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

Limiting Betti distributions of Hilbert schemes on nn points

Michael Griffin, Ken Ono, Larry Rolen, and Wei-Lun Tsai Department of Mathematics, Brigham Young University, Provo, UT 84602 [email protected] Department of Mathematics, University of Virginia, Charlottesville, VA 22904 [email protected] Department of Mathematics, Vanderbilt University, Nashville, TN 37240 [email protected] Department of Mathematics, University of Virginia, Charlottesville, VA 22904 [email protected]
Abstract.

Hausel and Rodriguez-Villegas [7] recently observed that work of Göttsche, combined with a classical result of Erdős and Lehner on integer partitions, implies that the limiting Betti distribution for the Hilbert schemes (2)[n](\mathbb{C}^{2})^{[n]} on nn points, as n+,n\rightarrow+\infty, is a Gumbel distribution. In view of this example, they ask for further such Betti distributions. We answer this question for the quasihomogeneous Hilbert schemes ((2)[n])Tα,β((\mathbb{C}^{2})^{[n]})^{T_{\alpha,\beta}} that are cut out by torus actions. We prove that their limiting distributions are also of Gumbel type. To obtain this result, we combine work of Buryak, Feigin, and Nakajima on these Hilbert schemes with our generalization of the result of Erdős and Lehner, which gives the distribution of the number of parts in partitions that are multiples of a fixed integer A2.A\geq 2. Furthermore, if pk(A;n)p_{k}(A;n) denotes the number of partitions of nn with exactly kk parts that are multiples of AA, then we obtain the asymptotic

pk(A,n)24k214(nAk)k2342(11A)k214k!Ak+12(2π)ke2π16(11A)(nAk),p_{k}(A,n)\sim\frac{24^{\frac{k}{2}-\frac{1}{4}}(n-Ak)^{\frac{k}{2}-\frac{3}{4}}}{\sqrt{2}\left(1-\frac{1}{A}\right)^{\frac{k}{2}-\frac{1}{4}}k!A^{k+\frac{1}{2}}(2\pi)^{k}}e^{2\pi\sqrt{\frac{1}{6}\left(1-\frac{1}{A}\right)(n-Ak)}},

a result which is of independent interest.

Key words and phrases:
Betti numbers, Hilbert schemes, partitions
2020 Mathematics Subject Classification:
14C05, 14F99, 11P82

1. Introduction and statement of results

We consider the Hilbert schemes of nn points on 2,\mathbb{C}^{2}, denoted X[n]=(2)[n],X^{[n]}=(\mathbb{C}^{2})^{[n]}, that have been studied by Göttsche [9, 10], and Buryak, Feigin, and Nakajima [2, 3]. Each X[n]X^{[n]} is a nonsingular, irreducible, quasiprojective dimension 2n2n algebraic variety. Moreover, they enjoy the convenient description

X[n]={I[x,y]:I is an ideal with dim([x,y]/I)=n},X^{[n]}=\left\{I\subset\mathbb{C}[x,y]\ :\ {\text{\rm$I$ is an ideal with $\dim_{\mathbb{C}}(\mathbb{C}[x,y]/I)=n$}}\right\}, (1.1)

which reduces the calculation of its Betti numbers to problems on integer partitions. To investigate these Betti numbers, it is natural to combine them to form the generating function

P(X[n];T):=j=02n2bj(n)Tj=j=02n2dim(Hj(X[n],))Tj,P\left(X^{[n]};T\right):=\sum_{j=0}^{2n-2}b_{j}(n)T^{j}=\sum_{j=0}^{2n-2}\dim\left(H_{j}\left(X^{[n]},\mathbb{Q}\right)\right)T^{j}, (1.2)

known as its Poincaré polynomial. Due to the connection with integer partitions, it turns out that these polynomial generating functions equivalently keep track of the number of parts among the size nn partitions.

In their work on the statistical properties of certain varieties, Hausel and Rodriguez-Villegas [7] observed that a classical result of Erdős and Lehner on partitions [4] gives (see Section 4.3 of [7]) the limiting distribution for the Betti numbers of X[n]X^{[n]} as n+n\rightarrow+\infty. Using Göttsche’s generating function [9, 10] for the P(X[n];T),P(X^{[n]};T), it is straightforward to compute examples that offer glimpses of this result. For example, we find that

P(X[50];T)=1+T2+2T4++5427T88+2611T90+920T92+208T94+25T96+T98.P\left(X^{[50]};T\right)=1+T^{2}+2T^{4}+\dots+5427T^{88}+2611T^{90}+920T^{92}+208T^{94}+25T^{96}+T^{98}.

The renormalized even degree111The coefficients b2j+1(n)b_{2j+1}(n) for odd degree terms identically vanish. coefficients are plotted in Figure 1. As P(X[50];1)=p(50),P\left(X^{[50]};1\right)=p(50), the number of partitions of 50,50, the plot consists of the points {(2m98,b2m(50)p(50)): 0m49}.\left\{\left(\frac{2m}{98},\frac{b_{2m}(50)}{p(50)}\right)\ :\ 0\leq m\leq 49\right\}.

[Uncaptioned image]
Figure 1. Betti distribution for X[50]X^{[50]}

These distributions, when properly renormalized, converge to a Gumbel distribution as n+.n\rightarrow+\infty.

Hausel and Rodriguez-Villegas asked for further such nn-aspect Betti distributions. We answer this question for the quasihomogeneous nn point Hilbert schemes that are cut out by torus actions. To define them, we use the torus (×)2(\mathbb{C}^{\times})^{2}-action on 2\mathbb{C}^{2} defined by scalar multiplication

(t1,t2)(x,y):=(t1x,t2y),(t_{1},t_{2})\cdot(x,y):=(t_{1}x,t_{2}y),

which lifts to X[n]=(2)[n].X^{[n]}=(\mathbb{C}^{2})^{[n]}. For relatively prime α,β,\alpha,\beta\in\mathbb{N}, we have the one-dimensional subtorus

Tα,β:={(tα,tβ):t×}.T_{\alpha,\beta}:=\{(t^{\alpha},t^{\beta})\ :\ t\in\mathbb{C}^{\times}\}.

The quasihomogeneous Hilbert scheme Xα,β[n]:=((2)[n])Tα,βX^{[n]}_{\alpha,\beta}:=((\mathbb{C}^{2})^{[n]})^{T_{\alpha,\beta}} is the fixed point set of X[n].X^{[n]}.

To define Betti distributions, we make use of the Poincaré polynomials

P(Xα,β[n];T):=j=02nα+βbj(α,β;n)Tj=j=02nα+βdim(Hj(Xα,β[n],))Tj.P\left(X^{[n]}_{\alpha,\beta};T\right):=\sum_{j=0}^{2\lfloor\frac{n}{\alpha+\beta}\rfloor}b_{j}(\alpha,\beta;n)T^{j}=\sum_{j=0}^{2\lfloor\frac{n}{\alpha+\beta}\rfloor}\dim\left(H_{j}\left(X^{[n]}_{\alpha,\beta},\mathbb{Q}\right)\right)T^{j}. (1.3)

As P(Xα,β[n];1)=p(n),P\left(X^{[n]}_{\alpha,\beta};1\right)=p(n), we have that the discrete measure dμα,β[n]d\mu^{[n]}_{\alpha,\beta} for Xα,β[n]X^{[n]}_{\alpha,\beta} is

Φn(α,β;x):=1p(n)x𝑑μα,β[n]=1p(n)jxbj(α,β;n).\Phi_{n}(\alpha,\beta;x):=\frac{1}{p(n)}\cdot\int_{-\infty}^{x}d\mu^{[n]}_{\alpha,\beta}=\frac{1}{p(n)}\cdot\sum_{j\leq x}b_{j}(\alpha,\beta;n). (1.4)

The following theorem gives the limiting Betti distributions (as functions in xx) we seek.

Theorem 1.1.

If α\alpha and β\beta are relatively prime positive integers, then

limn+Φn(α,β;2nx+δn(α,β))=exp(6π(α+β)exp(π(α+β)6x)),\lim_{n\rightarrow+\infty}\Phi_{n}(\alpha,\beta;2\sqrt{n}x+\delta_{n}(\alpha,\beta))=\exp\left(-\frac{\sqrt{6}}{\pi(\alpha+\beta)}\cdot\exp\left(-\frac{\pi(\alpha+\beta)}{\sqrt{6}}x\right)\right),

where δn(α,β):=6π(α+β)nlog(n).\delta_{n}(\alpha,\beta):=\frac{\sqrt{6}}{\pi(\alpha+\beta)}\sqrt{n}\log(n).

Two Remarks.

(1) The limiting cumulative distribution in Theorem 1.1 is of Gumbel type [5, 6]. Such distributions are often used to study the maximum (resp. minimum) of a number of samples of a random variable. Letting A:=α+β,A:=\alpha+\beta, we have mean 6Aπ(log(6Aπ)+γ),\frac{\sqrt{6}}{A\pi}\left(\log\left(\frac{\sqrt{6}}{A\pi}\right)+\gamma\right), where γ\gamma is the Euler-Mascheroni constant, and variance 1/A2.1/A^{2}.

(2) Gillman, Gonzalez, Schoenbauer and two of the authors studied a different kind of distribution for Hilbert schemes of surfaces in [8]. In that work equidistribution results were obtained for the Hodge numbers organized by congruence conditions.

Example.

For example, let α=1\alpha=1 and β=2.\beta=2. For n=20,n=20, we have

P(X1,2[20];T)=202+212T2+126T4+56T6+22T8+7T10+2T12.P\left(X^{[20]}_{1,2};T\right)=202+212T^{2}+126T^{4}+56T^{6}+22T^{8}+7T^{10}+2T^{12}.

This small degree polynomial is not very suggestive. However, for n=1000n=1000 the renormalized even degree222The odd degree coefficients terms identically vanish. coefficients displayed in Figure 2 is quite illuminating. As P(X1,2[1000];1)=p(1000),P\left(X^{[1000]}_{1,2};1\right)=p(1000), the plot consists of the 334 points {(2m666,b2m(1000)p(1000)): 0m333}.\left\{\left(\frac{2m}{666},\frac{b_{2m}(1000)}{p(1000)}\right)\ :\ 0\leq m\leq 333\right\}.

[Uncaptioned image]
Figure 2. Betti distribution for X1,2[1000]X^{[1000]}_{1,2}

Theorem 1.1 gives the cumulative distribution corresponding to such plots as n+.n\rightarrow+\infty. In this case, the theorem asserts that

limn+Φn(1,2;2nx+6n3πlog(n))=exp(63πexp(3πx6)).\lim_{n\rightarrow+\infty}\Phi_{n}\left(1,2;2\sqrt{n}x+\frac{\sqrt{6n}}{3\pi}\cdot\log(n)\right)=\exp\left(-\frac{\sqrt{6}}{3\pi}\cdot\exp\left(-\frac{3\pi x}{\sqrt{6}}\right)\right).

Theorem 1.1 follows from a result which is of independent interest that generalizes a theorem of Erdős and Lehner on the distribution of the number of parts in partitions of fixed size. Using the celebrated Hardy-Ramanujan asymptotic formula

p(n)14n3exp(Cn),p(n)\sim\frac{1}{4n\sqrt{3}}\cdot\exp(C\sqrt{n}),

where C:=π2/3,C:=\pi\sqrt{2/3}, Erdős and Lehner determined the distribution of the number of parts in partitions of size nn. More precisely, if kn=kn(x):=C1nlog(n)+nx,k_{n}=k_{n}(x):=C^{-1}\sqrt{n}\log(n)+\sqrt{n}x, they proved (see Theorem 1.1 of [4]) that

limn+pkn(n)p(n)=exp(2Ce12Cx),\lim_{n\rightarrow+\infty}\frac{p_{\leq k_{n}}(n)}{p(n)}=\exp\left(-\frac{2}{C}e^{-\frac{1}{2}Cx}\right), (1.5)

where pk(n)p_{\leq k}(n) denotes333We note that pk(n)p_{\leq k}(n) is denoted pk(n)p_{k}(n) in [4]. the number of partitions of nn with at most kk parts. In particular, the normal order for the number of parts of a partition of size nn is C1nlog(n).C^{-1}\sqrt{n}\log(n).

To prove Theorem 1.1, the generalization of the observation of Hausel and Rodriguez-Villegas, we require the distribution of the number of parts in partitions that are multiples of a fixed integer A2A\geq 2. The next theorem describes these distributions.

Theorem 1.2.

If A2A\geq 2 and pk(A;n)p_{\leq k}(A;n) denotes the number of partitions of nn with at most kk parts that are multiples of AA, then for kA,n=kA,n(x):=1ACnlog(n)+xn,k_{A,n}=k_{A,n}(x):=\frac{1}{AC}\sqrt{n}\log(n)+x\sqrt{n}, we have

limn+pkA,n(A;n)p(n)=exp(2ACexp(12xAC)).\lim_{n\rightarrow+\infty}\frac{p_{\leq k_{A,n}}(A;n)}{p(n)}=\exp\left(-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right).
Remark.

The distribution functions in Theorem 1.2 are of Gumbel type with mean 2AC(log(2AC)+γ)\frac{2}{AC}\left(\log\left(\frac{2}{AC}\right)+\gamma\right) and variance 1/A2.1/A^{2}.

Example.

Here we illustrate Theorem 1.2 with A=2A=2 and n=600.n=600. In this case we have

k2,600(x):=600log(600)2C+600x.k_{2,600}(x):=\frac{\sqrt{600}\log(600)}{2C}+\sqrt{600}x.

For real numbers x,x, we let

δk2,600(x):=#{λ600 with k2,n(x) many even parts}p(n).\delta_{k_{2,600}}(x):=\frac{\#\{{\text{\rm$\lambda\vdash 600$ with $\leq k_{2,n}(x)$ many even parts}}\}}{p(n)}.

The theorem indicates that these proportions are approximated by the Gumbel values

G2,600(x):=exp(1CeCx).G_{2,600}(x):=\exp\left(-\frac{1}{C}\cdot e^{-Cx}\right).

The table below illustrates the strength of these approximations for various values of xx.

xx k2,600(x)\lfloor k_{2,600}(x)\rfloor δk2,600(x)\delta_{k_{2,600}}(x) G2,600(x)G_{2,600}(x)
0.1-0.1 2828 0.5970.597\dots 0.6040.604\dots
0.0\ 0.0 3030 0.6630.663\dots 0.6770.677\dots
0.1\ 0.1 3232 0.7210.721\dots 0.7390.739\dots
0.2\ 0.2 3535 0.7910.791\dots 0.7920.792\dots
0.3\ 0.3 3737 0.8300.830\dots 0.8350.835\dots
\vdots \vdots \vdots \vdots
1.5\ 1.5 6767 0.9940.994\dots 0.9920.992\dots
2.0\ 2.0 7979 0.9980.998\dots 0.9980.998\dots

We note that Theorem 1.2 does not offer the asymptotics for pk(A;n),p_{k}(A;n), the number of partitions of nn with exactly kk parts that are multiples of AA. For completeness, we offer such asymptotics, a result which is of independent interest. To make this precise, we recall the qq-Pochhammer symbol

(a;q)k:=n=0k1(1aqn).(a;q)_{k}:=\displaystyle\prod_{n=0}^{k-1}(1-aq^{n}).
Theorem 1.3.

If A2A\geq 2 is an integer, then the following are true.

(1) We have that pk(A;n)p_{k}(A;n) is the coefficient of TkqnT^{k}q^{n} in the infinite product

(qA;qA)(q;q)(TqA;qA).\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(Tq^{A};q^{A})_{\infty}}.

(2) For every non-negative integer n,n, we have pk(A;n)=pk(A;nAk).p_{k}(A;n)=p_{\leq k}(A;n-Ak). Moreover, we have

(qA;qA)(q;q)(qA;qA)k=n0pk(A;n)qn.\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(q^{A};q^{A})_{k}}=\sum_{n\geq 0}p_{\leq k}(A;n)q^{n}.

(3) For fixed kk, as n+,n\rightarrow+\infty, we have the asymptotic formulas

pk(A;n)24k214nk2342(11A)k214k!Ak+12(2π)ke2π16(11A)n,pk(A;n)24k214(nAk)k2342(11A)k214k!Ak+12(2π)ke2π16(11A)(nAk).\begin{split}&p_{\leq k}(A;n)\sim\frac{24^{\frac{k}{2}-\frac{1}{4}}n^{\frac{k}{2}-\frac{3}{4}}}{\sqrt{2}\left(1-\frac{1}{A}\right)^{\frac{k}{2}-\frac{1}{4}}k!A^{k+\frac{1}{2}}(2\pi)^{k}}e^{2\pi\sqrt{\frac{1}{6}\left(1-\frac{1}{A}\right)n}},\\ &p_{k}(A;n)\sim\frac{24^{\frac{k}{2}-\frac{1}{4}}(n-Ak)^{\frac{k}{2}-\frac{3}{4}}}{\sqrt{2}\left(1-\frac{1}{A}\right)^{\frac{k}{2}-\frac{1}{4}}k!A^{k+\frac{1}{2}}(2\pi)^{k}}e^{2\pi\sqrt{\frac{1}{6}\left(1-\frac{1}{A}\right)(n-Ak)}}.\end{split}
Example.

Here we illustrate the convergence of the asymptotic for p1(3;n).p_{1}(3;n). Theorem 1.3 (3) gives

p1(3;n)16π(n3)14e2πn33.p_{1}(3;n)\sim\frac{1}{6\pi(n-3)^{\frac{1}{4}}}e^{\frac{2\pi\sqrt{n-3}}{3}}.

For convenience, we let p1(3;n)p^{*}_{1}(3;n) denote the right hand side of this asymptotic. The table below illustrates the convergence of the asymptotic.

nn p1(3;n)p_{1}(3;n) p1(3;n)p_{1}^{*}(3;n) p1(3;n)/p1(3;n)p_{1}(3;n)/p_{1}^{*}(3;n)
200200 9312582384793125823847 82738081118\approx 82738081118 1.126\approx 1.126
400400 1.718×1016\approx 1.718\times 10^{16} 1.579×1016\approx 1.579\times 10^{16} 1.088\approx 1.088
600600 1.928×1020\approx 1.928\times 10^{20} 1.799×1020\approx 1.799\times 10^{20} 1.071\approx 1.071
800800 5.058×1023\approx 5.058\times 10^{23} 4.764×1023\approx 4.764\times 10^{23} 1.062\approx 1.062
10001000 5.232×1026\approx 5.232\times 10^{26} 4.959×1026\approx 4.959\times 10^{26} 1.055\approx 1.055
Table 1. Asymptotics for p1(3;n)p_{1}(3;n)

This paper is organized as follows. In Section 2 we prove Theorem 1.2, the generalization of the classical limiting distribution (1.5) of Erdős and Lehner. In Section 3, we recall the work of Buryak, Feigin, and Nakajima [2, 3], which gives the infinite product generating functions for the Poincaré polynomials P(Xα,β[n];T).P\left(X^{[n]}_{\alpha,\beta};T\right). These generating functions relate the Betti numbers to the partition functions pk().p_{\leq k}(\cdot). We use these facts, combined with Theorem 1.2, to obtain Theorem 1.1. Finally, in Section 4 we obtain Theorem 1.3, the asymptotic formulas for the pk(A;n)p_{\leq k}(A;n) and pk(A;n)p_{k}(A;n) partition functions. These asymptotics follow from an application of Ingham’s Tauberian theorem.

Acknowledgements

The authors thank Kathrin Bringmann for helpful comments on an earlier version of this paper, and for pointing out the corrected version of the statement of Ingham’s Tauberian Theorem. The authors also thank Ole Warnaar for comments on an earlier version of this paper. The second author thanks the support of the Thomas Jefferson Fund and the NSF (DMS-1601306 and DMS-2055118), and the Kavli Institute grant NSF PHY-1748958. The third author is grateful for the support of a grant from the Simons Foundation (853830, LR) and a 2021-2023 Dean’s Faculty Fellowship from Vanderbilt University. Finally, the authors thank the referee for pointing out typographical errors in the original manuscript.

2. Generalization of a theorem of Erdős and Lehner

Here we prove Theorem 1.2. To prove the theorem we combine some elementary observations about integer partitions with delicate asymptotic analysis.

2.1. Elementary considerations

First we begin with an elementary convolution involving the partition functions pk(A;)p_{\leq k}(A;\cdot), pk(),p_{\leq k}(\cdot), and preg(A;n)p_{\mathrm{reg}}(A;n), the number of AA-regular partitions of size n.n. Recall that a partition is AA-regular if all of its parts are not multiples of AA.

Proposition 2.1.

If A2A\geq 2 is a positive integer, then for every positive integer nn we have

pk(A;n)=j=0nApk(j)preg(A;nAj).p_{\leq k}(A;n)=\sum_{j=0}^{\lfloor\frac{n}{A}\rfloor}p_{\leq k}(j)\cdot p_{\mathrm{reg}}(A;n-Aj).
Proof.

Every partition of nn with at most kk parts that are multiples of AA can be represented as the direct product of an AA-regular partition and a partition into at most kk parts that are all multiples of AA. If the sum of these multiples of AA is AjAj, then the AA-regular partition has size nAjn-Aj. Moreover, by dividing by AA, the multiples of AA are represented by a partition of jj into at most kk parts. This proves the claimed convolution. ∎

We also require an elegant inclusion-exclusion formula due to Erdős and Lehner [4] for pk(n).p_{\leq k}(n).

Proposition 2.2.

If kk and jj are positive integers, then

pk(j)=m=0(1)mSk(m;j),p_{\leq k}(j)=\sum_{m=0}^{\infty}(-1)^{m}S_{k}(m;j),

where444The Sk(m;j)/p(j)S_{k}(m;j)/p(j) are denoted SmS_{m} in [4].

Sk(m;j):=1r1<r2<<rmTmr1+r2++rmjmkp(ji=1m(k+ri))S_{k}(m;j):=\sum_{\begin{subarray}{c}1\leq r_{1}<r_{2}<\dots<r_{m}\\ T_{m}\leq r_{1}+r_{2}+\dots+r_{m}\leq j-mk\end{subarray}}p\left(j-\sum_{i=1}^{m}(k+r_{i})\right) (2.1)

and Tm:=m(m+1)/2.T_{m}:=m(m+1)/2.

Proof.

By definition, pk(j)p_{\leq k}(j) is the number of partitions of jj with at most kk parts. By considering conjugates of partitions, one can equivalently define pk(j)p_{\leq k}(j) as the number of partitions of jj with no parts k+1\geq k+1. Since the number of partitions of size jj that contain a part of size k+r,k+r, where r1r\geq 1, equals p(j(k+r)),p(j-(k+r)), we find that Sk(1,j)S_{k}(1,j) is generally an overcount for the number of partitions of jj with at least one part k+1.\geq k+1. Due to this overcounting, we find that

p(j)Sk(1;j)pk(j)p(j)Sk(1;j)+Sk(2;j),p(j)-S_{k}(1;j)\leq p_{\leq k}(j)\leq p(j)-S_{k}(1;j)+S_{k}(2;j),

which is obtained by taking into account those partitions which have at least two parts of distinct size k+1.\geq k+1. The claim follows in this way by inclusion-exclusion. ∎

2.2. Proof of Theorem 1.2

To prove Theorem 1.2, we require Propositions 2.1 and  2.2, and the asymptotics for preg(A;n).p_{\mathrm{reg}}(A;n). Thanks to the identity

n=1(1+qn+q2n++q(A1)n)=n=1(1qAn)(1qn)=n=0preg(A;n)qn,\prod_{n=1}^{\infty}(1+q^{n}+q^{2n}+\dots+q^{(A-1)n})=\prod_{n=1}^{\infty}\frac{(1-q^{An})}{(1-q^{n})}=\sum_{n=0}^{\infty}p_{\mathrm{reg}}(A;n)q^{n},

we find that preg(A;n)p_{\mathrm{reg}}(A;n) equals the number of partitions of nn where no part occurs more than A1A-1 times. Hagis [11] obtained asymptotics for the number of partitions where no part is repeated more than tt times, and letting t=A1t=A-1 in Corollary 4.2 of [11] gives the following theorem.

Theorem 2.3.

If A2,A\geq 2, then we have

preg(A;n)=CA(24n1+A)34exp(CA1A(n+A124))(1+O(n12)),p_{\mathrm{reg}}(A;n)=C_{A}(24n-1+A)^{-\frac{3}{4}}\exp\left(C\sqrt{\frac{A-1}{A}\left(n+\frac{A-1}{24}\right)}\right)\left(1+O(n^{-\frac{1}{2}})\right),

where C:=π2/3C:=\pi\sqrt{2/3} and CA:=12A34(A1)14,C_{A}:=\sqrt{12}A^{-\frac{3}{4}}(A-1)^{\frac{1}{4}}, and the implied constant is independent of AA.

Proof of Theorem 1.2.

Thanks to Propositions 2.1 and  2.2, we have that

pk(A;n)p(n)=j=0nA(m=0(1)mSk(m;j))preg(A;nAj)p(n).\frac{p_{\leq k}(A;n)}{p(n)}=\sum_{j=0}^{\lfloor\frac{n}{A}\rfloor}\frac{\left(\sum_{m=0}^{\infty}(-1)^{m}S_{k}(m;j)\right)p_{\mathrm{reg}}(A;n-Aj)}{p(n)}. (2.2)

The proof follows directly from this expression by a sequence of observations involving the asymptotics for p()p(\cdot) and preg(A;),p_{\mathrm{reg}}(A;\cdot), combined with the earlier work of Erdős and Lehner on the sums Sk(m;j).S_{k}(m;j). Thanks to the special choice of kn=kn(x)k_{n}=k_{n}(x), this expression yields the Taylor expansion of the claimed cumulative Gumbel distribution in x,x, as n+.n\rightarrow+\infty. In other words, these asymptotics conspire so that the dependence on nn vanishes in the limit.

For convenience, we let Sk(m;j):=Sk(m;j)/p(j).S^{*}_{k}(m;j):=S_{k}(m;j)/p(j). In terms of Sk(m,j),S^{*}_{k}(m,j), (2.2) becomes

pk(A;n)p(n)=j=0nA(m=0(1)mSk(m;j))p(j)preg(A;nAj)p(n).\frac{p_{\leq k}(A;n)}{p(n)}=\sum_{j=0}^{\lfloor\frac{n}{A}\rfloor}\frac{\left(\sum_{m=0}^{\infty}(-1)^{m}S^{*}_{k}(m;j)\right)p(j)p_{\mathrm{reg}}(A;n-Aj)}{p(n)}. (2.3)

To make use of this formula, we begin by employing the method of Erdős and Lehner mutatis mutandis, which we briefly recapitulate here. For k+,k\rightarrow+\infty, with jj and mm fixed, Erdős and Lehner proved (see (2.5) of [4]) that

Sk(m;j)=1m!(2Cjexp(C2jk))m+oj,m(1).S^{*}_{k}(m;j)=\frac{1}{m!}\left(\frac{2}{C}\sqrt{j}\exp\left(-\frac{C}{2\sqrt{j}}k\right)\right)^{m}+o_{j,m}(1). (2.4)

For every positive integer m,m, this effectively gives

Sk(m;j)=1m!Sk(1;j)m+oj,m(1)1m!Sk(1;j)m,S^{*}_{k}(m;j)=\frac{1}{m!}\cdot S^{*}_{k}(1;j)^{m}+o_{j,m}(1)\sim\frac{1}{m!}\cdot S_{k}^{*}(1;j)^{m},

which Erdős and Lehner show produces, as functions in xx, the asymptotic

m=0(1)mSkn(m;j)=exp(Skn(1;j))(1+on(1)).\sum_{m=0}^{\infty}(-1)^{m}S^{*}_{k_{n}}(m;j)=\exp(-S^{*}_{k_{n}}(1;j))\left(1+o_{n}(1)\right). (2.5)

We recall the choice of k=kA,n=kA,n(x)=1ACnlog(n)+xn.k=k_{A,n}=k_{A,n}(x)=\frac{1}{AC}\sqrt{n}\log(n)+x\sqrt{n}. This is the exponential which arises in the exponential of the claimed cumulative distribution.

To make use of (2.5), it is convenient to recenter the sum on jj in (2.3) by setting j=nA2+y.j=\frac{n}{A^{2}}+y. As (2.5) only involves Skn(1;j),S^{*}_{k_{n}}(1;j), it suffices to note that when m=1,m=1, (2.4) becomes

SkA,n(1;j)\displaystyle S^{*}_{{\color[rgb]{0,0,0}k_{A,n}}}(1;j) =2ACn+A2yexp(log(n)21+yA2/nxAC21+yA2/n)+on(1).\displaystyle=\frac{2}{AC}\sqrt{n+A^{2}y}\cdot\exp\left(-\frac{\log(n)}{2\sqrt{1+yA^{2}/n}}-\frac{xAC}{2\sqrt{1+yA^{2}/n}}\right)+o_{n}(1). (2.6)

As the proof relies on (2.3), we must also estimate the quotients

p(j)preg(A;nAj)p(n).\frac{p(j)p_{\mathrm{reg}}(A;n-Aj)}{p(n)}.

Thanks to the Hardy-Ramanujan asymptotic for p(n)p(n) and Theorem 2.3, we have

p(j)preg(A;nAj)p(n)\displaystyle\frac{p(j)p_{\mathrm{reg}}(A;n-Aj)}{p(n)}
=CA(24n24Aj1+A)34njexp(C(jn+A1A(nAj+A124)))(1+Oj(n12))\displaystyle=\frac{C_{A}}{(24n-24Aj-1+A)^{\frac{3}{4}}}\frac{n}{j}\exp{\left(C\left(\sqrt{j}-\sqrt{n}+\sqrt{\frac{A-1}{A}\left(n-Aj+\frac{A-1}{24}\right)}\right)\right)}\cdot\left(1+O_{j}(n^{-\frac{1}{2}})\right)
=CA(24n24n/A24Ay1+A)34A2nn+A2y\displaystyle=\frac{C_{A}}{(24n-24n/A-24Ay-1+A)^{\frac{3}{4}}}\frac{A^{2}n}{n+A^{2}y}
×exp(C(n/A2+yn+A1A(nn/AAy+A124)))(1+Oy(n12)).\displaystyle\ \ \ \ \ \ \ \ \ \ \times\exp{\left(C\left(\sqrt{n/A^{2}+y}-\sqrt{n}+\sqrt{\frac{A-1}{A}\left(n-n/A-Ay+\frac{A-1}{24}\right)}\right)\right)}\cdot\left(1+O_{y}(n^{-\frac{1}{2}})\right). (2.7)

The last manipulation uses the change of variable for jj.

We will make use of (2.5), (2.6) and (2.2) to complete the proof. To this end, we let j=n/A2+yj=\lfloor n/A^{2}\rfloor+y essentially as above, but now modified555We can ignore the difference between n/A2\lfloor n/A^{2}\rfloor with n/A2n/A^{2} as it makes no difference for our limit calculations. so that the yy are integers. We then rewrite (2.3) as

pkA,n(A;n)p(n)=Σ1+Σ2+Σ3,\frac{p_{\leq{\color[rgb]{0,0,0}k_{A,n}}}(A;n)}{p(n)}=\Sigma_{1}+\Sigma_{2}+\Sigma_{3},

where Σ1\Sigma_{1} is the sum over n/A2y<n3/4log(n)-n/A^{2}\leq y<-n^{3/4}\log(n), Σ2\Sigma_{2} is the sum over n3/4log(n)yn3/4log(n),-n^{3/4}\log(n)\leq y\leq n^{3/4}\log(n), and Σ3\Sigma_{3} is the sum over n3/4log(n)yn(1/A1/A2).n^{3/4}\log(n)\leq y\leq n(1/A-1/A^{2}). We shall show that the main contribution will come from Σ2,\Sigma_{2}, and that Σ1\Sigma_{1} and Σ3\Sigma_{3} vanish as n+.n\rightarrow+\infty.

To establish the vanishing of Σ1+Σ3,\Sigma_{1}+\Sigma_{3}, we consider the case that |y|>n3/4log(n).|y|>n^{3/4}\log(n). For such yy we have

n/A2+yn+A1A(nn/AAy+A124)=Oy(n),\sqrt{n/A^{2}+y}-\sqrt{n}+\sqrt{\frac{A-1}{A}\left(n-n/A-Ay+\frac{A-1}{24}\right)}=O_{y}(\sqrt{n}),

where the implied constant is negative. Moreover, (2.6) implies that SkA,n(1;n/A2+y)=O(n),S^{*}_{{\color[rgb]{0,0,0}k_{A,n}}}(1;n/A^{2}+y)=O(\sqrt{n}), where the implied constant is positive. Thus, for yy in these ranges, both p(j)p(n)preg(A;nAj)\frac{p(j)}{p(n)}p_{\mathrm{reg}}(A;n-Aj) and m=0(1)mSkA,n(m;j)\sum_{m=0}^{\infty}(-1)^{m}S^{*}_{{\color[rgb]{0,0,0}k_{A,n}}}(m;j) decay sub-exponentially, and so

limnΣ1+Σ3=0.\lim_{n\to\infty}\Sigma_{1}+\Sigma_{3}=0.

We now consider Σ2\Sigma_{2}, where |y|n3/4log(n).|y|\leq n^{3/4}\log(n). In this range, (2.6) becomes

SkA,n(1;j)\displaystyle S^{*}_{{\color[rgb]{0,0,0}k_{A,n}}}(1;j) =2ACn+A2yexp(log(n)21+yA2/nxAC21+yA2/n)+on(1)\displaystyle=\frac{2}{AC}\sqrt{n+A^{2}y}\cdot\exp\left(-\frac{\log(n)}{2\sqrt{1+yA^{2}/n}}-\frac{xAC}{2\sqrt{1+yA^{2}/n}}\right)+o_{n}(1) (2.8)
=2ACexp(12xAC)+on(1).\displaystyle=\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)+o_{n}(1). (2.9)

Using (2.5), we obtain

m=0(1)mSkA,n(m;j)=exp(2ACexp(12xAC))(1+on(1)).\sum_{m=0}^{\infty}(-1)^{m}S^{*}_{{\color[rgb]{0,0,0}k_{A,n}}}(m;j)=\exp\left(-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right)\left(1+o_{n}(1)\right). (2.10)

We now estimate (2.2) for these |y|n3/4log(n).|y|\leq n^{3/4}\log(n). Since we have

n/A2+yn+A1A(nn/AAy+A124)\displaystyle\sqrt{n/A^{2}+y}-\sqrt{n}+\sqrt{\frac{A-1}{A}\left(n-n/A-Ay+\frac{A-1}{24}\right)} =A48(A1)y2n3/2+OA(y3n5/2),\displaystyle=-\frac{A^{4}}{8(A-1)}y^{2}n^{-3/2}+O_{A}(y^{3}n^{-5/2}),

the hypothesis on yy allows us to turn (2.2) into

p(j)preg(A;nAj)p(n)=A2CA(24nA1A)3/4×exp(CA48(A1)y2n3/2)(1+OA(n14+ε)).\frac{p(j)p_{\mathrm{reg}}(A;n-Aj)}{p(n)}=\frac{A^{2}C_{A}}{(24n\frac{A-1}{A})^{3/4}}\times\exp{\left(-C\frac{A^{4}}{8(A-1)}\frac{y^{2}}{n^{3/2}}\right)}\cdot\left(1+O_{A}(n^{-\frac{1}{4}+\varepsilon})\right).

Combined with (2.10), and using CA=12A34(A1)14,C_{A}=\sqrt{12}A^{-\frac{3}{4}}(A-1)^{\frac{1}{4}}, we obtain

limnΣ2=limn|y|<n3/4log(n)A2961/4A11n3/4exp(CA48(A1)y2n3/22ACexp(12xAC))(1+oA(1)).\begin{split}&\lim_{n\to\infty}\Sigma_{2}\\ &\ \ =\lim_{n\to\infty}\sum_{|y|<n^{3/4}\log(n)}\frac{A^{2}}{96^{1/4}\sqrt{A-1}}\cdot\frac{1}{n^{3/4}}\cdot\exp\left(-\frac{CA^{4}}{8(A-1)}\frac{y^{2}}{n^{3/2}}-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right)\cdot\left(1+o_{A}(1)\right).\end{split}

Approximating the right hand side as a Riemann sum, we obtain

limn+Σ2=limn+A2961/4A1log(n)log(n)exp(CA48(A1)t22ACexp(12xAC))𝑑t,\lim_{n\rightarrow+\infty}\Sigma_{2}=\lim_{n\rightarrow+\infty}\frac{A^{2}}{96^{1/4}\sqrt{A-1}}\int_{-\log(n)}^{\log(n)}\exp\left(-\frac{CA^{4}}{8(A-1)}t^{2}-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right)dt, (2.11)

where nn only appears in the limits of integration. To obtain this, we have used the substitutions t=yn3/4t=yn^{-3/4} and dt=n3/4dy,dt=n^{-3/4}dy, and employ the fact that the widths of the subintervals defining the Riemann sums tend to 0. Expanding as an integral over \mathbb{R}, this expression simplifies to

exp(2ACexp(12xAC)).\exp\left(-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right).

Therefore, as a function in x,x, we have

limn+pkA,n(A;n)p(n)=exp(2ACexp(12xAC)).\lim_{n\rightarrow+\infty}\frac{p_{\leq k_{A,n}}(A;n)}{p(n)}=\exp\left(-\frac{2}{AC}\exp\left(-\frac{1}{2}xAC\right)\right).

This completes the proof of the theorem.

3. Application to the Hilbert schemes Xα,β[n]X_{\alpha,\beta}^{[n]}

Here we recall the relevant generating functions for the Poincaré polynomials of the Hilbert schemes that pertain to Theorem 1.1. For the various Hilbert schemes on nn points, Göttsche, Buryak, Feigin, and Nakajima [2, 3, 9, 10] proved infinite product generating functions for these Poincaré polynomials. For Theorem 1.1, we require the following theorem.

Theorem 3.1.

(Buryak and Feigin) If α,β\alpha,\beta\in\mathbb{N} are relatively prime, then we have that

Gα,β(T;q):=n=0P(Xα,β[n];T)qn=(qα+β;qα+β)(q;q)(T2qα+β;qα+β).G_{\alpha,\beta}(T;q):=\sum_{n=0}^{\infty}P\left(X^{[n]}_{\alpha,\beta};T\right)q^{n}=\frac{(q^{\alpha+\beta};q^{\alpha+\beta})_{\infty}}{(q;q)_{\infty}(T^{2}q^{\alpha+\beta};q^{\alpha+\beta})_{\infty}.}
Remark.

The Poincaré polynomials in these cases only have even degree terms (i.e. odd index Betti numbers are zero). Moreover, letting T=1T=1 in these generating functions give Euler’s generating function for p(n).p(n). Therefore, we directly see that

p(n)=P(Xα,β[n];1).p(n)=P\left(X^{[n]}_{\alpha,\beta};1\right).

Of course, the proof of Theorem 3.1 begins with partitions of size nn.

Corollary 3.2.

Assuming the notation and hypotheses above, if dμα,β[n]d\mu^{[n]}_{\alpha,\beta} is the discrete measure for Xα,β[n]X^{[n]}_{\alpha,\beta}, then

Φn(α,β;x)=1p(n)x𝑑μα,β[n]=px2(α+β;n)p(n).\Phi_{n}(\alpha,\beta;x)=\frac{1}{p(n)}\cdot\int_{-\infty}^{x}d\mu^{[n]}_{\alpha,\beta}=\frac{p_{\leq\frac{x}{2}}(\alpha+\beta;n)}{p(n)}.
Proof.

By Theorem 3.1, the Poincaré polynomial P(Xα,β[n];T)P\left(X^{[n]}_{\alpha,\beta};T\right) is the coefficient of qnq^{n} of

(qα+β;qα+β)(q;q)(T2qα+β;qα+β).\frac{(q^{\alpha+\beta};q^{\alpha+\beta})_{\infty}}{(q;q)_{\infty}(T^{2}q^{\alpha+\beta};q^{\alpha+\beta})_{\infty}}.

Part (1) of Theorem 1.3 applied to A=α+βA=\alpha+\beta gives that the coefficient of T2kT^{2k} in this expression is pk(α+β;n)p_{k}(\alpha+\beta;n) (the odd powers of TT do not appear in this product as it is a function of T2T^{2}). Therefore, (1.3) becomes

P(Xα,β[n];T)=j=0nα+βpj(α+β;n)T2j=j=02nα+βdim(Hj(Xα,β[n],))Tj.P\left(X^{[n]}_{\alpha,\beta};T\right)=\sum_{j=0}^{\lfloor\frac{n}{\alpha+\beta}\rfloor}p_{j}(\alpha+\beta;n)T^{2j}=\sum_{j=0}^{2\lfloor\frac{n}{\alpha+\beta}\rfloor}\dim\left(H_{j}\left(X^{[n]}_{\alpha,\beta},\mathbb{Q}\right)\right)T^{j}.

Thus, the sum of coefficients up to xx, divided by p(n)p(n), is

1p(n)jxbj(α,β;n)=1p(n)jx/2pj(α+β;n)=px2(α+β;n)p(n).\frac{1}{p(n)}\cdot\sum_{j\leq x}b_{j}(\alpha,\beta;n)=\frac{1}{p(n)}\cdot\sum_{j\leq x/2}p_{j}(\alpha+\beta;n)=\frac{p_{\leq\frac{x}{2}}(\alpha+\beta;n)}{p(n)}.

This completes the proof. ∎

Proof of Theorem 1.1.

To prove Theorem 1.1, we remind the reader that Theorem 1.2 gives the cumulative asymptotic distribution function for pk(A;n)p_{\leq k}(A;n) when A2.A\geq 2. Corollary 3.2, with A=α+β,A=\alpha+\beta, identifies this partition distribution with the Betti distribution for the nn point Hilbert schemes cut out by the α,β\alpha,\beta torus action. The theorem follows by combining these two results. ∎

4. Asymptotic formulae for the pk(A;n)p_{k}(A;n) partition functions

Here we prove Theorem 1.3. To this end, we make use of Ingham’s Tauberian theorem [12]. We note that this theorem is misstated in a number of places in the literature. Condition (3) in the statement below is often omitted. The reader is referred to the discussion in [1]. Here we use a special case666In the notation of [1], we let d=βd=\beta, N=γN=\gamma, and we let α=0\alpha=0 in the case of weak monotonicity of Theorem 1.1. of Theorem 1.1 of [1].

Theorem 4.1 (Ingham).

Let f(q)=n0a(n)qnf(q)=\sum_{n\geq 0}a(n)q^{n} be a holomorphic function in the unit disk
|q|<1|q|<1 satisfying the following conditions:

(1) The sequence {a(n)}n0\left\{a(n)\right\}_{n\geq 0} is positive and weakly monotonically increasing.

(2) There exist c,c\in\mathbb{C}, d,d\in\mathbb{R}, and N>0,N>0, such that as t0+t\rightarrow 0^{+} we have

f(et)λtdeNt.f(e^{-t})\sim\lambda\cdot t^{d}\cdot e^{\frac{N}{t}}.

(3) For any Δ>0\Delta>0, in the cone |y|Δx|y|\leq\Delta x with x>0x>0 and z=x+iyz=x+iy, we have, as z0z\rightarrow 0

f(ez)|z|deN|z|.f(e^{-z})\ll|z|^{d}\cdot e^{\frac{N}{|z|}}.

Then as n+n\rightarrow+\infty we have

a(n)λNd2+142πnd2+34e2Nn.a(n)\sim\frac{\lambda\cdot N^{\frac{d}{2}+\frac{1}{4}}}{2\sqrt{\pi}\cdot n^{\frac{d}{2}+\frac{3}{4}}}e^{2\sqrt{Nn}}.
Proof of Theorem 1.3.

We prove the claims one-by-one.

(1) We begin by recalling the qq-Pochhammer symbol

(a;q)k:=n=0k1(1aqn).(a;q)_{k}:=\displaystyle\prod_{n=0}^{k-1}(1-aq^{n}).

Clearly, we have

(qA;qA)(q;q)=j=1A11(qj;qA),\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}}=\prod_{j=1}^{A-1}\frac{1}{(q^{j};q^{A})_{\infty}},

which in turn gives

(qA;qA)(q;q)(TqA;qA)=n0(modA)11qn×n0(modA)11Tqn.\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(Tq^{A};q^{A})_{\infty}}=\prod_{n\not\equiv 0\,\,({\rm mod}\,\,{A})}\frac{1}{1-q^{n}}\quad\times\prod_{n\equiv 0\,\,({\rm mod}\,\,{A})}\frac{1}{1-Tq^{n}}.

Expanding each term as a geometric series, we find that the coefficient of TkT^{k} collects those partitions which have kk parts which are 0(modA)0\,\,({\rm mod}\,\,{A}).

(2) We make use of the qq-binomial theorem, which asserts that

n0(a;q)n(q;q)nzn=(az;q)(z;q).\sum_{n\geq 0}\frac{(a;q)_{n}}{(q;q)_{n}}z^{n}=\frac{(az;q)_{\infty}}{(z;q)_{\infty}}.

Hence, if we let [Tk][T^{k}] denote the coefficient of TkT^{k}, this theorem allows us to conclude that

(qA;qA)(q;q)[Tk](1(TqA;qA))=(qA;qA)(q;q)[Tk](n0(TqA)n(qA;qA)n)=qAk(qA;qA)(q;q)(qA;qA)k.\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}}[T^{k}]\left(\frac{1}{(Tq^{A};q^{A})_{\infty}}\right)=\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}}[T^{k}]\left(\sum_{n\geq 0}\frac{(Tq^{A})^{n}}{(q^{A};q^{A})_{n}}\right)=\frac{q^{Ak}(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(q^{A};q^{A})_{k}}.

Arguing as in the proof of (1), we find the claimed generating function identity

(qA;qA)(q;q)(qA;qA)k=n0pk(A;n)qn.\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(q^{A};q^{A})_{k}}=\sum_{n\geq 0}p_{\leq k}(A;n)q^{n}. (4.1)

These two qq-series identities, combined with (1), imply that pk(A;n)=pk(A;nAk).p_{k}(A;n)=p_{\leq k}(A;n-Ak).

(3) To establish the desired asymptotics, we apply Theorem 4.1 to (4.1), which is facilitated by the modularity of Dedekind’s eta-function

η(τ):=q124(q;q).\eta(\tau):=q^{\frac{1}{24}}(q;q)_{\infty}.

This function is well-known to satisfy

η(1τ)=τiη(τ).\eta\left(-\frac{1}{\tau}\right)=\sqrt{\frac{\tau}{i}}\eta(\tau).

As a consequence of this transformation and the qq-expansion η(τ)=q124+O(q2524)\eta(\tau)=q^{\frac{1}{24}}+O(q^{\frac{25}{24}}) near τ=i\tau=i\infty (for example, see p. 53 of [14]), for q=etq=e^{-t}, t0+,t\rightarrow 0^{+}, we find that

log(1(q;q))=π26t12log(2πt)+O(t).\log\left(\frac{1}{(q;q)_{\infty}}\right)=\frac{\pi^{2}}{6t}-\frac{1}{2}\log\left(\frac{2\pi}{t}\right)+O(t). (4.2)

Thus, letting tAtt\mapsto At and taking a difference yields

log((qA;qA)(q;q))=π26t(11A)log(A)2+O(t).\log\left(\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}}\right)=\frac{\pi^{2}}{6t}\left(1-\frac{1}{A}\right)-\frac{\log(A)}{2}+O(t). (4.3)

This calculation gives the behavior in the radial limit as t0+t\rightarrow 0^{+} of the infinite Pochhammer symbols in (4.1).

To satisfy condition (3) of Theorem 4.1, we also need to estimate the quotient on the left hand side of (4.3) for the regions |y|Δx|y|\leq\Delta x. This is given directly in Section 3.1 of [1]. Namely, they show that in these regions, one has

1(ez;ez)=z2πeπ26z(1+OΔ(|e4π2z|))\frac{1}{(e^{-z};e^{-z})_{\infty}}=\sqrt{\frac{z}{2\pi}}\cdot\ e^{\frac{\pi^{2}}{6z}}\left(1+O_{\Delta}\left(\left|e^{-\frac{4\pi^{2}}{z}}\right|\right)\right)

and

e1ze1(1+Δ2)|z|.e^{-\frac{1}{z}}\leq e^{-\frac{1}{(1+\Delta^{2})|z|}}.

Thus, we have

1(ez;ez)=z2πeπ26z(1+OΔ(e4π2(1+Δ2)|z|)).\displaystyle\frac{1}{(e^{-z};e^{-z})_{\infty}}=\sqrt{\frac{z}{2\pi}}\cdot\ e^{\frac{\pi^{2}}{6z}}\left(1+O_{\Delta}\left(e^{-\frac{4\pi^{2}}{(1+\Delta^{2})|z|}}\right)\right). (4.4)

Changing variables to let zAzz\mapsto Az, we then find

(eAz;eAz)(ez;ez)=Aeπ26z(11A)(1+OΔ(e4π2A(1+Δ2)|z|))(1+OΔ(e4π2(1+Δ2)|z|))=Aeπ26z(11A)(1+OΔ(e4π2A(1+Δ2)|z|)).\frac{(e^{-Az};e^{-Az})_{\infty}}{(e^{-z};e^{-z})_{\infty}}=\sqrt{A}\cdot e^{-\frac{\pi^{2}}{6z}\left(1-\frac{1}{A}\right)}\cdot\frac{\left(1+O_{\Delta}\left(e^{-\frac{4\pi^{2}}{A(1+\Delta^{2})|z|}}\right)\right)}{\left(1+O_{\Delta}\left(e^{-\frac{4\pi^{2}}{(1+\Delta^{2})|z|}}\right)\right)}=\sqrt{A}\cdot e^{-\frac{\pi^{2}}{6z}\left(1-\frac{1}{A}\right)}\left(1+O_{\Delta}\left(e^{-\frac{4\pi^{2}}{A(1+\Delta^{2})|z|}}\right)\right). (4.5)

Now we turn to estimating the remaining factor in (4.1), namely, 1/(qA;qA)k1/(q^{A};q^{A})_{k}. On the line t0+t\rightarrow 0^{+}, an important result of Zhang (see Theorem 2 of [15]) gives that for 0<t00<t\rightarrow 0 and ww\in\mathbb{C},

(ewt;et)2πΓ(w)eπ26t(w12)log(t).(e^{-wt};e^{-t})_{\infty}\sim\frac{\sqrt{2\pi}}{\Gamma(w)}e^{-\frac{\pi^{2}}{6t}-\left(w-\frac{1}{2}\right)\log(t)}.

Letting w=k+1w=k+1 and combining with (4.2), we conclude that

1(q;q)k=(qk+1;q)(q;q)2πk!eπ26ε(k+1/2)log(t)+π26t12log(2π/t)=tkk!.\frac{1}{(q;q)_{k}}=\frac{(q^{k+1};q)_{\infty}}{(q;q)_{\infty}}\sim\frac{\sqrt{2}\pi}{k!}e^{-\frac{\pi^{2}}{6\varepsilon}-(k+1/2)\log(t)+\frac{\pi^{2}}{6t}-\frac{1}{2}\log(2\pi/t)}=\frac{t^{-k}}{k!}.

Letting tAtt\mapsto At, we have

1(qA;qA)k1k!Aktk.\frac{1}{(q^{A};q^{A})_{k}}\sim\frac{1}{k!A^{k}}t^{-k}. (4.6)

Turning to estimate 1/(qA;qA)k1/(q^{A};q^{A})_{k} in the regions |y|Δx|y|\leq\Delta x, we use the same argument in the proof of Theorem 2 of [15]. One merely modifies the proof by replacing xx with |z||z| in Zhang’s setting to obtain

(eA(k+1)z;eAz)2πk!eπ26|z|(k+112)log|z|,{(e^{-A(k+1)z};e^{-Az})_{\infty}}\ll\frac{\sqrt{2\pi}}{k!}e^{-\frac{\pi^{2}}{6|z|}-\left(k+1-\frac{1}{2}\right)\log|z|},

as z0.z\rightarrow 0. Moreover, by combining with (4.4), we have

1(eAz;eAz)k=(eA(k+1)z;eAz)(ez;ez)|z|kk!.\frac{1}{(e^{-Az};e^{-Az})_{k}}=\frac{(e^{-A(k+1)z};e^{-Az})_{\infty}}{(e^{-z};e^{-z})_{\infty}}\ll\frac{|z|^{-k}}{k!}. (4.7)

Then multiplying (4.5) and (4.7), we find that

(eAz;eAz)(ez;ez)(eAz;eAz)kAk!|z|keπ26|z|(11A),\frac{(e^{-Az};e^{-Az})_{\infty}}{(e^{-z};e^{-z})_{\infty}(e^{-Az};e^{-Az})_{k}}\ll\frac{\sqrt{A}}{k!}|z|^{-k}e^{\frac{\pi^{2}}{6|z|}\left(1-\frac{1}{A}\right)}, (4.8)

which shows that condition (3) of Theorem 4.1 is satisfied.

Multiplying (4.3) with (4.6), where q:=etq:=e^{-t}, we obtain

(qA;qA)(q;q)(qA;qA)k1k!Ak+12tkeπ26t(11A).\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(q^{A};q^{A})_{k}}\sim\frac{1}{k!A^{k+\frac{1}{2}}}t^{-k}e^{\frac{\pi^{2}}{6t}\left(1-\frac{1}{A}\right)}.

Moreover, the coefficients (qA;qA)(q;q)(qA;qA)k\frac{(q^{A};q^{A})_{\infty}}{(q;q)_{\infty}(q^{A};q^{A})_{k}} are clearly positive as they count partitions. They are weakly increasing as there is an easy injection from the set of partitions of nn with at most kk parts which are multiples of AA into the set of partitions of n+1n+1 which have at most kk parts which are multiples of AA; simply add 11 to the partition, which doesn’t affect the number of multiples of AA among the parts.

We are thus in the situation of Theorem 4.1, where we interprete (4.8) with

λ=1k!Ak+12,d=k,N=π26(11A).\lambda=\frac{1}{k!A^{k+\frac{1}{2}}},\quad d=-k,\quad N=\frac{\pi^{2}}{6}\left(1-\frac{1}{A}\right).

Plugging these into the Theorem 4.1 gives the desired asymptotic for pk(A;n).p_{\leq k}(A;n). The asymptotics for pk(A;n)p_{k}(A;n) follows from the identity pk(A;n)=pk(A;nAk)p_{k}(A;n)=p_{\leq k}(A;n-Ak) obtained in (2).

References

  • [1] K. Bringmann, C. Jennings-Shaffer, K. Mahlburg, On a Tauberian theorem of Ingham and Euler-Maclaurin summation, Ramanujan J. (2021), https://doi.org/10.1007/s11139-020-00377-5.
  • [2] A. Buryak and B. Feigin, Generating series of the Poincaré polynomials of quasihomogeneous Hilbert schemes, Integrable systems and representations, Proc. Math. Stat., Springer, 2013, 15–33.
  • [3] A. Buryak, B. Feigin, and H. Nakajima, A simple proof of the formula for the Betti numbers of the quasihomogeneous Hilbert schemes, Int. Math. Res. Notices 13 (2015), 4708–4715.
  • [4] P. Erdős and J. Lehner, The distribution of the number of summands in the partitions of a positive integer, Duke Math. J. 8 (1941), 335-345.
  • [5] E. J. Gumbel, Les valeurs extrêmes des distributions statistiques, Annales de L’Institut Henri Poincaré 5 (1935), 115-158.
  • [6] E. J. Gumbel, The return period of flood flows, Annals of Mathematical Statistics 12 (1941), 163-190.
  • [7] T. Hausel and F. Rodriguez-Villegas, Cohomology of large semiprojective hyperkähler varieties, Astérisque No. 370 (2015), 113-156.
  • [8] N. Gillman, X. Gonzalez, K. Ono, L. Rolen, and M. Schoenbauer, From partitions to Hodge numbers of Hilbert schemes of surfaces, Phil. Trans. Royal Soc., Series A 378 (2020), 20180435.
  • [9] L. Göttsche, Hilbert schemes of zero-dimensional subschemes of smooth varieties 1572, Springer Lect. Notes Math., 1994.
  • [10] L. Göttsche, Hilbert schemes of points on surfaces, ICM Proceedings, Vol. II (Beijing, 2002), 483–494.
  • [11] P. Hagis, Jr., Partitions with a restriction on the multiplicity of the summands, Trans. Amer. Math. Soc. 155 (1971), 375-384.
  • [12] A. Ingham, A Tauberian theorem for partitions. Ann. Math. 42 (1941), 1075-1090.
  • [13] H. Nakajima, Lectures on Hilbert Schemes of Points on Surfaces, vol 18 of University Lecture Series. Amer. Math. Soc., Providence, RI, 1999.
  • [14] D. Zagier The dilogarithm function, In Frontiers in Number Theory, Physics and Geometry II, P. Cartier, B. Julia, P. Moussa, P. Vanhove (eds.), Springer-Verlag, Berlin-Heidelberg-New York (2006), 3–65.
  • [15] R. Zhang, On asymptotics of the q-exponential and q-gamma functions, Journal of Mathematical Analysis and Applications 411 (2014), 522-529.