\displaystyle 2\sqrt{2n/3} terms. Hence, it is an effective formula to calculate p(n)p(n). By using ω1(k)\omega_{1}(k) and ω2(k)\omega_{2}(k) up to k=100k=100, one can calculate all p(n)p(n) for 1n150501\leq n\leq 15050. Selected values of p(n)p(n) for nn in this range are tabled in Table 4.

In 1918, Hardy and Ramanujan [HR18] showed that p(n)p(n) satisfies the asymptotic formula:

p(n)14n3exp(π2n3)whenn.p(n)\sim\frac{1}{4n\sqrt{3}}\exp\left(\pi\sqrt{\frac{2n}{3}}\right)\hskip 28.45274pt\text{when}\;n\rightarrow\infty.

More precisely, if we let

P1(n)=14n3exp(π2n3),\displaystyle P_{1}(n)=\frac{1}{4n\sqrt{3}}\exp\left(\pi\sqrt{\frac{2n}{3}}\right), (1.3)

then

limnp(n)P1(n)=1.\lim_{n\rightarrow\infty}\frac{p(n)}{P_{1}(n)}=1.

This was discovered independently by Uspensky [Usp20] in 1920.

In fact, in [HR18], Hardy and Ramanujan obtained the following aymptotic formula

p(n)=k<αnPk(n)+O(n1/4),\displaystyle p(n)=\sum_{k<\alpha\sqrt{n}}P_{k}(n)+O\left(n^{-1/4}\right), (1.4)

where α\alpha is a positive constant, and P1(n)P_{1}(n) is the leading term given by (1.3). The terms P2(n)P_{2}(n), P3(n)P_{3}(n), \cdots, have similar forms, but with a constant smaller than π2/3\displaystyle\pi\sqrt{2/3} in the exponential term.

A drawback of the asymptotic formula (1.4) is that the infinite series

k=1Pk(n)\sum_{k=1}^{\infty}P_{k}(n)

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 n1n\geq 1, the partition function p(n)p(n) is represented by the convergent series

p(n)=k=11π2Ak(n)kddn(sinh{πk23(n124)}n124),\displaystyle p(n)=\sum_{k=1}^{\infty}\frac{1}{\pi\sqrt{2}}A_{k}(n)\sqrt{k}\frac{d}{dn}\left(\frac{\displaystyle\sinh\left\{\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}\right\}}{\displaystyle\sqrt{n-\frac{1}{24}}}\right),

where Ak(n)A_{k}(n) 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 NN is of order n\sqrt{n}, then the remainder after NN terms is of order n1/4n^{-1/4}.

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 Poincare´\acute{\text{e}} 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 ={τ|Imτ>0}\displaystyle\mathbb{H}=\left\{\tau\,|\,\text{Im}\,\tau>0\right\} by the equation

η(τ)=eπiτ/12n=1(1e2πinτ).\displaystyle\eta(\tau)=e^{\pi i\tau/12}\prod_{n=1}^{\infty}\left(1-e^{2\pi in\tau}\right). (2.1)

The generating function F(x)F(x) (1.1) for the sequence {p(n)}\{p(n)\} is related to η(τ)\eta(\tau) by

η(τ)=eπiτ/12F(e2πiτ).\displaystyle\eta(\tau)=e^{\pi i\tau/12}F(e^{2\pi i\tau}). (2.2)

The eta function η(τ)\eta(\tau) is closely related to the theory of modular forms [Apo90]. Its 24th power, η(τ)24\eta(\tau)^{24}, 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

τaτ+bcτ+d\tau\mapsto\frac{a\tau+b}{c\tau+d}

defined by an element [abcd]\displaystyle\begin{bmatrix}a&b\\ c&d\end{bmatrix} of the modular group Γ=PSL(2,)\Gamma=\text{PSL}\,(2,\mathbb{Z}).

Before stating the transformation formula, let us define the Dedekind sum.

Definition 2.1.

If hh is an integer, kk is a positive integer larger than 1, the Dedekind sum s(h,k)s(h,k) is defined as

s(h,k)=r=1k1rk(hrkhrk12).\displaystyle s(h,k)=\sum_{r=1}^{k-1}\frac{r}{k}\left(\frac{hr}{k}-\left\lfloor\frac{hr}{k}\right\rfloor-\frac{1}{2}\right). (2.3)

When k=1k=1, s(h,1)s(h,1) is defined as 0 for any integer hh.

Theorem 2.2 (The Dedekind’s Functional Equation).

Let a,b,c,da,b,c,d be integers with adbc=1ad-bc=1 and c>0c>0. Under the fractional linear transformation

τaτ+bcτ+d,\tau\mapsto\frac{a\tau+b}{c\tau+d},

the Dedekind eta function satisfies the transformation formula.

η(aτ+bcτ+d)=exp{πi(a+d12c+s(d,c))}{i(cτ+d)}1/2η(τ).\eta\left(\frac{a\tau+b}{c\tau+d}\right)=\exp\left\{\pi i\left(\frac{a+d}{12c}+s(-d,c)\right)\right\}\left\{-i(c\tau+d)\right\}^{1/2}\eta(\tau). (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 ν\nu be a complex number. The modified Bessel function Iν(x)I_{\nu}(x) is defined as

Iν(x)=(x2)νj=01j!Γ(ν+j+1)(x2)2j.\displaystyle I_{\nu}(x)=\left(\frac{x}{2}\right)^{\nu}\sum_{j=0}^{\infty}\frac{1}{j!\,\Gamma(\nu+j+1)}\left(\frac{x}{2}\right)^{2j}. (2.5)

Iν(x)I_{\nu}(x) has a contour integral representation given by

Iν(x)=(x2)ν12πicic+itν1exp(t+x24t)𝑑t,I_{\nu}(x)=\left(\frac{x}{2}\right)^{\nu}\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}t^{-\nu-1}\exp\left(t+\frac{x^{2}}{4t}\right)dt,

where cc is a positive number. To prove this, we need the Hankel’s formula.

Let KK be the contour which is a loop around the negative real axis. As shown in Figure 1, it consists of three parts K1K_{1}, K2K_{2} and K3K_{3}, where K1K_{1} and K3K_{3} are the lower and upper edges of a cut in the zz-plane along the negative real-axis, and K2K_{2} is a positively oriented circle of radius c<2πc<2\pi about the origin.

Refer to caption
Figure 1. The contour KK.
Lemma 2.3 (Hankel’s Formula).

When ss\in\mathbb{C}, we have

1Γ(s)=12πiKzsez𝑑z.\displaystyle\frac{1}{\Gamma(s)}=\frac{1}{2\pi i}\int_{K}z^{-s}e^{z}dz.
Proof.

The integral defines an analytic function for all ss\in\mathbb{C}. It is sufficient to prove the formula for ss with s<0s<0. Then the result follows from analytic continuation. When s<0s<0, the integral over K2K_{2} goes to 0 as c0+c\rightarrow 0^{+}. In this limit, we have

12πiKzsez𝑑z\displaystyle\frac{1}{2\pi i}\int_{K}z^{-s}e^{z}dz =eπi12πi0(eπit)set𝑑t+eπi12πi0(eπit)set𝑑t\displaystyle=-e^{-\pi i}\frac{1}{2\pi i}\int_{0}^{\infty}(e^{-\pi i}t)^{-s}e^{-t}dt+e^{\pi i}\frac{1}{2\pi i}\int_{0}^{\infty}(e^{\pi i}t)^{-s}e^{-t}dt
=eπiseπis2πi0tset𝑑t\displaystyle=\frac{e^{\pi is}-e^{-\pi is}}{2\pi i}\int_{0}^{\infty}t^{-s}e^{-t}dt
=sinπsπΓ(1s)\displaystyle=\frac{\sin\pi s}{\pi}\Gamma(1-s)
=1Γ(s).\displaystyle=\frac{1}{\Gamma(s)}.

Using this lemma, we can prove the contour integral formula for the modified Bessel function Iν(x)I_{\nu}(x).

Proposition 2.4.

Let ν\nu be a complex number and let cc be a positive number. If xx is real, then

Iν(x)=(x2)ν12πicic+itν1exp(t+x24t)𝑑t.\displaystyle I_{\nu}(x)=\left(\frac{x}{2}\right)^{\nu}\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}t^{-\nu-1}\exp\left(t+\frac{x^{2}}{4t}\right)dt. (2.6)
Proof.

The integrand

tν1exp(t+x24t)t^{-\nu-1}\exp\left(t+\frac{x^{2}}{4t}\right)

is an analytic function of tt on the domain {t|Imt=0,Ret0}\mathbb{C}-\left\{t\,|\,\text{Im}\,t=0,\;\text{Re}\,t\leq 0\right\}. It decays exponentially fast in the region Retc\text{Re}\,t\leq c when |t||t|\rightarrow\infty. Using Cauchy residue theorem, we can take the limit c0+c\rightarrow 0^{+}, and change the contour to KK. Namely,

f(x)\displaystyle f(x) =(x2)ν12πicic+itν1exp(t+x24t)𝑑t\displaystyle=\left(\frac{x}{2}\right)^{\nu}\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}t^{-\nu-1}\exp\left(t+\frac{x^{2}}{4t}\right)dt
=(x2)ν12πiKtν1exp(t+x24t)𝑑t.\displaystyle=\left(\frac{x}{2}\right)^{\nu}\frac{1}{2\pi i}\int_{K}t^{-\nu-1}\exp\left(t+\frac{x^{2}}{4t}\right)dt.

Now, using the Taylor series of the exponential function, we have

f(x)\displaystyle f(x) =(x2)νj=01j!(x2)2j12πiKtνj1et𝑑t\displaystyle=\left(\frac{x}{2}\right)^{\nu}\sum_{j=0}^{\infty}\frac{1}{j!}\left(\frac{x}{2}\right)^{2j}\frac{1}{2\pi i}\int_{K}t^{-\nu-j-1}e^{t}dt
=(x2)νj=01j!(x2)2j1Γ(ν+j+1),\displaystyle=\left(\frac{x}{2}\right)^{\nu}\sum_{j=0}^{\infty}\frac{1}{j!}\left(\frac{x}{2}\right)^{2j}\frac{1}{\Gamma(\nu+j+1)},

where we have used Lemma 2.3. It follows from (2.5) that

f(x)=Iν(x).f(x)=I_{\nu}(x).

For our applications, we are interested in the case where ν=3/2\nu=3/2.

Proposition 2.5.

I32(x)I_{\frac{3}{2}}(x) has an explicit formula given by

I32(x)=2xπddxsinhxx.\displaystyle I_{\frac{3}{2}}(x)=\sqrt{\frac{2x}{\pi}}\frac{d}{dx}\frac{\sinh x}{x}.
Proof.

This is a direct computation using (2.5). Using the fact that

Γ(j+52)=(2j+3)(2j+1)×12j+2Γ(12)=π(2j+3)(2j+1)×12j+2,\Gamma\left(j+\frac{5}{2}\right)=\frac{(2j+3)(2j+1)\ldots\times 1}{2^{j+2}}\Gamma\left(\frac{1}{2}\right)=\sqrt{\pi}\frac{(2j+3)(2j+1)\ldots\times 1}{2^{j+2}},

we have

I32(x)\displaystyle I_{\frac{3}{2}}(x) =(x2)32j=01j!Γ(j+52)(x2)2j\displaystyle=\left(\frac{x}{2}\right)^{\frac{3}{2}}\sum_{j=0}^{\infty}\frac{1}{j!\,\Gamma(j+\frac{5}{2})}\left(\frac{x}{2}\right)^{2j}
=2π(x2)12j=02j+2(2j+3)!x2j+1\displaystyle=\frac{2}{\sqrt{\pi}}\left(\frac{x}{2}\right)^{\frac{1}{2}}\sum_{j=0}^{\infty}\frac{2j+2}{(2j+3)!}x^{2j+1}
=2xπddxj=01(2j+3)!x2j+2\displaystyle=\sqrt{\frac{2x}{\pi}}\frac{d}{dx}\sum_{j=0}^{\infty}\frac{1}{(2j+3)!}x^{2j+2}
=2xπddxj=01(2j+1)!x2j\displaystyle=\sqrt{\frac{2x}{\pi}}\frac{d}{dx}\sum_{j=0}^{\infty}\frac{1}{(2j+1)!}x^{2j}
=2xπddxsinhxx.\displaystyle=\sqrt{\frac{2x}{\pi}}\frac{d}{dx}\frac{\sinh x}{x}.


2.3. Farey Series and Ford Circles

Let nn be a positive integer. The set of Farey fractions of order nn, denoted by FnF_{n}, is the set of all reduced fractions in the closed interval [0,1][0,1] whose denominator is not larger than nn, listed in increasing order of magnitude.

Example 2.6.

The sets FnF_{n}, 1n101\leq n\leq 10, are given by

F1:\displaystyle F_{1}: 01,11\displaystyle\;\frac{0}{1},\frac{1}{1}
F2:\displaystyle F_{2}: 01,12,11\displaystyle\;\frac{0}{1},\frac{1}{2},\frac{1}{1}
F3:\displaystyle F_{3}: 01,13,12,23,11\displaystyle\frac{0}{1},\frac{1}{3},\frac{1}{2},\frac{2}{3},\frac{1}{1}
F4:\displaystyle F_{4}: 01,14,13,12,23,34,11\displaystyle\frac{0}{1},\frac{1}{4},\frac{1}{3},\frac{1}{2},\frac{2}{3},\frac{3}{4},\frac{1}{1}
F5:\displaystyle F_{5}: 01,15,14,13,25,12,35,23,34,45,11\displaystyle\frac{0}{1},\frac{1}{5},\frac{1}{4},\frac{1}{3},\frac{2}{5},\frac{1}{2},\frac{3}{5},\frac{2}{3},\frac{3}{4},\frac{4}{5},\frac{1}{1}
F6:\displaystyle F_{6}: 01,16,15,14,13,25,12,35,23,34,45,56,11\displaystyle\frac{0}{1},\frac{1}{6},\frac{1}{5},\frac{1}{4},\frac{1}{3},\frac{2}{5},\frac{1}{2},\frac{3}{5},\frac{2}{3},\frac{3}{4},\frac{4}{5},\frac{5}{6},\frac{1}{1}
F7:\displaystyle F_{7}: 01,17,16,15,14,27,13,25,37,12,47,35,23,57,34,45,56,67,11\displaystyle\frac{0}{1},\frac{1}{7},\frac{1}{6},\frac{1}{5},\frac{1}{4},\frac{2}{7},\frac{1}{3},\frac{2}{5},\frac{3}{7},\frac{1}{2},\frac{4}{7},\frac{3}{5},\frac{2}{3},\frac{5}{7},\frac{3}{4},\frac{4}{5},\frac{5}{6},\frac{6}{7},\frac{1}{1}
F8:\displaystyle F_{8}: 01,18,17,16,15,14,27,13,38,25,37,12,47,35,58,23,57,34,45,56,67,78,11\displaystyle\frac{0}{1},\frac{1}{8},\frac{1}{7},\frac{1}{6},\frac{1}{5},\frac{1}{4},\frac{2}{7},\frac{1}{3},\frac{3}{8},\frac{2}{5},\frac{3}{7},\frac{1}{2},\frac{4}{7},\frac{3}{5},\frac{5}{8},\frac{2}{3},\frac{5}{7},\frac{3}{4},\frac{4}{5},\frac{5}{6},\frac{6}{7},\frac{7}{8},\frac{1}{1}
F9:\displaystyle F_{9}: 01,19,18,17,16,15,29,14,27,13,38,25,37,49,12,59,47,35,58,23,57,34,79,45,56,67,78,89,11\displaystyle\frac{0}{1},\frac{1}{9},\frac{1}{8},\frac{1}{7},\frac{1}{6},\frac{1}{5},\frac{2}{9},\frac{1}{4},\frac{2}{7},\frac{1}{3},\frac{3}{8},\frac{2}{5},\frac{3}{7},\frac{4}{9},\frac{1}{2},\frac{5}{9},\frac{4}{7},\frac{3}{5},\frac{5}{8},\frac{2}{3},\frac{5}{7},\frac{3}{4},\frac{7}{9},\frac{4}{5},\frac{5}{6},\frac{6}{7},\frac{7}{8},\frac{8}{9},\frac{1}{1}
F10:\displaystyle F_{10}: 01,110,19,18,17,16,15,29,14,27,310,13,38,25,37,49,12,59,47,35,58,23,710,57,34,79,45,56,67,78,89,910,11\displaystyle\frac{0}{1},\frac{1}{10},\frac{1}{9},\frac{1}{8},\frac{1}{7},\frac{1}{6},\frac{1}{5},\frac{2}{9},\frac{1}{4},\frac{2}{7},\frac{3}{10},\frac{1}{3},\frac{3}{8},\frac{2}{5},\frac{3}{7},\frac{4}{9},\frac{1}{2},\frac{5}{9},\frac{4}{7},\frac{3}{5},\frac{5}{8},\frac{2}{3},\frac{7}{10},\frac{5}{7},\frac{3}{4},\frac{7}{9},\frac{4}{5},\frac{5}{6},\frac{6}{7},\frac{7}{8},\frac{8}{9},\frac{9}{10},\frac{1}{1}

Obviously, for n2n\geq 2, Fn1F_{n-1} is a subset of FnF_{n}. Moreover,

|Fn||Fn1|=φ(n),|F_{n}|-|F_{n-1}|=\varphi(n),

which is the number of positive integers less than or equal to nn that are relatively prime to nn.

To be more precise, one adds in φ(n)\varphi(n) fractions of the form hn\displaystyle\frac{h}{n} with 1hn1\leq h\leq n and (h,n)=1(h,n)=1 into the set Fn1F_{n-1} to obtain FnF_{n}. To determine the precise places to insert these fractions, one need the following lemmas.

Lemma 2.7.

If aa, bb, cc and dd are positive integers such that

ab<cd,\displaystyle\frac{a}{b}<\frac{c}{d},

then

ab<a+cb+d<cd.\frac{a}{b}<\frac{a+c}{b+d}<\frac{c}{d}.
Proof.

We have

cdab=bcadbd>0\displaystyle\frac{c}{d}-\frac{a}{b}=\frac{bc-ad}{bd}>0

Since bb and dd are positive, we have

bcad>0.bc-ad>0.

Hence,

a+cb+dab\displaystyle\frac{a+c}{b+d}-\frac{a}{b} =bcadb(b+d)>0,\displaystyle=\frac{bc-ad}{b(b+d)}>0,
cda+cb+d\displaystyle\frac{c}{d}-\frac{a+c}{b+d} =bcadd(b+d)>0.\displaystyle=\frac{bc-ad}{d(b+d)}>0.

This proves the assertion. ∎

Lemma 2.8.

If aa, bb, cc, dd, hh and kk are positive integers such that

ab<hk<cd\displaystyle\frac{a}{b}<\frac{h}{k}<\frac{c}{d} (2.7)

and bcad=1bc-ad=1, then

kb+d.k\geq b+d.
Proof.

The inequalities in (2.7) imply that

bhak\displaystyle bh-ak 1,\displaystyle\geq 1,
ckdh\displaystyle ck-dh 1.\displaystyle\geq 1.

Therefore,

k\displaystyle k =k(bcad)\displaystyle=k(bc-ad)
=b(ckdh)+d(bhak)\displaystyle=b(ck-dh)+d(bh-ak)
b+d.\displaystyle\geq b+d.

From this, we can deduce the following theorem.

Theorem 2.9.

Let aa, bb, cc, dd be positive integers such that

ab<cd,\displaystyle\frac{a}{b}<\frac{c}{d},

and bcad=1bc-ad=1. If

max{b,d}nb+d1,\max\{b,d\}\leq n\leq b+d-1,

then a/ba/b and c/dc/d are consecutive terms in FnF_{n}.

Proof.

Since max{b,d}n\max\{b,d\}\leq n, bb and dd are less than or equal to nn. Hence, a/ba/b and c/dc/d are in FnF_{n}.
By Lemma 2.8, no fractions h/kh/k with kb+d1k\leq b+d-1 can satisfy

ab<hk<cd.\frac{a}{b}<\frac{h}{k}<\frac{c}{d}.

Since FnF_{n} only contains reduced fractions with denominator at most nn and nb+d1n\leq b+d-1, this shows that a/ba/b and c/dc/d are consecutive terms in FnF_{n}. ∎

Now we can give the precise algorithm to construct FnF_{n} from Fn1F_{n-1}. First notice that for the two terms

ab=01andcd=11\frac{a}{b}=\frac{0}{1}\quad\text{and}\quad\frac{c}{d}=\frac{1}{1}

in F1F_{1}, a=0a=0, b=c=d=1b=c=d=1 and thus

bcad=1.bc-ad=1.

Now to construct F2F_{2}, we insert the term

hk=a+cb+d=12\frac{h}{k}=\frac{a+c}{b+d}=\frac{1}{2}

in between them. It is easy to check that if bcad=1bc-ad=1, h=a+ch=a+c and k=b+dk=b+d, then

bhak=1andckdh=1.bh-ak=1\quad\text{and}\quad ck-dh=1.

In general, assume that Fn1F_{n-1} has been constructed and for any two consecutive terms a/ba/b and c/dc/d in Fn1F_{n-1}, bcad=1bc-ad=1. To construct FnF_{n}, we need to insert the φ(n)\varphi(n) fractions h/nh/n with 1hn1\leq h\leq n and (h,n)=1(h,n)=1 into Fn1F_{n-1}. To do this, we look for consecutive pairs of reduced fractions a/ba/b and c/dc/d in Fn1F_{n-1} for which a/b<c/da/b<c/d and b+d=nb+d=n. Let h=a+ch=a+c. Since 0ab0\leq a\leq b and 0cd0\leq c\leq d, we have

0h=a+cb+d.0\leq h=a+c\leq b+d.

Lemma 2.7 implies that

ab<hn<cd.\frac{a}{b}<\frac{h}{n}<\frac{c}{d}.

By induction hypothesis, bcad=1bc-ad=1. As we have shown, this implies that

bhan=1andcndh=1.bh-an=1\quad\text{and}\quad cn-dh=1.

It remains to show that we can find exactly φ(n)\varphi(n) pairs of consecutive terms a/ba/b and c/dc/d in Fn1F_{n-1} such that b+d=nb+d=n.

For n2n\geq 2, there are exactly φ(n)\varphi(n) positive integers bb with 1bn1\leq b\leq n and (b,n)=1(b,n)=1. In fact, we must also have bn1b\leq n-1. For each of these integers bb, d=nbd=n-b is also relatively prime to nn and 1dn11\leq d\leq n-1. Since b+d=nb+d=n, bb and dd must also be relatively prime. There exists s0s_{0} and t0t_{0} such that

bs0dt0=1.bs_{0}-dt_{0}=1.

Any pairs of integers (s,t)(s,t) with bsdt=1bs-dt=1 can be written as

s=s0+du,t=t0+bu\displaystyle s=s_{0}+du,\quad t=t_{0}+bu

for some integer uu. Let u0u_{0} be the smallest one so that c=s0+du01c=s_{0}+du_{0}\geq 1. Then cdc\leq d. If a=t0+bu0a=t_{0}+bu_{0}, then bcad=1bc-ad=1. Since

ad=bc1<bcbd,ad=bc-1<bc\leq bd,

we find that a<ba<b. Since b1,c1b\geq 1,c\geq 1, we have ad0ad\geq 0 and thus a0a\geq 0. In other words, we have shown that for each of the φ(n)\varphi(n) positive integers bb so that 1bn1\leq b\leq n and (b,n)=1(b,n)=1, there are unique integers a,ca,c and dd such that

0ab<cd10\leq\frac{a}{b}<\frac{c}{d}\leq 1

and bcad=1bc-ad=1. This shows that we can find exactly φ(n)\varphi(n) consecutive pairs in Fn1F_{n-1} for us to insert the fractions h/nh/n with 1hn1\leq h\leq n and (h,n)=1(h,n)=1.

Given a reduced fraction h/kh/k, the Ford circle C(h,k)C(h,k) (Figure 2) is the circle in the complex plane with center at

ch,k=hk+i2k2c_{h,k}=\frac{h}{k}+\frac{i}{2k^{2}}

and radius

rh,k=12k2.r_{h,k}=\frac{1}{2k^{2}}.

Obviously, each Ford circle C(h,k)C(h,k) touches the xx-axis at the point (h/k,0)(h/k,0).

Refer to caption
Figure 2. The Ford circle C(h,k)C(h,k).
Theorem 2.10.

Two distinct Ford circles C(a,b)C(a,b) and C(c,d)C(c,d) are either tangent to each other or disjoint. They are tangent if and only if bcad=±1bc-ad=\pm 1. In particular, if a/ba/b and c/dc/d are two consecutive Farey fractions, then the two Ford circles C(a,b)C(a,b) and C(c,d)C(c,d) are tangent to each other.

Proof.
Refer to caption
Figure 3. The circles C(a,b)C(a,b) and C(c,d)C(c,d).

Let DD be the distance between the centers of C(a,b)C(a,b) and C(c,d)C(c,d). Then

D2=(abcd)2+(12b212d2)2.\displaystyle D^{2}=\left(\frac{a}{b}-\frac{c}{d}\right)^{2}+\left(\frac{1}{2b^{2}}-\frac{1}{2d^{2}}\right)^{2}.

The sum of the radii of the two circles is

S=12b2+12d2.S=\frac{1}{2b^{2}}+\frac{1}{2d^{2}}.

Notice that the two circles are disjoint if and only if D>SD>S, and the two circles are tangent to each other if and only if D=SD=S. A straightforward computation gives

D2S2\displaystyle D^{2}-S^{2} =(bcad)21b2d2.\displaystyle=\frac{(bc-ad)^{2}-1}{b^{2}d^{2}}.

The two circles are distinct, so bcad0bc-ad\neq 0. Therefore, (bcad)21(bc-ad)^{2}\geq 1. This shows that

D2S2D^{2}\geq S^{2}

and so DSD\geq S. Moreover, D=SD=S if and only if (bcad)2=1(bc-ad)^{2}=1, if and only if bcad=±1bc-ad=\pm 1. ∎

Now we study the points of tangency of two Ford circles that are tangent to each other.

Theorem 2.11.

Let

h1k1<hk<h2k2\frac{h_{1}}{k_{1}}<\frac{h}{k}<\frac{h_{2}}{k_{2}}

be three consecutive Farey fractions. The points of tangency of C(h,k)C(h,k) with C(h1,k1)C(h_{1},k_{1}) and C(h2,k2)C(h_{2},k_{2}) are given by

α1(h,k)\displaystyle\alpha_{1}(h,k) =hkk1k(k2+k12)+ik2+k12,\displaystyle=\frac{h}{k}-\frac{k_{1}}{k(k^{2}+k_{1}^{2})}+\frac{i}{k^{2}+k_{1}^{2}}, (2.8a)
α2(h,k)\displaystyle\alpha_{2}(h,k) =hk+k2k(k2+k22)+ik2+k22.\displaystyle=\frac{h}{k}+\frac{k_{2}}{k(k^{2}+k_{2}^{2})}+\frac{i}{k^{2}+k_{2}^{2}}. (2.8b)
Proof.

Notice that the condition of tangency implies that

hk1kh1=1,kh2hk2=1.hk_{1}-kh_{1}=1,\quad kh_{2}-hk_{2}=1.
Refer to caption
Figure 4. The circles C(a,b)C(a,b) and C(c,d)C(c,d).

As shown in Figure 4, the point of tangency α1(h,k)\alpha_{1}(h,k) between C(h1,k1)C(h_{1},k_{1}) and C(h,k)C(h,k) is the unique point on the line segment joining ch1,k1c_{h_{1},k_{1}} and ch,kc_{h,k} such that

ch,kα1(h,k)ch,kch1,k1=rh,krh1,k1+rh,k.\frac{c_{h,k}-\alpha_{1}(h,k)}{c_{h,k}-c_{h_{1},k_{1}}}=\frac{r_{h,k}}{r_{h_{1},k_{1}}+r_{h,k}}.

A direct computation proves (2.8a). Similarly, α2(h,k)\alpha_{2}(h,k) is the unique point on the line segment joining ch,kc_{h,k} to ch2,k2c_{h_{2},k_{2}} such that

α2(h,k)ch,kch2,k2ch,k=rh,krh2,k2+rh,k.\frac{\alpha_{2}(h,k)-c_{h,k}}{c_{h_{2},k_{2}}-c_{h,k}}=\frac{r_{h,k}}{r_{h_{2},k_{2}}+r_{h,k}}.

This gives (2.8b). ∎

For each positive integer NN, we construct a path P(N)P(N) joining the points ii and 1+i1+i in the complex plane as follows. Let

0,h1k1,,hmN1kmN1,10,\frac{h_{1}}{k_{1}},\ldots,\frac{h_{m_{N}-1}}{k_{m_{N}-1}},1

be the Farey fractions in FNF_{N} listed in increasing order. Let

h0k0=0andhmNkmN=1.\frac{h_{0}}{k_{0}}=0\quad\text{and}\quad\frac{h_{m_{N}}}{k_{m_{N}}}=1.

For each 1jmN11\leq j\leq m_{N}-1, let AjA_{j} be the point of tangency between C(hj1,kj1)C(h_{j-1},k_{j-1}) and C(hj,kj)C(h_{j},k_{j}); and let BjB_{j} be the point of tangency between C(hj,kj)C(h_{j},k_{j}) and C(hj+1,kj+1)C(h_{j+1},k_{j+1}). Then Bj=Aj+1B_{j}=A_{j+1} for 1jmN21\leq j\leq m_{N}-2. AjA_{j} and BjB_{j} divide the circle C(hj,kj)C(h_{j},k_{j}) into two arcs, the upper arc and the lower arc. Let UjU_{j} be the upper arc. Finally, let U0U_{0} be the arc of the circle C(h0,k0)C(h_{0},k_{0}) that joins the point ii to A1A_{1} that is in the right half plane Rez0\text{Re}\,z\geq 0, and let UmNU_{m_{N}} be the arc of the circle C(hmN,kmN)C(h_{m_{N}},k_{m_{N}}) that joins BmN1B_{m_{N}-1} to the point 1+i1+i that is in the left half-plane Rez1\text{Re}\;z\leq 1. The path P(N)P(N) is defined as

P(N)=j=0mNUj,\displaystyle P(N)=\bigcup_{j=0}^{m_{N}}U_{j}, (2.9)

following the orientation from the point ii to the point 1+i1+i.

Refer to caption
Figure 5. The path P(3)P(3) for the Farey series F3:01,13,12,23,11\displaystyle F_{3}:\frac{0}{1},\frac{1}{3},\frac{1}{2},\frac{2}{3},\frac{1}{1}.
Refer to caption
Figure 6. The path P(4)P(4) for the Farey series F4:01,14,13,12,23,34,11\displaystyle F_{4}:\frac{0}{1},\frac{1}{4},\frac{1}{3},\frac{1}{2},\frac{2}{3},\frac{3}{4},\frac{1}{1}.
Refer to caption
Figure 7. The path P(5)P(5) for the Farey series F5:01,15,14,13,25,12,35,23,34,45,11\displaystyle F_{5}:\frac{0}{1},\frac{1}{5},\frac{1}{4},\frac{1}{3},\frac{2}{5},\frac{1}{2},\frac{3}{5},\frac{2}{3},\frac{3}{4},\frac{4}{5},\frac{1}{1}.

Let us study the image of P(N)P(N) under the mapping z=e2πiτz=e^{2\pi i\tau}.

Lemma 2.12.

Let

𝔻={z||z|<1}\mathbb{D}=\left\{z\in\mathbb{C}\,|\,|z|<1\right\}

be the unit disc on the complex plane and let

𝒮={τ| 0Reτ1,Imτ>0}\mathcal{S}=\left\{\tau\in\mathbb{C}\,|\,0\leq\text{Re}\,\tau\leq 1,\text{Im}\,\tau>0\right\}

be an infinite vertical strip on the upper half plane. The transformation

z=e2πiτz=e^{2\pi i\tau}

is a conformal mapping that maps the vertical strip 𝒮\mathcal{S} in the τ\tau-plane onto the punctured unit disc 𝔻{0}\mathbb{D}\setminus\{0\} in the zz-plane. The mapping is one-to-one on the interior of 𝒮\mathcal{S}. The image of P(N)P(N) under this mapping is a simple closed contour 𝒞\mathcal{C} that enclosed the origin.

Proof.
Refer to caption
Figure 8. The map z=e2πiτz=e^{2\pi i\tau}.

The mapping

z=e2πiτz=e^{2\pi i\tau}

is clearly analytic, and so it is a conformal mapping. Let z=x+iy=reiθz=x+iy=re^{i\theta} and τ=σ+iλ\tau=\sigma+i\lambda. Then we find that

reiθ=e2πiσe2πλ.re^{i\theta}=e^{2\pi i\sigma}e^{-2\pi\lambda}.

As λ>0\lambda>0, we find that

0<e2πλ<1.0<e^{-2\pi\lambda}<1.

On the other hand, as σ\sigma varies from 0 to 1, θ=2πσ\theta=2\pi\sigma varies from 0 to 2π2\pi. Hence, we see that the mapping transforms 𝒮\mathcal{S} onto 𝔻{0}\mathbb{D}\setminus\{0\}, and it is one-to-one except on the two vertical lines σ=0\sigma=0 and σ=1\sigma=1.

Let 𝒞\mathcal{C} be the image of P(N)P(N) under the map z=e2πiτz=e^{2\pi i\tau}. Since the initial and end points of P(N)P(N) are ii and 1+i1+i, and they are mapped to the same point in the unit disc 𝔻\mathbb{D}, and so 𝒞\mathcal{C} is a closed curve. Except for the initial and end points, the rest of P(N)P(N) lies in the interior of 𝒮\mathcal{S}. Since P(N)P(N) is a simple curve and the mapping z=e2πiτz=e^{2\pi i\tau} is one-to-one in the interior of 𝒮\mathcal{S}, 𝒞\mathcal{C} is a simple closed curve. ∎

To derive the Rademacher’s formula, we are extracting the coefficient p(n)p(n) from the generating function F(z)=F(e2πiτ)F(z)=F(e^{2\pi i\tau}). As a function of τ\tau, F(e2πiτ)F(e^{2\pi i\tau}) is periodic modulo 1. Hence, we can extend the path P(N)P(N) periodically modulo 1.

The Farey fractions in FNF_{N} are given by

h0k0,h1k1,,hmN1kmN1,hmNkmN,\frac{h_{0}}{k_{0}},\frac{h_{1}}{k_{1}},\ldots,\frac{h_{m_{N}-1}}{k_{m_{N}}-1},\frac{h_{m_{N}}}{k_{m_{N}}},

where

h0k0=01,h1k1=1N,hmN1kmN1=N1N,hmNkmN=11.\frac{h_{0}}{k_{0}}=\frac{0}{1},\quad\frac{h_{1}}{k_{1}}=\frac{1}{N},\quad\frac{h_{m_{N}-1}}{k_{m_{N}}-1}=\frac{N-1}{N},\quad\frac{h_{m_{N}}}{k_{m_{N}}}=\frac{1}{1}.

Define

hmN+1kmN+1=h1k1+1=N+1N.\frac{h_{m_{N}+1}}{k_{m_{N}}+1}=\frac{h_{1}}{k_{1}}+1=\frac{N+1}{N}.

By (2.8), U0U_{0} is an arc joining the point ii to

u1=N+iN2+1;u_{1}=\frac{N+i}{N^{2}+1};

and UmNU_{m_{N}} is an arc joining the point

u2=N2N+1+iN2+1u_{2}=\frac{N^{2}-N+1+i}{N^{2}+1}

to 1+i1+i. Translating U0U_{0} by 1 and combine with UmNU_{m_{N}}, we obtain a single arc of the circle |z1i|=1/2|z-1-i|=1/2 joining u2u_{2} to u1+1u_{1}+1.

Henceforth, we regard the path P(N)P(N) as a path consists of the union of mNm_{N} arcs U~1\widetilde{U}_{1}, \ldots, U~mN\widetilde{U}_{m_{N}}, corresponding to the fractions hj/kjh_{j}/k_{j} with 1jmN1\leq j\leq m_{N}.

P(N)=j=1mNU~j,\displaystyle P(N)=\bigcup_{j=1}^{m_{N}}\widetilde{U}_{j}, (2.10)

For 1jmN11\leq j\leq m_{N}-1, U~j=Uj\widetilde{U}_{j}=U_{j}; whereas U~mN\widetilde{U}_{m_{N}} is the union of UmNU_{m_{N}} and U0+1U_{0}+1 that we have just described. If we take hmN1/kmN1h_{m_{N}-1}/k_{m_{N}-1}, hmN/kmNh_{m_{N}}/k_{m_{N}} and hmN+1/kmN+1h_{m_{N}+1}/k_{m_{N}+1} as consecutive Farey fractions, U~mN\widetilde{U}_{m_{N}} is the upper arc of the circle C(hmN,kmN)C(h_{m_{N}},k_{m_{N}}) that joins the point α1(hmN,kmN)\alpha_{1}(h_{m_{N}},k_{m_{N}}) to α2(hmN,kmN)\alpha_{2}(h_{m_{N}},k_{m_{N}}) given by (2.8).

Before ending this section, we study further properties of the arc U~j\widetilde{U}_{j} in the Ford circle C(hj,kj)C(h_{j},k_{j}) under the fractional linear transformation

w=ik2(τhk).w=-ik^{2}\left(\tau-\frac{h}{k}\right).
Theorem 2.13.

The mapping

w=ik2(τhk)\displaystyle w=-ik^{2}\left(\tau-\frac{h}{k}\right) (2.11)

transforms the Ford circle C(h,k)C(h,k) in the τ\tau-plane onto a circle C~\tilde{C} in the ww-plane. The center of C~\tilde{C} is the point w=1/2w=1/2 and the radius is 1/21/2. If h1/k1,h/k,h2/k2h_{1}/k_{1},h/k,h_{2}/k_{2} are three consecutive Farey fractions in FNF_{N}, α1(h,k)\alpha_{1}(h,k) is the point of tangency between C(h1,k1)C(h_{1},k_{1}) and C(h,k)C(h,k), α2(h,k)\alpha_{2}(h,k) is the point of tangency between C(h,k)C(h,k) and C(h2,k2)C(h_{2},k_{2}), the images of α1(h,k)\alpha_{1}(h,k) and α2(h,k)\alpha_{2}(h,k) under the mapping (2.11) are w1(h,k)w_{1}(h,k) and w2(h,k)w_{2}(h,k) given by

w1(h,k)=k2k2+k12+ikk1k2+k12,w2(h,k)=k2k2+k22ikk2k2+k22.\begin{split}w_{1}(h,k)&=\frac{k^{2}}{k^{2}+k_{1}^{2}}+i\frac{kk_{1}}{k^{2}+k_{1}^{2}},\\ w_{2}(h,k)&=\frac{k^{2}}{k^{2}+k_{2}^{2}}-i\frac{kk_{2}}{k^{2}+k_{2}^{2}}.\end{split} (2.12)

The upper arc on C(h,k)C(h,k) joining α1(h,k)\alpha_{1}(h,k) to α2(h,k)\alpha_{2}(h,k) is mapped to the arc on C~\tilde{C} joining w1(h,k)w_{1}(h,k) to w2(h,k)w_{2}(h,k) that does not touch the imaginary axis.

Refer to caption
Figure 9. The circle C~\tilde{C} in the ww-plane, which is the image of C(h,k)C(h,k) in the τ\tau-plane under the transformation w=ik2(τhk)\displaystyle w=-ik^{2}\left(\tau-\frac{h}{k}\right).
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 C(h,k)C(h,k) joining α1(h,k)\alpha_{1}(h,k) to α2(h,k)\alpha_{2}(h,k), is the arc on C~\tilde{C} joining w1(h,k)w_{1}(h,k) to w2(h,k)w_{2}(h,k) that is away from the imaginary axis. ∎

Proposition 2.14.

Let C~\tilde{C} be the circle on the complex plane with center at 1/21/2 and radius 1/21/2. For any zC~z\in\tilde{C}, Rez0\text{Re}\,z\geq 0 and

Re(1z)=1.\text{Re}\,\left(\frac{1}{z}\right)=1.
Proof.

If z=x+iyz=x+iy is a point on C~\tilde{C},

(x12)2+y2=14.\left(x-\frac{1}{2}\right)^{2}+y^{2}=\frac{1}{4}.

This implies that

|x12|12\left|x-\frac{1}{2}\right|\leq\frac{1}{2}

and hence

x0.x\geq 0.

On the other hand, zz can be written as

z=12(1+eiθ)z=\frac{1}{2}(1+e^{i\theta})

for some θ[0,2π]\theta\in[0,2\pi]. It follows that

Re{1z}=2(1+cosθ)(1+cosθ)2+sin2θ=1.\text{Re}\,\left\{\frac{1}{z}\right\}=\frac{2(1+\cos\theta)}{(1+\cos\theta)^{2}+\sin^{2}\theta}=1.

Theorem 2.15.

Let h1/k1,h/k,h2/k2h_{1}/k_{1},h/k,h_{2}/k_{2} be three consecutive Farey fractions in FNF_{N}, and let w1(h,k)w_{1}(h,k) and w2(h,k)w_{2}(h,k) be the points given in (2.12) in Theorem 2.13. The moduli of these points are given by

|w1(h,k)|=kk2+k12,|w2(h,k)|=kk2+k22.\displaystyle|w_{1}(h,k)|=\frac{k}{\sqrt{k^{2}+k_{1}^{2}}},\hskip 28.45274pt|w_{2}(h,k)|=\frac{k}{\sqrt{k^{2}+k_{2}^{2}}}. (2.13)

For any point ww on the chord joining w1(h,k)w_{1}(h,k) and w2(h,k)w_{2}(h,k),

|w|2kN+1.|w|\leq\frac{\sqrt{2}k}{N+1}.

Hence, the length the chord is not larger than 22k/(N+1)2\sqrt{2}k/(N+1). On the other hand, the real part of ww is bounded below by k2/(2N2)k^{2}/(2N^{2}), and hence,

Re{1w}>14.\displaystyle\text{Re}\,\left\{\frac{1}{w}\right\}>\frac{1}{4}.
Proof.

Eq. (2.13) is easily verified. To prove the second assertion, notice that if k+k1Nk+k_{1}\leq N, then (h+h1)/(k+k1)(h+h_{1})/(k+k_{1}) is a Farey fraction in FnF_{n} and it lies between h1/k1h_{1}/k_{1} and h/kh/k. This contradicts to h1/k1h_{1}/k_{1} and h/kh/k are consecutive in FNF_{N}. Therefore, we must have k+k1N+1k+k_{1}\geq N+1. Similarly, k+k2N+1k+k_{2}\geq N+1. Hence,

k2+k12(k+k1)22(N+1)22.\displaystyle k^{2}+k_{1}^{2}\geq\frac{(k+k_{1})^{2}}{2}\geq\frac{(N+1)^{2}}{2}.

Similarly,

k2+k22(N+1)22.k^{2}+k_{2}^{2}\geq\frac{(N+1)^{2}}{2}.

If ww is on the chord joining w1(h,k)w_{1}(h,k) and w2(h,k)w_{2}(h,k), then there is a real number tt in the interval [0,1][0,1] such that

w=tw1(h,k)+(1t)w2(h,k).\displaystyle w=tw_{1}(h,k)+(1-t)w_{2}(h,k).

It follows that

|w|\displaystyle|w| t|w1(h,k)|+(1t)|w2(h,k)|\displaystyle\leq t|w_{1}(h,k)|+(1-t)|w_{2}(h,k)|
=tkk2+k12+(1t)kk2+k22\displaystyle=t\frac{k}{\sqrt{k^{2}+k_{1}^{2}}}+(1-t)\frac{k}{\sqrt{k^{2}+k_{2}^{2}}}
2tkN+1+2(1t)kN+1\displaystyle\leq\frac{\sqrt{2}tk}{N+1}+\frac{\sqrt{2}(1-t)k}{N+1}
=2kN+1.\displaystyle=\frac{\sqrt{2}k}{N+1}.

By triangle inequality, the length of the chord is not larger than |w1|+|w2||w_{1}|+|w_{2}|, and hence, it is not larger than 22k/(N+1)2\sqrt{2}k/(N+1).

Now, for the assertion about the real part, we use the fact that k,k1,k2k,k_{1},k_{2} are all not larger than NN. Then

Rew\displaystyle\text{Re}\,w =tRew1(h,k)+(1t)Rew2(h,k)\displaystyle=t\text{Re}\,w_{1}(h,k)+(1-t)\text{Re}\,w_{2}(h,k)
=tk2k2+k12+(1t)k2k2+k22\displaystyle=\frac{tk^{2}}{k^{2}+k_{1}^{2}}+\frac{(1-t)k^{2}}{k^{2}+k_{2}^{2}}
tk22N2+(1t)k22N2\displaystyle\geq t\frac{k^{2}}{2N^{2}}+(1-t)\frac{k^{2}}{2N^{2}}
=k22N2.\displaystyle=\frac{k^{2}}{2N^{2}}.

It follows that

Re{1w}=Rew|w|2>k22N2×N22k2=14.\displaystyle\text{Re}\,\left\{\frac{1}{w}\right\}=\frac{\text{Re}\,w}{|w|^{2}}>\frac{k^{2}}{2N^{2}}\times\frac{N^{2}}{2k^{2}}=\frac{1}{4}.

Finally, we have the following elementary proposition.

Proposition 2.16.

Let CC be the circle with center at 1/21/2 and radius 1/21/2. Given a point ww on CC, let 𝒲\mathcal{W} be the minor arc joining the point 0 to the point ww. Then (𝒲)\ell(\mathcal{W}), the arc-length of 𝒲\mathcal{W}, is bounded above by π|w|/2.\pi|w|/2. For any zz on 𝒲\mathcal{W}, |z||w||z|\leq|w|.

Proof.
Refer to caption
Figure 10. The circle CC.

Any point zz on CC can be written as

z=12(1+eiθ)\displaystyle z=\frac{1}{2}\left(1+e^{i\theta}\right) (2.14)

for some θ[0,2π]\theta\in[0,2\pi]. Let

w=12(1+eiθw).w=\frac{1}{2}\left(1+e^{i\theta_{w}}\right).

By symmetry of CC, we can assume that θw[0,π)\theta_{w}\in[0,\pi). Then

|w|=12(1+cosθw)2+sin2θw=122+2cosθw=cosθw2.|w|=\frac{1}{2}\sqrt{(1+\cos\theta_{w})^{2}+\sin^{2}\theta_{w}}=\frac{1}{2}\sqrt{2+2\cos\theta_{w}}=\cos\frac{\theta_{w}}{2}.

The arc-length of 𝒲\mathcal{W} is

(𝒲)=12α,\ell(\mathcal{W})=\frac{1}{2}\alpha,

where α=πθw\alpha=\pi-\theta_{w} is the central angle of 𝒲\mathcal{W}. It is elementary to show that if 0<tπ/20<t\leq\pi/2,

2πsintt1.\frac{2}{\pi}\leq\frac{\sin t}{t}\leq 1.

Therefore,

(𝒲)|w|=α2sinα2π2.\displaystyle\frac{\ell(\mathcal{W})}{|w|}=\frac{\alpha}{\displaystyle 2\sin\frac{\alpha}{2}}\leq\frac{\pi}{2}.

This proves the first assertion.

Now, if zz is a point on 𝒲\mathcal{W}, then zz is given by (2.14) for some θ[θw,π]\theta\in[\theta_{w},\pi]. If follows that

|z|=cosθ2cosθw2|w|.|z|=\cos\frac{\theta}{2}\leq\cos\frac{\theta_{w}}{2}\leq|w|.


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 n1n\geq 1, the partition function p(n)p(n) is represented by the convergent series

p(n)=k=1Rk(n),\displaystyle p(n)=\sum_{k=1}^{\infty}R_{k}(n), (3.1)

where

Rk(n)=1π2Ak(n)kddn(sinh{πk23(n124)}n124),\displaystyle R_{k}(n)=\frac{1}{\pi\sqrt{2}}A_{k}(n)\sqrt{k}\frac{d}{dn}\left(\frac{\displaystyle\sinh\left\{\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}\right\}}{\displaystyle\sqrt{n-\frac{1}{24}}}\right),
Ak(n)=1hk(h,k)=1exp(πis(h,k)2πinhk),\displaystyle A_{k}(n)=\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\exp\left(\pi is(h,k)-\frac{2\pi inh}{k}\right),

and s(h,k)s(h,k) is the Dedekind sum defined by

s(h,k)=r=1k1rk(hrkhrk12).\displaystyle s(h,k)=\sum_{r=1}^{k-1}\frac{r}{k}\left(\frac{hr}{k}-\left\lfloor\frac{hr}{k}\right\rfloor-\frac{1}{2}\right). (3.2)

We break the proof into a few lemmas.

Lemma 3.2.

Let

F(z)=m=111zm,\displaystyle F(z)=\prod_{m=1}^{\infty}\frac{1}{1-z^{m}}, (3.3)

and let 𝒞\mathcal{C} be any positively oriented simple closed curve in the unit disc 𝔻\mathbb{D} surrounding the point z=0z=0. Given a positive integer integer nn,

p(n)=12πi𝒞F(z)zn+1𝑑z.\displaystyle p(n)=\frac{1}{2\pi i}\oint_{\mathcal{C}}\frac{F(z)}{z^{n+1}}dz.
Proof.

The infinite product in (3.3) converges absolutely when |z|<1|z|<1. Hence, F(z)F(z) is an analytic function on the unit disc 𝔻\mathbb{D}. Since the Taylor series of F(z)F(z) is given by

F(z)=n=0p(n)zn.\displaystyle F(z)=\sum_{n=0}^{\infty}p(n)z^{n}.

By Cauchy integral formula,

p(n)=12πi𝒞F(z)zn+1𝑑z.\displaystyle p(n)=\frac{1}{2\pi i}\oint_{\mathcal{C}}\frac{F(z)}{z^{n+1}}dz.

As we have mentioned in the proof of Lemma 3.2, the function F(z)F(z) is analytic in the unit disc 𝔻\mathbb{D}. By the product expansion, we see that it has singularities at the points e2πih/ke^{2\pi ih/k}, where hh and kk are positive integers. In the following theorem, we present a transformation formula for F(z)F(z) near a singularity point e2πih/ke^{2\pi ih/k}.

Theorem 3.3.

Let

F(z)=m=111zm,\displaystyle F(z)=\prod_{m=1}^{\infty}\frac{1}{1-z^{m}}, (3.4)

and let hh and kk be integers with 1hk1\leq h\leq k and (h,k)=1(h,k)=1. Let HH be a positive integer not larger than kk such that hH1modkhH\equiv-1\;\text{mod}\;k. Finally, let zz be a complex number with Re(z)>0\text{Re}\,(z)>0. Define

w=exp(2πihk2πzk2),w=exp(2πiHk2πz).w=\exp\left(\frac{2\pi ih}{k}-\frac{2\pi z}{k^{2}}\right),\hskip 28.45274ptw^{\prime}=\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right).

Then

F(w)=eπis(h,k)(zk)12exp(π12zπz12k2)F(w),\displaystyle F(w)=e^{\pi is(h,k)}\left(\frac{z}{k}\right)^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}-\frac{\pi z}{12k^{2}}\right)F(w^{\prime}), (3.5)

where s(h,k)s(h,k) is given by (3.2).

When |z||z| is small, ww is a point that is close to the point e2πih/ke^{2\pi ih/k}, whereas ww^{\prime} is a point that is close to the origin. Since F(0)=1F(0)=1, this theorem says that when ww is close to the singularity point e2πih/ke^{2\pi ih/k},

F(w)eπis(h,k)(zk)12exp(π12z).\displaystyle F(w)\sim e^{\pi is(h,k)}\left(\frac{z}{k}\right)^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}\right).
Proof.

Recall that F(e2πiτ)=eπiτ12η(τ)1\displaystyle F(e^{2\pi i\tau})=e^{\frac{\pi i\tau}{12}}\eta(\tau)^{-1}, where η(τ)\eta(\tau) is the Dedekind eta function (2.1). It follows that

F(w)=eπiτ12η(τ),whereτ=hk+izk2.\displaystyle F(w)=\frac{e^{\frac{\pi i\tau}{12}}}{\eta(\tau)},\hskip 28.45274pt\text{where}\;\;\tau=\frac{h}{k}+i\frac{z}{k^{2}}.

Since

hH=1modk,hH=-1\mod k,

there exists an integer KK such that

kKhH=1.kK-hH=1.

This implies that

[HKkh]\begin{bmatrix}H&-K\\ k&-h\end{bmatrix}

is an element of the modular group Γ=SL(2,)\Gamma=\text{SL}\,(2,\mathbb{Z}). Notice that

τ\displaystyle\tau^{\prime} =HτKkτh\displaystyle=\frac{H\tau-K}{k\tau-h}
=HhKkk+iHzk2izk\displaystyle=\frac{\displaystyle\frac{Hh-Kk}{k}+i\frac{Hz}{k^{2}}}{\displaystyle\frac{iz}{k}}
=Hk+iz.\displaystyle=\frac{H}{k}+\frac{i}{z}.

Hence, w=e2πiτw^{\prime}=e^{2\pi i\tau^{\prime}}. The transformation formula for the Dedekind eta function Theorem 2.2 says that

η(τ)1=η(τ)1{i(kτh)}1/2exp{πi(Hh12k+s(h,k))}.\displaystyle\eta(\tau)^{-1}=\eta(\tau^{\prime})^{-1}\bigl{\{}-i(k\tau-h)\bigr{\}}^{1/2}\exp\left\{\pi i\left(\frac{H-h}{12k}+s(h,k)\right)\right\}.

It follows that

F(w)\displaystyle F(w) =eπiτ12{i(kτh)}1/2exp{πi(Hh12k+s(h,k))}eπiτ12F(w)\displaystyle=e^{\frac{\pi i\tau}{12}}\bigl{\{}-i(k\tau-h)\bigr{\}}^{1/2}\exp\left\{\pi i\left(\frac{H-h}{12k}+s(h,k)\right)\right\}e^{-\frac{\pi i\tau^{\prime}}{12}}F(w^{\prime})
=(zk)12exp{πi(Hh12k+s(h,k)+hH12k+iz12k2i12z)}F(w)\displaystyle=\left(\frac{z}{k}\right)^{\frac{1}{2}}\exp\left\{\pi i\left(\frac{H-h}{12k}+s(h,k)+\frac{h-H}{12k}+\frac{iz}{12k^{2}}-\frac{i}{12z}\right)\right\}F(w^{\prime})
=eπis(h,k)(zk)12exp(π12zπz12k2)F(w).\displaystyle=e^{\pi is(h,k)}\left(\frac{z}{k}\right)^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}-\frac{\pi z}{12k^{2}}\right)F(w^{\prime}).

Proof of Theorem 3.1.

Fixed a positive integer NN. We let 𝒞\mathcal{C} be the image of the path P(N)P(N) defined by (2.10) under the map z=e2πiτ.z=e^{2\pi i\tau}.

By Lemma 3.2,

p(n)=12πi𝒞F(z)zn+1𝑑z.\displaystyle p(n)=\frac{1}{2\pi i}\oint_{\mathcal{C}}\frac{F(z)}{z^{n+1}}dz.

Making the change of variables z=e2πiτz=e^{2\pi i\tau}, we have

p(n)\displaystyle p(n) =P(N)F(e2πiτ)e2πinτ𝑑τ\displaystyle=\int_{P(N)}F(e^{2\pi i\tau})e^{-2\pi in\tau}d\tau
=j=1mNU~jF(e2πiτ)e2πinτ𝑑τ.\displaystyle=\sum_{j=1}^{m_{N}}\int_{\widetilde{U}_{j}}F(e^{2\pi i\tau})e^{-2\pi in\tau}d\tau.

Each U~j\widetilde{U}_{j} is the upper arc of the circle C(hj,kj)C(h_{j},k_{j}) with initial point α1(hj,kj)\alpha_{1}(h_{j},k_{j}) and end point α2(hj,kj)\alpha_{2}(h_{j},k_{j}), where

α1(hj,kj)\displaystyle\alpha_{1}(h_{j},k_{j}) =hjkjkj1kj(kj2+kj12)+ik2+kj12,\displaystyle=\frac{h_{j}}{k_{j}}-\frac{k_{j-1}}{k_{j}(k_{j}^{2}+k_{j-1}^{2})}+\frac{i}{k^{2}+k_{j-1}^{2}},
α2(hj,kj)\displaystyle\alpha_{2}(h_{j},k_{j}) =hjkj+kj+1k(k2+kj+12)+ik2+kj+12.\displaystyle=\frac{h_{j}}{k_{j}}+\frac{k_{j+1}}{k(k^{2}+k_{j+1}^{2})}+\frac{i}{k^{2}+k_{j+1}^{2}}.

For the integral over the arc U~j\widetilde{U}_{j}, we make a change of variables

z=ikj2(τhjkj)z=-ik_{j}^{2}\left(\tau-\frac{h_{j}}{k_{j}}\right)

so that

τ=hjkj+izkj2.\tau=\frac{h_{j}}{k_{j}}+i\frac{z}{k_{j}^{2}}.

Then

U~jF(e2πiτ)e2πinτ𝑑τ\displaystyle\int_{\widetilde{U}_{j}}F(e^{2\pi i\tau})e^{-2\pi in\tau}d\tau =ik2VjF(exp(2πihjkj2πzkj2))e2πinhjkje2πnzkj2𝑑z.\displaystyle=\frac{i}{k^{2}}\int_{V_{j}}F\left(\exp\left(\frac{2\pi ih_{j}}{k_{j}}-\frac{2\pi z}{k_{j}^{2}}\right)\right)e^{-\frac{2\pi inh_{j}}{k_{j}}}e^{\frac{2\pi nz}{k_{j}^{2}}}dz. (3.6)

Here VjV_{j} is the arc of the circle C~\tilde{C} with center at 1/21/2 and radius 1/21/2, which do not touch the imaginary axis, and joining the points w1(hj,kj)w_{1}(h_{j},k_{j}) and w2(hj,kj)w_{2}(h_{j},k_{j}) given by

w1(hj,kj)=kj2kj2+kj12+ikjkj1kj2+kj12,w2(hj,kj)=kj2kj2+kj+12ikjkj+1kj2+kj+12.\begin{split}w_{1}(h_{j},k_{j})&=\frac{k_{j}^{2}}{k_{j}^{2}+k_{j-1}^{2}}+i\frac{k_{j}k_{j-1}}{k_{j}^{2}+k_{j-1}^{2}},\\ w_{2}(h_{j},k_{j})&=\frac{k_{j}^{2}}{k_{j}^{2}+k_{j+1}^{2}}-i\frac{k_{j}k_{j+1}}{k_{j}^{2}+k_{j+1}^{2}}.\end{split} (3.7)

Now we use the transformation formula given in Theorem 3.3,

F(exp(2πihk2πzk2))=eπis(h,k)(zk)12exp(π12zπz12k2)F(exp(2πiHk2πz)),\displaystyle F\left(\exp\left(\frac{2\pi ih}{k}-\frac{2\pi z}{k^{2}}\right)\right)=e^{\pi is(h,k)}\left(\frac{z}{k}\right)^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}-\frac{\pi z}{12k^{2}}\right)F\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right), (3.8)

where HH is a positive integer not larger than kk and such that hH1hH\equiv-1 mod kk.

Let

G(z)=F(z)1=m=1p(m)zm.\displaystyle G(z)=F(z)-1=\sum_{m=1}^{\infty}p(m)z^{m}. (3.9)

Then (3.6) and (3.8) imply that

U~jF(e2πiτ)e2πinτ𝑑τ\displaystyle\int_{\widetilde{U}_{j}}F(e^{2\pi i\tau})e^{-2\pi in\tau}d\tau =𝒜hj,kj+hj,kj,\displaystyle=\mathcal{A}_{h_{j},k_{j}}+\mathcal{B}_{h_{j},k_{j}},

where

𝒜h,k\displaystyle\mathcal{A}_{h,k} =ik5/2Veπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)𝑑z,\displaystyle=\frac{i}{k^{5/2}}\int_{V}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)dz,
h,k\displaystyle\mathcal{B}_{h,k} =ik5/2Veπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)G(exp(2πiHk2πz))𝑑z.\displaystyle=\frac{i}{k^{5/2}}\int_{V}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)G\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right)dz.

As jj runs from 11 to mNm_{N}, (h,k)(h,k) runs through all pairs of positive integers with 1hkN1\leq h\leq k\leq N and (h,k)=1(h,k)=1. Therefore,

p(n)=k=1N1hk(h,k)=1(𝒜h,k+h,k).\displaystyle p(n)=\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\left(\mathcal{A}_{h,k}+\mathcal{B}_{h,k}\right).

We estimate h,k\mathcal{B}_{h,k} first. Given h/kh/k in FNF_{N}, let WW be the chord joining the point w1(h,k)w_{1}(h,k) to the point w2(h,k)w_{2}(h,k). Then VWV-W is a closed curve that does not enclose the origin. Since the integrand

𝒢(z,h,k,n)=ik5/2eπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)G(exp(2πiHk2πz))\displaystyle\mathcal{G}(z,h,k,n)=\frac{i}{k^{5/2}}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)G\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right) (3.10)

is analytic in an open set containing the region enclosed by the closed curve VjWjV_{j}-W_{j}, by residue theorem, we have

VW𝒢(z,h,k,n)𝑑z=0.\oint_{V-W}\mathcal{G}(z,h,k,n)dz=0.

Hence,

h,k\displaystyle\mathcal{B}_{h,k} =W𝒢(z,h,k,n)𝑑z.\displaystyle=\int_{W}\mathcal{G}(z,h,k,n)dz.

For any zWz\in W, Theorem 2.15 says that

|z|2kN+1,\displaystyle|z|\leq\frac{\sqrt{2}k}{N+1}, (3.11)

and the arclength of WW, (W)\ell(W), is not larger than 22k/(N+1)2\sqrt{2}k/(N+1). If we can find an MM such that

|𝒢(z,h,k,n)|Mfor all zW,\displaystyle\left|\mathcal{G}(z,h,k,n)\right|\leq M\hskip 28.45274pt\text{for all }\;z\in W, (3.12)

then

|h,k|M(W)22kN+1M.\displaystyle\left|\mathcal{B}_{h,k}\right|\leq M\ell(W)\leq\frac{2\sqrt{2}k}{N+1}M.

Let us now find an MM satisfying (3.12). By definition (3.10),

|𝒢(z,h,k,n)|=1k5/2|z|12|exp(π12z+2πnzkπz12k2)||G(exp(2πiHk2πz))|.\displaystyle\left|\mathcal{G}(z,h,k,n)\right|=\frac{1}{k^{5/2}}|z|^{\frac{1}{2}}\left|\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k}-\frac{\pi z}{12k^{2}}\right)\right|\left|G\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right)\right|.

The estimate of |z||z| is given by (3.11). Now,

|exp(π12z+2πnzkπz12k2)|\displaystyle\left|\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k}-\frac{\pi z}{12k^{2}}\right)\right|
=exp(Re{π12z+2πnzkπz12k2})\displaystyle=\exp\left(\text{Re}\,\left\{\frac{\pi}{12z}+\frac{2\pi nz}{k}-\frac{\pi z}{12k^{2}}\right\}\right)
=exp(π12Re{1z})exp(Re{2πnzk})exp(Re{πz12k2}).\displaystyle=\exp\left(\frac{\pi}{12}\text{Re}\,\left\{\frac{1}{z}\right\}\right)\exp\left(\text{Re}\,\left\{\frac{2\pi nz}{k}\right\}\right)\exp\left(-\text{Re}\,\left\{\frac{\pi z}{12k^{2}}\right\}\right).

For zWz\in W, Proposition 2.14 implies that Rez>0\text{Re}\,z>0. Hence,

exp(Re{πz12k2})1.\exp\left(-\text{Re}\,\left\{\frac{\pi z}{12k^{2}}\right\}\right)\leq 1.

On the other hand,

Re{2πnzk}2πnk|z|2πnk2πn.\displaystyle\text{Re}\,\left\{\frac{2\pi nz}{k}\right\}\leq\frac{2\pi n}{k}|z|\leq\frac{2\pi n}{k}\leq 2\pi n.

Therefore,

exp(Re{2πnzk})exp(2πn).\exp\left(\text{Re}\,\left\{\frac{2\pi nz}{k}\right\}\right)\leq\exp\left(2\pi n\right).

Next, we estimate

|G(exp(2πiHk2πz))|.\displaystyle\left|G\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right)\right|.

Using definition (3.9) and triangle inequality, we have

|G(exp(2πiHk2πz))|\displaystyle\left|G\left(\exp\left(\frac{2\pi iH}{k}-\frac{2\pi}{z}\right)\right)\right| m=1p(m)|exp(2πimHk2πmz)|\displaystyle\leq\sum_{m=1}^{\infty}p(m)\left|\exp\left(\frac{2\pi imH}{k}-\frac{2\pi m}{z}\right)\right|
=m=1p(m)exp(2πmRe{1z}).\displaystyle=\sum_{m=1}^{\infty}p(m)\exp\left(-2\pi m\text{Re}\,\left\{\frac{1}{z}\right\}\right).

Therefore,

|𝒢(z,h,k,n)|21/4k2N+1exp(2πn)m=1p(m)exp(2π[m124]Re{1z}).\displaystyle\left|\mathcal{G}(z,h,k,n)\right|\leq\frac{2^{1/4}}{k^{2}\sqrt{N+1}}\exp\left(2\pi n\right)\sum_{m=1}^{\infty}p(m)\exp\left(-2\pi\left[m-\frac{1}{24}\right]\text{Re}\,\left\{\frac{1}{z}\right\}\right).

For a point zWz\in W, Theorem 2.15 says that

x=Re{1z}14.x=\text{Re}\,\left\{\frac{1}{z}\right\}\geq\frac{1}{4}.

Using the fact that 0<p(m1)<p(m2)0<p(m_{1})<p(m_{2}) when m1<m2m_{1}<m_{2}, we find that

0\displaystyle 0 m=1p(m)exp(2π[m124]Re{1z})\displaystyle\leq\sum_{m=1}^{\infty}p(m)\exp\left(-2\pi\left[m-\frac{1}{24}\right]\text{Re}\,\left\{\frac{1}{z}\right\}\right)
=m=1p(24m1)exp(πx12(24m1))\displaystyle=\sum_{m=1}^{\infty}p(24m-1)\exp\left(-\frac{\pi x}{12}(24m-1)\right)
k=1p(k)exp(πkx12)\displaystyle\leq\sum_{k=1}^{\infty}p(k)\exp\left(-\frac{\pi kx}{12}\right)
=F(exp(πx12))1.\displaystyle=F\left(\exp\left(-\frac{\pi x}{12}\right)\right)-1.

Since all the coefficients of F(z)F(z) are positive, we find that 0<F(a1)<F(a2)0<F(a_{1})<F(a_{2}) if 0<a1<a20<a_{1}<a_{2}. Hence,

F(exp(πx12))1F(eπ48)1=𝒞0.F\left(\exp\left(-\frac{\pi x}{12}\right)\right)-1\leq F\left(e^{-\frac{\pi}{48}}\right)-1=\mathscr{C}_{0}.

It follows that

|𝒢(z,h,k,n)|𝒞021/4k2N+1exp(2πn).\displaystyle\left|\mathcal{G}(z,h,k,n)\right|\leq\mathscr{C}_{0}\frac{2^{1/4}}{k^{2}\sqrt{N+1}}\exp\left(2\pi n\right).

Hence,

|k=1N1hk(h,k)=1h,k|\displaystyle\left|\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\mathcal{B}_{h,k}\right| 𝒞0k=1N1hk(h,k)=122kN+1×21/4k2N+1exp(2πn)\displaystyle\leq\mathscr{C}_{0}\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\frac{2\sqrt{2}k}{N+1}\times\frac{2^{1/4}}{k^{2}\sqrt{N+1}}\exp\left(2\pi n\right)
𝒞1(N+1)3/2k=1N1hk(h,k)=11k,\displaystyle\leq\frac{\mathscr{C}_{1}}{(N+1)^{3/2}}\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\frac{1}{k},

where

𝒞1=274𝒞0exp(2πn)\mathscr{C}_{1}=2^{\frac{7}{4}}\mathscr{C}_{0}\exp\left(2\pi n\right)

is a fixed constant that only depends on nn. It is easy to see that

0\displaystyle 0 k=1N1hk(h,k)=11kk=1N1=N,\displaystyle\leq\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\frac{1}{k}\leq\sum_{k=1}^{N}1=N,

since the number of positive integer hh such that 1hk1\leq h\leq k and (h,k)=1(h,k)=1 is not larger than kk.

Hence, we have shown that

|k=1N1hk(h,k)=1h,k|𝒞1N+1,\left|\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\mathcal{B}_{h,k}\right|\leq\frac{\mathscr{C}_{1}}{\sqrt{N+1}},

which is of order N1/2N^{-1/2}. It follows that

p(n)=k=1N1hk(h,k)=1𝒜h,k+O(𝒞1N1/2).\displaystyle p(n)=\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\ \end{subarray}}\mathcal{A}_{h,k}+O^{*}(\mathscr{C}_{1}N^{-1/2}).

Here we write

f(N)=O(g(N))f(N)=O^{*}(g(N))

to mean

|f(N)||g(N)|.|f(N)|\leq|g(N)|.

To compute 𝒜h,k\mathcal{A}_{h,k}, we notice that when zz traverses the circle C~\tilde{C} from w1(h,k)w_{1}(h,k) to w2(h,k)w_{2}(h,k) along the arc that is away from the imaginary axis, it traverses clockwise. We can write this arc as the whole circle C~-\tilde{C} (the minus sign is for the clockwise orientation) minus the two arcs 𝒲1-\mathcal{W}_{1} and 𝒲2\mathcal{W}_{2}, where 𝒲1\mathcal{W}_{1} is the arc from 0 to w1(h,k)w_{1}(h,k) which is clockwise, and 𝒲2\mathcal{W}_{2} is the arc 0 to w2(h,k)w_{2}(h,k) which is anticlockwise. This gives,

𝒜h,k\displaystyle\mathcal{A}_{h,k} =ik5/2C~eπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)𝑑z\displaystyle=-\frac{i}{k^{5/2}}\int_{\tilde{C}}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)dz
ik5/2𝒲1eπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)𝑑z\displaystyle-\frac{i}{k^{5/2}}\int_{\mathcal{W}_{1}}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)dz
+ik5/2𝒲2eπis(h,k)2πinhkz12exp(π12z+2πnzk2πz12k2)𝑑z\displaystyle+\frac{i}{k^{5/2}}\int_{\mathcal{W}_{2}}e^{\pi is(h,k)-\frac{2\pi inh}{k}}z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)dz
=𝒜0(h,k)+𝒜1(h,k)+𝒜2(h,k).\displaystyle=\mathcal{A}_{0}(h,k)+\mathcal{A}_{1}(h,k)+\mathcal{A}_{2}(h,k).

For any z𝒲1𝒲2z\in\mathcal{W}_{1}\cup\mathcal{W}_{2}, Proposition 2.14 and Proposition 2.16 imply that

|z12exp(π12z+2πnzk2πz12k2)|\displaystyle\left|z^{\frac{1}{2}}\exp\left(\frac{\pi}{12z}+\frac{2\pi nz}{k^{2}}-\frac{\pi z}{12k^{2}}\right)\right| 21/4k1/2N+1exp(π12Re{1z}+2πnk2Rezπ12k2Rez)\displaystyle\leq\frac{2^{1/4}k^{1/2}}{\sqrt{N+1}}\exp\left(\frac{\pi}{12}\text{Re}\,\left\{\frac{1}{z}\right\}+\frac{2\pi n}{k^{2}}\text{Re}\,z-\frac{\pi}{12k^{2}}\text{Re}\,z\right)
21/4k1/2N+1exp(π12+2πn).\displaystyle\leq\frac{2^{1/4}k^{1/2}}{\sqrt{N+1}}\exp\left(\frac{\pi}{12}+2\pi n\right).

Theorem 2.15 and Proposition 2.16 say that (𝒲1)\ell(\mathcal{W}_{1}) and (𝒲2)\ell(\mathcal{W}_{2}) are bounded above by

π2×2kN+1.\frac{\pi}{2}\times\frac{\sqrt{2}k}{N+1}.

It follows that

|k=1N1hk(h,k)=1(𝒜1(h,k)+𝒜2(h,k))|\displaystyle\left|\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\left(\mathcal{A}_{1}(h,k)+\mathcal{A}_{2}(h,k)\right)\right|
k=1N1hk(h,k)=12×π2×2kN+1×1k5/2×21/4k1/2N+1exp(π12+2πn)\displaystyle\leq\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}2\times\frac{\pi}{2}\times\frac{\sqrt{2}k}{N+1}\times\frac{1}{k^{5/2}}\times\frac{2^{1/4}k^{1/2}}{\sqrt{N+1}}\exp\left(\frac{\pi}{12}+2\pi n\right)
𝒞2N+1,\displaystyle\leq\frac{\mathscr{C}_{2}}{\sqrt{N+1}},

where

𝒞2=23/4πexp(π12+2πn).\mathscr{C}_{2}=2^{3/4}\pi\exp\left(\frac{\pi}{12}+2\pi n\right).

This shows that

p(n)=k=1N1hk(h,k)=1𝒜0(h,k)+O((𝒞1+𝒞2)N1/2).\displaystyle p(n)=\sum_{k=1}^{N}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\mathcal{A}_{0}(h,k)+O^{*}((\mathscr{C}_{1}+\mathscr{C}_{2})N^{-1/2}). (3.13)

Taking NN\rightarrow\infty limit, we find that

p(n)\displaystyle p(n) =k=11hk(h,k)=1𝒜0(h,k)\displaystyle=\sum_{k=1}^{\infty}\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}\mathcal{A}_{0}(h,k)
=ik=11k5/2Ak(n)C~z12exp{π12z+2πzk2(n124)}𝑑z\displaystyle=-i\sum_{k=1}^{\infty}\frac{1}{k^{5/2}}A_{k}(n)\int_{\tilde{C}}z^{\frac{1}{2}}\exp\left\{\frac{\pi}{12z}+\frac{2\pi z}{k^{2}}\left(n-\frac{1}{24}\right)\right\}dz

where

Ak(n)=1hk(h,k)=1eπis(h,k)2πinhk.A_{k}(n)=\sum_{\begin{subarray}{c}1\leq h\leq k\\ (h,k)=1\end{subarray}}e^{\pi is(h,k)-\frac{2\pi inh}{k}}.

In particular, A1(n)=1A_{1}(n)=1. Making a change of variables

w=1z,w=\frac{1}{z},

the circle C~\tilde{C} is mapped to the line Rew=1\text{Re}\,w=1, with Imw\text{Im}\,w goes from \infty to -\infty. Hence,

p(n)=ik=11k5/2Ak(n)Rew=1w52exp{πw12+2πk2(n124)1w}𝑑w.\displaystyle p(n)=-i\sum_{k=1}^{\infty}\frac{1}{k^{5/2}}A_{k}(n)\int_{\text{Re}\,w=1}w^{-\frac{5}{2}}\exp\left\{\frac{\pi w}{12}+\frac{2\pi}{k^{2}}\left(n-\frac{1}{24}\right)\frac{1}{w}\right\}dw.

Making a change of variables

t=πw12,t=\frac{\pi w}{12},

we find that

p(n)=iπ32243k=11k5/2Ak(n)Ret=π12t52exp(t+xk24t)𝑑t,\displaystyle p(n)=-\frac{i\pi^{\frac{3}{2}}}{24\sqrt{3}}\sum_{k=1}^{\infty}\frac{1}{k^{5/2}}A_{k}(n)\int_{\text{Re}\,t=\frac{\pi}{12}}t^{-\frac{5}{2}}\exp\left(t+\frac{x_{k}^{2}}{4t}\right)dt,

where

xk\displaystyle x_{k} =πk23(n124).\displaystyle=\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}.

By (2.6), we find that

p(n)=k=1Rk(n),\displaystyle p(n)=\sum_{k=1}^{\infty}R_{k}(n),

where

Rk(n)=π521231k5/2Ak(n)(xk2)32I32(xk).\displaystyle R_{k}(n)=\frac{\pi^{\frac{5}{2}}}{12\sqrt{3}}\frac{1}{k^{5/2}}A_{k}(n)\left(\frac{x_{k}}{2}\right)^{-\frac{3}{2}}I_{\frac{3}{2}}(x_{k}).

By Proposition 2.5,

Rk(n)\displaystyle R_{k}(n) =π521231k5/2Ak(n)(xk2)322xkπ1dxkdnddn(sinh{πk23(n124)}πk23(n124))\displaystyle=\frac{\pi^{\frac{5}{2}}}{12\sqrt{3}}\frac{1}{k^{5/2}}A_{k}(n)\left(\frac{x_{k}}{2}\right)^{-\frac{3}{2}}\sqrt{\frac{2x_{k}}{\pi}}\frac{1}{\displaystyle\frac{dx_{k}}{dn}}\frac{d}{dn}\left(\frac{\displaystyle\sinh\left\{\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}\right\}}{\displaystyle\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}}\right)
=1π2Ak(n)kddn(sinh{πk23(n124)}n124).\displaystyle=\frac{1}{\pi\sqrt{2}}A_{k}(n)\sqrt{k}\frac{d}{dn}\left(\frac{\displaystyle\sinh\left\{\frac{\pi}{k}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}\right\}}{\displaystyle\sqrt{n-\frac{1}{24}}}\right).

Notice that (3.13) shows that the series k=1Rk(n)\displaystyle\sum_{k=1}^{\infty}R_{k}(n) indeed converges to p(n)p(n). This completes the proof of the theorem.

Remark 3.4.

Fixed a positive integer nn. Given a positive integer NN, define

QN(n)=k=N+1Rk(n),Q_{N}(n)=\sum_{k=N+1}^{\infty}R_{k}(n),

so that

p(n)=k=1NRk(n)+QN(n).p(n)=\sum_{k=1}^{N}R_{k}(n)+Q_{N}(n).

Namely, QN(n)Q_{N}(n) is the remainder if we approximate p(n)p(n) by the sum of the first NN terms in the series (3.1).

By (3.13)

|QN(n)|𝒞N,|Q_{N}(n)|\leq\frac{\mathscr{C}}{\sqrt{N}},

where

𝒞=27/4(F(eπ48)1)e2πn+23/4πexp(π12+2πn).\displaystyle\mathscr{C}=2^{7/4}\left(F(e^{-\frac{\pi}{48}})-1\right)e^{2\pi n}+2^{3/4}\pi\exp\left(\frac{\pi}{12}+2\pi n\right).

Since QN(n)0Q_{N}(n)\rightarrow 0 as NN\rightarrow\infty, the series (3.1) converges absolutely.

4. Leading Asymptotic of the Partition Function

In the following, we use the exact formula for p(n)p(n) we obtained in the previous section to find the leading asymptotic of p(n)p(n) as a function of nn when nn\rightarrow\infty.

Theorem 4.1.

The sequence {p(n)}\{p(n)\} satisfies the following asymptotic formula when nn\rightarrow\infty.

limnp(n)14n3exp(π2n3)=1.\lim_{n\rightarrow\infty}\frac{p(n)}{\displaystyle\frac{1}{4n\sqrt{3}}\exp\left(\pi\sqrt{\frac{2n}{3}}\right)}=1. (4.1)

Given a positive integer nn, let

L(n)=14n3exp(π2n3).L(n)=\frac{1}{4n\sqrt{3}}\exp\left(\pi\sqrt{\frac{2n}{3}}\right).

Eq. (4.1) asserts that

limnp(n)L(n)=1.\lim_{n\rightarrow\infty}\frac{p(n)}{L(n)}=1.

To prove this, let

S(n)=k=2Rk(n).S(n)=\sum_{k=2}^{\infty}R_{k}(n).

Then by (3.1),

p(n)=R1(n)+S(n).p(n)=R_{1}(n)+S(n).

To prove (4.1), it is sufficient to show that

limnR1(n)L(n)=1\lim_{n\rightarrow\infty}\frac{R_{1}(n)}{L(n)}=1 (4.2)

and

limnS(n)L(n)=0.\lim_{n\rightarrow\infty}\frac{S(n)}{L(n)}=0. (4.3)

By a straightforward computation, we find that

Rk(n)=πk32n124Ak(n)α(n)kcoshα(n)ksinhα(n)kα(n)2,R_{k}(n)=\frac{\pi\sqrt{k}}{3\sqrt{2}\displaystyle\sqrt{n-\frac{1}{24}}}A_{k}(n)\frac{\displaystyle\frac{\alpha(n)}{k}\cosh\frac{\alpha(n)}{k}-\sinh\frac{\alpha(n)}{k}}{\alpha(n)^{2}}, (4.4)

where

α(n)=π23(n124).\alpha(n)=\pi\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}.

We first prove (4.2).

Lemma 4.2.

Let nn be a positive integer. Then

limnR1(n)L(n)=1\lim_{n\rightarrow\infty}\frac{R_{1}(n)}{L(n)}=1
Proof.

By (4.4),

R1(n)=143(n124){(11α(n))eα(n)+(1+1α(n))eα(n)}.R_{1}(n)=\frac{1}{4\sqrt{3}\displaystyle\left(n-\frac{1}{24}\right)}\left\{\left(1-\frac{1}{\alpha(n)}\right)e^{\alpha(n)}+\left(1+\frac{1}{\alpha(n)}\right)e^{-\alpha(n)}\right\}.

Since

limnα(n)=,\lim_{n\rightarrow\infty}\alpha(n)=\infty,

we find that

limnR1(n)143(n124)eα(n)=1.\lim_{n\rightarrow\infty}\frac{R_{1}(n)}{\displaystyle\frac{1}{4\sqrt{3}\displaystyle\left(n-\frac{1}{24}\right)}e^{\alpha(n)}}=1.

It follows that

limnR1(n)L(n)\displaystyle\lim_{n\rightarrow\infty}\frac{R_{1}(n)}{L(n)} =limn143(n124)eα(n)14n3exp(π2n3)\displaystyle=\lim_{n\rightarrow\infty}\frac{\displaystyle\frac{1}{4\sqrt{3}\displaystyle\left(n-\frac{1}{24}\right)}e^{\alpha(n)}}{\displaystyle\frac{1}{4n\sqrt{3}}\exp\left(\pi\sqrt{\frac{2n}{3}}\right)}
=limnexp(π23(n124)π2n3).\displaystyle=\lim_{n\rightarrow\infty}\exp\left(\pi\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}-\pi\sqrt{\frac{2n}{3}}\right).

Since

limn(π23(n124)π2n3)\displaystyle\lim_{n\rightarrow\infty}\left(\pi\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}-\pi\sqrt{\frac{2n}{3}}\right)
=π23limnn124nn124+n\displaystyle=\pi\sqrt{\frac{2}{3}}\lim_{n\rightarrow\infty}\frac{\displaystyle n-\frac{1}{24}-n}{\displaystyle\sqrt{n-\frac{1}{24}}+\sqrt{n}}
=0,\displaystyle=0,

we conclude that

limnR1(n)L(n)=1.\lim_{n\rightarrow\infty}\frac{R_{1}(n)}{L(n)}=1.

To prove (4.3), we need the following.

Lemma 4.3.

For u>0u>0,

|ucoshusinhuu2|ucoshu2.\left|\frac{u\cosh u-\sinh u}{u^{2}}\right|\leq\frac{u\cosh u}{2}. (4.5)
Proof.

Let

g(u)=ucoshusinhu,h(u)=u2.g(u)=u\cosh u-\sinh u,\hskip 28.45274pth(u)=u^{2}.

By Cauchy mean value theorem, there exists v(0,u)v\in(0,u) such that

ucoshusinhuu2=g(u)h(u)=g(v)h(v).\frac{u\cosh u-\sinh u}{u^{2}}=\frac{g(u)}{h(u)}=\frac{g^{\prime}(v)}{h^{\prime}(v)}.

Now,

g(u)=usinhu,h(u)=2u.\displaystyle g^{\prime}(u)=u\sinh u,\quad h^{\prime}(u)=2u.

Thus,

ucoshusinhuu2\displaystyle\frac{u\cosh u-\sinh u}{u^{2}} =sinhv2\displaystyle=\frac{\sinh v}{2}

By mean value theorem, there exists w(0,v)w\in(0,v) such that

sinhvv=coshw.\frac{\sinh v}{v}=\cosh w.

Since coshu\cosh u is a strictly increasing function, coshwcoshvcoshu\cosh w\leq\cosh v\leq\cosh u. Thus,

ucoshusinhuu2\displaystyle\frac{u\cosh u-\sinh u}{u^{2}} =sinhv2\displaystyle=\frac{\sinh v}{2}
vcoshw2\displaystyle\leq\frac{v\cosh w}{2}
ucoshu2.\displaystyle\leq\frac{u\cosh u}{2}.

The assertion follows. ∎

Proof of Theorem 4.1.

To complete the proof of Theorem 4.1, it remains to prove (4.3).

By (4.5) and (4.4), and using the fact that |Ak(n)|k|A_{k}(n)|\leq k, we have

|Rk(n)|πα(n)62n1241k3/2coshα(n)k.|R_{k}(n)|\leq\frac{\pi\alpha(n)}{6\sqrt{2}\displaystyle\sqrt{n-\frac{1}{24}}}\frac{1}{k^{3/2}}\cosh\frac{\alpha(n)}{k}.

Using the fact that coshu<eu\cosh u<e^{u} when u>0u>0, we have

|Rk(n)|π2631k3/2exp(α(n)k).|R_{k}(n)|\leq\frac{\pi^{2}}{6\sqrt{3}}\frac{1}{k^{3/2}}\exp\left(\frac{\alpha(n)}{k}\right).

Hence,

|S(n)|\displaystyle|S(n)| k=2|Rk(n)|\displaystyle\leq\sum_{k=2}^{\infty}|R_{k}(n)|
k=2π2631k3/2exp(α(n)k)\displaystyle\leq\sum_{k=2}^{\infty}\frac{\pi^{2}}{6\sqrt{3}}\frac{1}{k^{3/2}}\exp\left(\frac{\alpha(n)}{k}\right)
π263exp(α(n)2)k=21k3/2\displaystyle\leq\frac{\pi^{2}}{6\sqrt{3}}\exp\left(\frac{\alpha(n)}{2}\right)\sum_{k=2}^{\infty}\frac{1}{k^{3/2}}
=Cπ263exp(α(n)2),\displaystyle=C\frac{\pi^{2}}{6\sqrt{3}}\exp\left(\frac{\alpha(n)}{2}\right),

where

C=ζ(32)1.C=\zeta\left(\frac{3}{2}\right)-1.

It follows that

|S(n)L(n)|\displaystyle\left|\frac{S(n)}{L(n)}\right| 2Cπ2n3exp(π223(n124)π2n3)\displaystyle\leq\frac{2C\pi^{2}n}{3}\exp\left(\frac{\pi}{2}\sqrt{\frac{2}{3}\left(n-\frac{1}{24}\right)}-\pi\sqrt{\frac{2n}{3}}\right)
2Cπ2n3exp(π223nπ2n3)\displaystyle\leq\frac{2C\pi^{2}n}{3}\exp\left(\frac{\pi}{2}\sqrt{\frac{2}{3}n}-\pi\sqrt{\frac{2n}{3}}\right)
=2Cπ2n3exp(π223n).\displaystyle=\frac{2C\pi^{2}n}{3}\exp\left(-\frac{\pi}{2}\sqrt{\frac{2}{3}n}\right).

It is easy to verify that

limn2Cπ2n3exp(π223n)=2Cπ23limnnexp(π223n)=0.\lim_{n\rightarrow\infty}\frac{2C\pi^{2}n}{3}\exp\left(-\frac{\pi}{2}\sqrt{\frac{2}{3}n}\right)=\frac{2C\pi^{2}}{3}\lim_{n\rightarrow\infty}\frac{n}{\displaystyle\exp\left(\frac{\pi}{2}\sqrt{\frac{2}{3}n}\right)}=0.

Hence,

limnS(n)L(n)=0.\lim_{n\rightarrow\infty}\frac{S(n)}{L(n)}=0.

This completes the proof of (4.1).

Theorem 4.1 states that the ratio of p(n)p(n) to L(n)L(n) approaches 1 when nn approaches infinity. This does not mean when we use L(n)L(n) to approximate p(n)p(n), the error in the approximation, defined as p(n)L(n)p(n)-L(n), is small. In fact, this error also tends to infinity when nn approaches infinity. What we assert is that the relative error

e(n)=p(n)L(n)p(n)e(n)=\frac{p(n)-L(n)}{p(n)}

goes to 0. The percentage relative error is defined as

ε(n)=p(n)L(n)p(n)×100%.\varepsilon(n)=\frac{p(n)-L(n)}{p(n)}\times 100\%.

Table 1 shows some values of p(n)p(n) as well as the percentage relative errors in the approximation of p(n)p(n) by L(n)L(n). Figure 11 depicts the percentage relative errors graphically.

Refer to caption
Figure 11. The percentage relative error in using L(n)L(n) to approximate p(n)p(n).
nn p(n)p(n)
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 p(n)p(n).
nn p(n)p(n) percentage relative error ε(n)\varepsilon(n)
10 4242 14.53-14.53
50 204226204226 6.54-6.54
100 190569292190569292 4.57-4.57
200 3.97×10123.97\times 10^{12} 3.2-3.2
500 2.30×10212.30\times 10^{21} 2.01-2.01
1000 2.41×10312.41\times 10^{31} 1.42-1.42
2000 4.72×10454.72\times 10^{45} 1-1
3000 4.96×10564.96\times 10^{56} 0.81-0.81
4000 1.02×10661.02\times 10^{66} 0.7-0.7
5000 1.70×10741.70\times 10^{74} 0.63-0.63
6000 4.67×10814.67\times 10^{81} 0.57-0.57
7000 3.29×10883.29\times 10^{88} 0.53-0.53
8000 7.84×10947.84\times 10^{94} 0.5-0.5
9000 7.71×101007.71\times 10^{100} 0.47-0.47
10000 3.62×101063.62\times 10^{106} 0.44-0.44
12000 1.29×101171.29\times 10^{117} 0.41-0.41
15000\quad 15000\quad 2.63×10131\quad 2.63\times 10^{131}\quad 0.36-0.36
Table 1. The values of p(n)p(n) and the percentage relative errors in using L(n)L(n) to approximate p(n)p(n).

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 p(n)p(n), 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 p(n)p(n), 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 p3p\geq 3, 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 P>3P>3, Amer. J. Math. 67 (1945), 194–208. MR 12101
  • [MLP12] James Mc Laughlin and Scott Parsell, A Hardy-Ramanujan-Rademacher-type formula for (r,s)(r,s)-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 p(n)p(n), 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 nn into at most mm 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