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

Cycles of given lengths in unicyclic components in sparse random graphs

Marc Noy Department of Mathematics, Universitat Politècnica de Catalunya
and Institut de Matemàtiques de la UPC-BarcelonaTech (IMTech), Spain.
[email protected]
Vonjy Rasendrahasina Ecole Normale Supérieure, Université d’Antananarivo, Madagascar. [email protected] Vlady Ravelomanana IRIF, Université de Paris, France. [email protected]  and  Juanjo Rué Department of Mathematics, Universitat Politècnica de Catalunya
and Barcelona Graduate School of Mathematics (BgsMath), Spain.
[email protected]
Abstract.

Let LL be subset of {3,4,}\{3,4,\dots\} and let Xn,M(L)X_{n,M}^{(L)} be the number of cycles belonging to unicyclic components whose length is in LL in the random graph G(n,M)G(n,M). We find the limiting distribution of Xn,M(L)X_{n,M}^{(L)} in the subcritical regime M=cnM=cn with c<1/2c<1/2 and the critical regime M=n2(1+μn1/3)M=\frac{n}{2}\left(1+\mu n^{-1/3}\right) with μ=O(1)\mu=O(1). Depending on the regime and a condition involving the series Lz/(2)\sum_{\ell\in L}z^{\ell}/(2\ell), we obtain in the limit either a Poisson or a normal distribution as nn\to\infty.

MSC: Primary 05A16, 01C80; Secondary 05C10

Keywords: Random graphs, random variables, analytic combinatorics.

1. Introduction

A graph is unicyclic if it is connected and has a unique cycle. We say that a cycle in a graph is isolated if it is the unique cycle in a unicyclic connected component. Let G(n,M)G(n,M) be the random graph with nn vertices and exactly MM edges drawn uniformly at random from the set of (n2){n\choose 2} possible edges. This is the model introduced in the seminal paper of Erdős and Rényi [3], in which each graph has the same probability

((n2)M)1.\binom{\binom{n}{2}}{M}^{-1}.

We are interested in the number of isolated cycles in G(n,M)G(n,M) whose lengths are restricted to take certain values. More precisely, let 3={3,4,}\mathbb{N}_{\geqslant 3}=\{3,4,\dots\} and LL a subset of 3\mathbb{N}_{\geqslant 3}. We denote by Xn,M(L)X_{n,M}^{(L)} the random variable equal to the number of isolated cycles in G(n,M)G(n,M) whose lengths lie in LL. Our main result gives the limiting distribution of Xn,M(L)X_{n,M}^{(L)} for various values of MM, corresponding to the so-called subcritical and critical regimes. Depending on the regime and a condition involving the generating function λ(z)=Lz/(2)\lambda(z)=\sum_{\ell\in L}z^{\ell}/(2\ell), we obtain in the limit as nn\to\infty either a Poisson or a normal distribution.

The number of cycles in G(n,M)G(n,M) has been studied since the appearance of [3]. When M=cnM=cn, Erdős and Rényi showed [3, Theorem 3b] that the number of cycles of length kk converges to a Poisson law with parameter (2c)k/(2k)(2c)^{k}/(2k). Let Xn,MX_{n,\,M} be the random variable equal to the number of isolated cycles in G(n,M)G(n,M). When M=cnM=cn and c<1/2c<1/2, asymptotically almost surely (that is, with probability tending to 1 as nn\to\infty) all cycles are isolated. As a consequence we have

limn𝔼[Xn,cn]=k3(2c)k2k=12log112ccc2.\lim_{n\to\infty}{\mathbb{E}}[X_{n,cn}]=\sum_{k\geqslant 3}\frac{(2c)^{k}}{2k}=\frac{1}{2}\log\frac{1}{1-2c}-c-c^{2}.

We next recall the different regimes for sparse random graphs (see for instance [7, 1]). The following results hold asymptotically almost surely (shortened to a.a.s.).

  • Subcritical regime. When M=cnM=cn with c<1/2c<1/2, the connected components of G(n,M)G(n,M) are either trees or unicyclic graphs.

  • Barely subcritical regime. When M=n2(1μn1/3)M=\frac{n}{2}\left(1-\mu n^{-1/3}\right) with μ\mu\to\infty and μ=o(n1/3)\mu=o\left(n^{1/3}\right),

  • Critical regime. This is when M=n2(1+μn1/3)M=\frac{n}{2}\left(1+\mu n^{-1/3}\right) and μ=O(1)\mu=O(1). In this regime the connected components of G(n,M)G(n,M) are trees, unicyclic graphs, and complex components. A complex component is obtained from a connected cubic multigraph KK by performing the following operations: first replace edges in KK by induced paths of any length so that to obtain a simple graph CC, and then attach rooted trees to the vertices of CC.

  • Supercritical regime. When M=cnM=cn with c>1/2c>1/2, there exists a unique component LL of linear size and the remaining components are either trees or unicyclic graphs. The ‘Symmetry principle’ (see [7, Section 5.6]) says that in this case G(n.M)\LG(n.M)\backslash L in some sense ‘looks like’ a subcritical random graph with suitable parameters.

In the barely subcritical regime Kolchin showed that if r0=16logn12logμr_{0}=\frac{1}{6}\log n-\frac{1}{2}\log\mu, then the normalized random variable (Xn,Mr0)/r0(X_{n,M}-r_{0})/\sqrt{r_{0}} tends in distribution to a Gaussian law (see [8, Theorem 1.1.15]). In the critical regime, Flajolet, Knuth and Pittel [4, Corollary 6] showed that 𝔼[Xn,M]16logn{\mathbb{E}}[X_{n,M}]\sim\frac{1}{6}\log n. By the so-called symmetry property  [7, Theorem 5.24], Xn,MX_{n,M} properly normalized should also be Gaussian when M=n2(1+μn1/3)M=\frac{n}{2}(1+\mu n^{-1/3}) and μ\mu\to\infty with μ=o(n1/3)\mu=o(n^{1/3}).

Some results have been obtained fixing a set LL of positive integers as possible cycle lengths. Following [4], define an LL-cycle as an isolated cycle whose length is in LL. Let Xn,M(L)X_{n,\,M}^{(L)} be the number of LL-cycles in G(n,M)G(n,\,M). It is shown in [4, Corollary 7] that if limn2Mn=λ<1\lim_{n\rightarrow\infty}\frac{2M}{n}=\lambda<1, then the probability that a graph (or multigraph) with nn vertices and MM edges has no LL-cycle is equal to

1λexp(l1,Lλ2l)+O(n1/2)=exp(l1,Lλ2l)+O(n1/2).\sqrt{1-\lambda}\,\exp{\left(\sum_{l\geqslant 1,\ell\notin L}\frac{\lambda^{\ell}}{2l}\right)}+O\left(n^{-1/2}\right)=\exp{\left(-\sum_{l\geqslant 1,\ell\in L}\frac{\lambda^{\ell}}{2l}\right)}+O\left(n^{-1/2}\right). (1)

Our results concern the distribution of the random variables Xn,M(L)X_{n,M}^{(L)}. In particular, we obtain full limiting distributions both in the subcritical and the critical regimes.

Theorem 1.1.

Let L3L\subseteq\mathbb{N}_{\geqslant 3} and set λL(z)=Lz2\lambda_{L}(z)=\sum_{\ell\in L}\frac{z^{\ell}}{2\ell}, considered as a function of one complex variable in the unit disk |z|<1|z|<1. Let Xn,M(L)X_{n,M}^{(L)} be the random variable equal to the number of LL-cycles in G(n,M)G(n,M). Then the following holds:

  1. (A)

    (Subcritical regime). Let c=c(n)c=c(n) be such that 0<lim supnc<1/20<\limsup_{n\to\infty}c<1/2 and M=cnM=cn. Then

    Xn,M(L)λL(2c)𝑑Poisson(1), as n.\frac{X_{n,\,M}^{(L)}}{\lambda_{L}(2c)}\mathbin{\stackrel{{\scriptstyle\mathop{d}}}{{\longrightarrow}}}\mathbin{\mathop{\mathrm{Poisson}}}\left(1\right),\quad\hbox{ as $n\to\infty$}. (2)
  2. (B)

    (Barely subcritical regime). Let M=n2(1μn1/3)M=\frac{n}{2}(1-\mu n^{-1/3}) with limμ=+\lim\mu=+\infty and μ=o(n1/3)\mu=o(n^{1/3}). Then two situations may happen: if limnλL(2Mn)<+\lim_{n\to\infty}\lambda_{L}(\frac{2M}{n})<+\infty, then

    Xn,M(L)λL(2Mn)𝑑Poisson(1), as n.\frac{X_{n,\,M}^{(L)}}{\lambda_{L}\left(\frac{2M}{n}\right)}\mathbin{\stackrel{{\scriptstyle\mathop{d}}}{{\longrightarrow}}}\mathbin{\mathop{\mathrm{Poisson}}}\left(1\right),\quad\hbox{ as $n\to\infty$}. (3)

    Otherwise, if limnλL(2Mn)=+\lim_{n\to\infty}\lambda_{L}(\frac{2M}{n})=+\infty, then

    Xn,M(L)λL(2Mn)λL(2Mn)𝑑𝒩(0,1), as n.\frac{X_{n,\,M}^{(L)}-\lambda_{L}\left(\frac{2M}{n}\right)}{\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)}}\mathbin{\stackrel{{\scriptstyle\mathop{d}}}{{\longrightarrow}}}\mathcal{N}(0,1),\quad\hbox{ as $n\to\infty$}. (4)
  3. (C)

    (Critical regime). Let M=n2(1+μn1/3)M=\frac{n}{2}(1+\mu n^{-1/3}), with μ=O(1)\mu=O(1). Let α\alpha be the unique positive solution of μ=1αα\mu=\frac{1}{\alpha}-\alpha. Then two situations may happen: if limnλL(eαn1/3)<+\lim_{n\to\infty}\lambda_{L}(e^{-\alpha n^{-1/3}})<+\infty, then

    Xn,M(L)λL(eαn1/3)𝑑Poisson(1), as n.\frac{X_{n,\,M}^{(L)}}{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}\mathbin{\stackrel{{\scriptstyle\mathop{d}}}{{\longrightarrow}}}\mathbin{\mathop{\mathrm{Poisson}}}\left(1\right),\quad\hbox{ as $n\to\infty$}. (5)

    Otherwise, if limnλL(eαn1/3)=+\lim_{n\to\infty}\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)=+\infty, then

    Xn,M(L)λL(eαn1/3)λL(eαn1/3)𝑑𝒩(0,1), as n.\frac{X_{n,\,M}^{(L)}-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}{\sqrt{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}}\mathbin{\stackrel{{\scriptstyle\mathop{d}}}{{\longrightarrow}}}\mathcal{N}(0,1),\quad\hbox{ as $n\to\infty$}. (6)

Points (A), (B) and (C) in Theorem 1.1 are the contents of Theorems 3.1, 3.2 and 3.4 given in the Section 3. We remark that in the previous statement there is no discontinuity between equations (2)–(3)–(5) and equations (4)–(6): the Taylor expansion of the term eαn1/3e^{-\alpha n^{-1/3}} in the statement for the critical regime is equal to 1αn1/3+o(n1/3)1-\alpha n^{-1/3}+o(n^{-1/3}), which coincides with the term 1μn1/31-\mu n^{-1/3} in the barely subcritical region.

The proofs are based on estimating coefficients of generating functions by means of Cauchy integrals along suitable contours and applying the saddle-point method.

Remarks. Observe that (1) follows directly from (2). Let us mention that technical refinements of our techniques would provide similar results for the region just before the supercritical regime, namely M=n2(1+μn1/3)M=\frac{n}{2}(1+\mu n^{-1/3}) when μ\mu\to\infty, μ=o(n1/12)\mu=o(n^{1/12}). We do not include the analysis of this region because the computations become too involved.

Finally, one may wonder why in the previous theorem we do not have a corresponding result for the supercritical regime. The reason is that in this case our techniques, based on the detailed structure of G(n,p)G(n,p) together with saddle-point estimates for the associated generating functions, do not apply in this situation. Given the Symmetry principle mentioned above, one should expect the number of LL-cycles in the supercritical regime follows a limit Poisson law as in the subcritical regime, but the tools provided by the Symmetry principle do no seem precise enough to prove such a statement.

2. Preliminaries and notation

All graphs considered in this paper are labelled. The size of a graph is the number of vertices. The excess of a graph GG is the number of vertices minus the number of edges. In G(n,M)G(n,M) the excess is MnM-n.

2.1. Analytic combinatorics of graphs

We use the language of analytic combinatorics as in [5]. Given a generating function A(x)=n0anxnA(x)=\sum_{n\geqslant 0}a_{n}x^{n}, we write [xn]A(x)=an[x^{n}]A(x)=a_{n}. If A(x)=n0anxnA(x)=\sum_{n\geqslant 0}a_{n}x^{n} and B(x)=n0bnxnB(x)=\sum_{n\geqslant 0}b_{n}x^{n}, we write A(x)B(x)A(x)\preceq B(x) if there exists n0n_{0} such that [xn]A(x)[xn]B(x)[x^{n}]A(x)\leqslant[x^{n}]B(x) for nn0n\geqslant n_{0}. All the generating functions that appear in this work are exponential generating functions of the form n0anxn/n!\sum_{n\geqslant 0}a_{n}x^{n}/n!, or EGF for short (see [5, Chapter 2]).

We denote by T(x)T(x) and W1(x)W_{-1}(x) the EGF of rooted and unrooted labelled trees, respectively. It is well known that

T(x)=xeT(x)=n=1nn1xnn!,W1(x)=T(x)T(x)22.T(x)=xe^{T(x)}=\sum_{n=1}^{\infty}n^{n-1}\frac{x^{n}}{n!},\qquad W_{-1}(x)=T(x)-\frac{T(x)^{2}}{2}. (7)

The EGF W0(x)W_{0}(x) of unicyclic graphs (connected graphs with nn vertices and nn edges) is given by (see for instance [6, Equation (3.5)])

W0(x)=k3T(x)k2k=12log(1T(x))T(x)2T(x)24.W_{0}(x)=\sum_{k\geqslant 3}\frac{T(x)^{k}}{2k}=-\frac{1}{2}\log{(1-T(x))}-\frac{T(x)}{2}-\frac{T(x)^{2}}{4}. (8)

We write λ(t)=k3tk2k=12log(1t)t/2t2/4\lambda(t)=\sum_{k\geqslant 3}\frac{t^{k}}{2k}=-\frac{1}{2}\log(1-t)-{t}/{2}-{t^{2}}/{4}, so that W0(x)=λ(T(x))W_{0}(x)=\lambda(T(x)).

2.2. From Poisson parametrizations to central limit theorems

We include the following result by Kolchin that provides an approximation to a normal law by a Poisson parametrization.

Theorem 2.1 ([8, Theorem 1.1.15]).

Let k=λn+ρnλnk=\lambda_{n}+\rho_{n}\sqrt{\lambda_{n}}. If (1+ρn)6/λn0(1+\rho_{n})^{6}/\lambda_{n}\to 0 as nn\to\infty then

eλnλnkk!=12πλneρn2/2(1+ρn3ρn6λn+O(1+ρn6λn)).e^{-\lambda_{n}}\frac{\lambda_{n}^{k}}{k!}=\frac{1}{\sqrt{2\pi\lambda_{n}}}e^{-\rho_{n}^{2}/2}\left(1+\frac{\rho_{n}^{3}-\rho_{n}}{6\sqrt{\lambda_{n}}}+O\left(\frac{1+\rho_{n}^{6}}{\lambda_{n}}\right)\right).

3. Proof of Theorem 1.1

We present separately the proof for each regime in Theorem 1.1. The main idea in all proofs is to encode the typical structure of random graphs in the regime under consideration using generating functions and then obtain large power estimates by means of saddle point bounds.

3.1. Subcritical regime

In this regime, the connected components of G(n,M)G(n,M) are a.a.s.  a set of acyclic graphs (a forest) together with a set of unicyclic graphs. We exploit this property in order to get the following result which refines the first statement in Theorem 1.1:

Theorem 3.1.

Let cc such that 0<c<1/20<c<1/2, and M=cnM=cn. Let L3L\subseteq\mathbb{N}_{\geqslant 3} and λL(z)=Lz2\lambda_{L}(z)=\sum_{\ell\in L}\frac{z^{\ell}}{2\ell}. Then the random variable Xn,M(L)X_{n,\,M}^{(L)} equal to the number of LL-cycles satisfies

Pr[Xn,M(L)=k]=eλL(2c)λL(2c)kk!(1+O(n1)).\Pr\left[X_{n,\,M}^{(L)}=k\right]=e^{-\lambda_{L}(2c)}\frac{\lambda_{L}(2c)^{k}}{k!}\left(1+O\left(n^{-1}\right)\right).

Moreover, if kk\to\infty as nn\to\infty then

Pr[Xn,M(L)=k]=O(kk).\Pr\left[X_{n,\,M}^{(L)}=k\right]=O(k^{-k}).
Proof.

It suffices to consider graphs whose connected components are trees and unicyclic graphs. Using the symbolic method we obtain that the probability that G(n,M)G(n,M) contains exactly kk unicyclic components containing an LL-cycle is equal to

Pr[Xn,M(L)=k]=n!((n2)M)[xn]W1(x)nM(nM)!λL(T(x))kk!eW0(x)λL(T(x)).\Pr\left[X_{n,\,M}^{(L)}=k\right]=\frac{n!}{\binom{\binom{n}{2}}{M}}[x^{n}]\frac{W_{-1}(x)^{n-M}}{(n-M)!}\frac{\lambda_{L}(T(x))^{k}}{k!}e^{W_{0}(x)-\lambda_{L}(T(x))}\,. (9)

The term λL(T(x))kk!\frac{\lambda_{L}(T(x))^{k}}{k!} encodes the components containing an LL-cycle, while the term eW0(x)λL(T(x))e^{W_{0}(x)-\lambda_{L}(T(x))} encodes the rest of unicyclic components (whose lengths do not belong to LL). Using Cauchy integral’s formula we get

[xn]W1(x)nMλL(T(x))keW0(x)λL(T(x))=\displaystyle[x^{n}]W_{-1}(x)^{n-M}\lambda_{L}(T(x))^{k}e^{W_{0}(x)-\lambda_{L}(T(x))}= (10)
2Mn2πi(2W1(x))nMλL(T(x))keW0(x)λL(T(x))dxxn+1.\displaystyle\frac{2^{M-n}}{2\pi i}\oint\left(2W_{-1}(x)\right)^{n-M}\lambda_{L}(T(x))^{k}e^{W_{0}(x)-\lambda_{L}(T(x))}\frac{dx}{x^{n+1}}.

After the change of variables z=T(x)z=T(x), it becomes

[xn]W1(x)nMλL(T(x))keW0(x)λL(T(x))=2Mn2πig(z)λL(z)kenh(z)dzz,[x^{n}]W_{-1}(x)^{n-M}\lambda_{L}(T(x))^{k}e^{W_{0}(x)-\lambda_{L}(T(x))}=\frac{2^{M-n}}{2\pi i}\oint g(z)\lambda_{L}(z)^{k}e^{nh(z)}\frac{dz}{z}\,, (11)

where

g(z)\displaystyle g(z) =\displaystyle= (1z)eλ(z)λL(z),\displaystyle(1-z)e^{\lambda(z)-\lambda_{L}(z)}, (12)
h(z)\displaystyle h(z) =\displaystyle= zlogz+(1Mn)log(2zz2).\displaystyle z-\log z+\left(1-\tfrac{M}{n}\right)\log\left(2z-z^{2}\right). (13)

Note that the function h(z)h(z) given by (13) is exactly the same as [2, Equation (30)], which satisfies the conditions h(2c)=h(1)=0h^{\prime}(2c)=h^{\prime}(1)=0. In the range M=cnM=cn with 0<c<120<c<\frac{1}{2}, we can apply saddle-point methods by choosing a circular path {2ceiθ,θ[π,π)}\{2ce^{i\theta},\theta\in[-\pi,\pi)\} as the contour of integration. As shown in [4], we split the integral in (11) into three parts, namely πθ0+θ0θ0+θ0π\int_{-\pi}^{-\theta_{0}}+\int_{-\theta_{0}}^{\theta_{0}}+\int_{\theta_{0}}^{\pi}. It suffices to integrate from θ0-\theta_{0} to θ0\theta_{0}, for a convenient value of θ0\theta_{0}, because the remaining integrals can be bounded by the magnitude of the central integrand. Following the proof of [2, Theorem 3.2] and choosing θ0=n2/5\theta_{0}=n^{-2/5} (so that nθ02n\theta_{0}^{2}\to\infty but nθ030n\theta_{0}^{3}\to 0 as nn\to\infty) we have

exp(nh(2ceiθ))=exp(nh(2c)nc(12c)2(1c)θ2)(1+iO(nθ3)+O(nθ4)),\exp\left(nh(2ce^{i\theta})\right)=\exp\left(nh(2c)-\tfrac{nc(1-2c)}{2(1-c)}\theta^{2}\right)\left(1+iO(n\theta^{3})+O(n\theta^{4})\right), (14)

and for all choices of θ\theta in [π,θ0][θ0,π)[-\pi,-\theta_{0}]\cup[\theta_{0},\pi) we have

|exp(nh(2ceiθ)nh(2c))|=exp(O(n1/5)).\left|\exp\left(nh(2ce^{i\theta})-nh(2c)\right)\right|=\exp\left(-O(n^{1/5})\right). (15)

As 2c<12c<1 in the vicinity of θ0\theta_{0}, we have

g(2ceiθ)=g(2c)(1+iO(θ)+O(θ2)),g\left(2ce^{i\theta}\right)=g(2c)\left(1+iO(\theta)+O(\theta^{2})\right), (16)

and

λL(2ceiθ)k=λL(2c)k(1+iO(θ)+O(θ2))\lambda_{L}(2ce^{i\theta})^{k}=\lambda_{L}(2c)^{k}\left(1+iO(\theta)+O(\theta^{2})\right) (17)

for fixed k0k\geqslant 0. Using expansions (14), (16), (17) and the bound (15) we have

g(z)λL(z)kenh(z)dzz=iθ0θ0g(2ceiθ)λL(2ceiθ)kenh(2ceiθ)𝑑θ(1+eO(n1/5))\oint g(z)\lambda_{L}(z)^{k}e^{nh(z)}\frac{dz}{z}=i\int_{-\theta_{0}}^{\theta_{0}}g(2ce^{i\theta})\lambda_{L}(2ce^{i\theta})^{k}e^{nh(2ce^{i\theta})}d\theta\left(1+e^{-O(n^{1/5})}\right)
=ig(2c)λL(2c)kenh(2c)θ0+θ0enσθ22(1+iO(θ)+O(θ2)+iO(nθ3)+O(nθ4))𝑑θ(1+eO(n1/5)),=ig(2c)\lambda_{L}(2c)^{k}e^{nh(2c)}\int_{-\theta_{0}}^{+\theta_{0}}e^{-n\sigma\tfrac{\theta^{2}}{2}}\cdot(1+iO(\theta)+O(\theta^{2})+iO(n\theta^{3})+O(n\theta^{4}))d\theta\left(1+e^{-O(n^{1/5})}\right),
g(z)λL(z)kenh(z)dzz=iθ0θ0g(2ceiθ)λL(2ceiθ)kenh(2ceiθ)𝑑θ(1+eO(n1/5))=ig(2c)λL(2c)kenh(2c)θ0+θ0enσθ22(1+iO(θ)+O(θ2)+iO(nθ3)+O(nθ4))𝑑θ(1+eO(n1/5)),\begin{split}\oint g(z)\lambda_{L}(z)^{k}e^{nh(z)}\frac{dz}{z}=i\int_{-\theta_{0}}^{\theta_{0}}g(2ce^{i\theta})\lambda_{L}(2ce^{i\theta})^{k}e^{nh(2ce^{i\theta})}d\theta\left(1+e^{-O(n^{1/5})}\right)\\ =ig(2c)\lambda_{L}(2c)^{k}e^{nh(2c)}\int_{-\theta_{0}}^{+\theta_{0}}e^{-n\sigma\tfrac{\theta^{2}}{2}}\cdot(1+iO(\theta)+O(\theta^{2})+iO(n\theta^{3})+O(n\theta^{4}))d\theta\left(1+e^{-O(n^{1/5})}\right),\end{split}

where σ=c(12c)1c\sigma=\,{\frac{c\left(1-2c\right)}{1-c}}. If we set x=nσθx=\sqrt{n\sigma}\thetathe integral in the above equation becomes

1nσσ1/2n1/10σ1/2n1/10ex22(1+iO(xnσ)+O(x2nσ)+iO(nx3nσ3)+O(x4nσ2))𝑑x.\frac{1}{\sqrt{n\sigma}}\int_{-\sigma^{1/2}n^{1/10}}^{\sigma^{1/2}n^{1/10}}e^{-\tfrac{x^{2}}{2}}\left(1+iO\left(\tfrac{x}{\sqrt{n\sigma}}\right)+O\left(\tfrac{x^{2}}{n\sigma}\right)+iO\left(n\tfrac{x^{3}}{\sqrt{n\sigma}^{3}}\right)+O\left(\tfrac{x^{4}}{n\sigma^{2}}\right)\right)dx. (18)

Observe that σ=O(1)\sigma=O(1) and the estimate (18) is a real number (because (10) is a real number). Hence (18) is equal to

1σnσ1/2n1/10σ1/2n1/10ex22(1+O(x2n)+O(x4n))𝑑x.\frac{1}{\sqrt{\sigma n}}\int_{-\sigma^{1/2}n^{1/10}}^{\sigma^{1/2}n^{1/10}}e^{-\tfrac{x^{2}}{2}}\left(1+O\left(\frac{x^{2}}{n}\right)+O\left(\frac{x^{4}}{n}\right)\right)dx.

It follows that

θ0θ0g(2ceiθ)λL(2ceiθ)kenh(2ceiθ)𝑑θ=2πσng(2c)λL(2c)kenh(2c)(1+O(n1)+eO(n1/5)).\int_{-\theta_{0}}^{\theta_{0}}g(2ce^{i\theta})\lambda_{L}(2ce^{i\theta})^{k}e^{nh(2ce^{i\theta})}d\theta=\sqrt{\frac{2\pi}{\sigma n}}g(2c)\lambda_{L}(2c)^{k}e^{nh(2c)}\left(1+O\left(n^{-1}\right)+e^{-O(n^{1/5})}\right).

That is

[xn]W1(x)nMλL(T(x))keW0(x)λL(T(x))=2Mn12πσng(2c)λL(2c)kenh(2c)(1+O(n1))[x^{n}]W_{-1}(x)^{n-M}\lambda_{L}(T(x))^{k}e^{W_{0}(x)-\lambda_{L}(T(x))}=2^{M-n}\frac{1}{\sqrt{2\pi\sigma n}}g(2c)\lambda_{L}(2c)^{k}e^{nh(2c)}\left(1+O\left(n^{-1}\right)\right) (19)

Using Stirling’s formula for the corresponding range of MM, we have

1((n2)M)n!(nM)!k!=1k!2πnMnM2MnnMMn2M(nM)nMexp(2M+Mn+M2n2)(1+O(n1)).\frac{1}{\binom{\binom{n}{2}}{M}}\frac{n!}{(n-M)!k!}=\frac{1}{k!}\sqrt{\frac{2\pi nM}{n-M}}\frac{2^{M}n^{n}M^{M}}{n^{2M}(n-M)^{n-M}}\exp\left(-2M+\frac{M}{n}+\frac{M^{2}}{n^{2}}\right)\left(1+O\left(n^{-1}\right)\right). (20)

Multiplying (19) and (20), after cancellations we obtain

Pr[Xn,M(L)=k]=eλL(2c)λL(2c)kk!(1+O(n1)).\Pr\left[X_{n,\,M}^{(L)}=k\right]=e^{-\lambda_{L}(2c)}\frac{\lambda_{L}(2c)^{k}}{k!}\left(1+O\left(n^{-1}\right)\right).

This proves the first part of the theorem.

Now, suppose that kk\to\infty as nn\to\infty. The previous arguments work in a similar way. Instead of using the estimate (17), which is only valid when θ\theta is small enough, we exploit the fact that λL\lambda_{L} has non-negative Taylor coefficients. Hence, Equation (17) can be replaced by the relation

|λL(2ceiθ)k|λL(2c)k,\left|\lambda_{L}(2ce^{i\theta})^{k}\right|\leqslant\lambda_{L}(2c)^{k},

which is valid for each choice of θ[π,π)\theta\in[-\pi,\pi). Applying the same arguments as before and that 1k!<ekkk\frac{1}{k!}<\frac{e^{k}}{k^{k}} for large kk, we conclude that

Pr[Xn,M(L)=k]eλL(2c)λL(2c)kk!(1+O(n1))<CeλL(2c)(eλL(2c))kkk,\Pr\left[X_{n,\,M}^{(L)}=k\right]\leqslant e^{-\lambda_{L}(2c)}\frac{\lambda_{L}(2c)^{k}}{k!}\left(1+O\left(n^{-1}\right)\right)<Ce^{-\lambda_{L}(2c)}\frac{(e\lambda_{L}(2c))^{k}}{k^{k}},

for a suitable constant CC. The second result in the theorem follows from the fact that c<12c<\frac{1}{2}, and hence λL(2c)\lambda_{L}(2c) is bounded. ∎

3.2. Barely subcritical regime

In the barely subcritical regime the asymptotic structure of G(n,M)G(n,M) is the same as in the subcritical regime. However, the integration countour we use is slightly more complicated in order to encode cycles of arbitrary length.

Theorem 3.2.

Let M=n2(1μn1/3)M=\frac{n}{2}(1-\mu n^{-1/3}) with μ\mu tending to infinity with μ=o(n1/3)\mu=o\left(n^{1/3}\right). Let L3L\subseteq\mathbb{N}_{\geqslant 3} and λL(z)=Lz2\lambda_{L}(z)=\sum_{\ell\in L}\frac{z^{\ell}}{2\ell}. Then the random variable Xn,M(L)X_{n,\,M}^{(L)} equal to the number of LL-cycles satisfies

Pr[Xn,M(L)=k]=eλL(2Mn)λL(2Mn)kk!(1+O(μ3)).\mathrm{Pr}\left[X_{n,M}^{(L)}=k\right]=e^{-\lambda_{L}\left(\frac{2M}{n}\right)}\frac{\lambda_{L}\left(\frac{2M}{n}\right)^{k}}{k!}\left(1+O\left(\mu^{-3}\right)\right). (21)

Assume moreover that limnλL(2Mn)=\lim_{n}\lambda_{L}\left(\frac{2M}{n}\right)=\infty. Then for fixed real numbers y0<y1y_{0}<y_{1}

Pr[y0Xn,MλL(2Mn)λL(2Mn)y1]12πy0y1eu2/2𝑑u,as n.\Pr\left[y_{0}\leqslant\frac{X_{n,\,M}-\lambda_{L}\left(\frac{2M}{n}\right)}{\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)}}\leqslant y_{1}\right]\to\frac{1}{\sqrt{2\pi}}\int_{y_{0}}^{y_{1}}e^{-u^{2}/2}du,\quad\hbox{as $n\to\infty$}. (22)
Proof.

The arguments and notation are similar to the ones in the proof of Theorem 3.1. As mentioned in the proof of Theorem 3.1, a.a.s. in this regime G(n,M)G(n,M) contains only trees and unicyclic graphs as components. We need estimates for (9) in this new range of MM. We use again the same methods as in the proof of [2, Theorem 3.2]. Let

ω(n)=(n2M)1/4n1/6,τ=n(nM)M(n2M),θ0=τnω(n).\omega(n)=\frac{(n-2M)^{1/4}}{n^{1/6}},\,\qquad\tau=\frac{n(n-M)}{M(n-2M)},\,\qquad\theta_{0}=\sqrt{\frac{\tau}{n}}\omega(n).

Then nθ2n\theta^{2}\to\infty and nθ30n\theta^{3}\to 0 as nn\to\infty. The expansion of hh in the vicinity of θ0\theta_{0} is

h(2Mneiθ)=h(2Mn)M(n2M)2n(nM)θ2i(n25nM+2M2)M6(nM)2θ3+O(θ4).h\left(\frac{2M}{n}e^{i\theta}\right)=h\left(\frac{2M}{n}\right)-\frac{M(n-2M)}{2n(n-M)}\theta^{2}-i\frac{(n^{2}-5nM+2M^{2})M}{6(n-M)^{2}}\theta^{3}+O(\theta^{4}). (23)

For θ[θ0,+θ0]\theta\in[-\theta_{0},+\theta_{0}], k=Θ(λL(2Mn))k=\Theta\left(\lambda_{L}(\frac{2M}{n})\right), the expansion of λL\lambda_{L} in the vicinity of θ0\theta_{0} is

λL(2Mneiθ)kλL(2Mn)k=1+iO(kλL(2Mn)n(n2M)θ)+O(k2λL(2Mn)2n2(n2M)2θ2)=1+iO(n(n2M)θ)+O(n2(n2M)2θ2).\begin{split}\frac{\lambda_{L}\left(\frac{2M}{n}e^{i\theta}\right)^{k}}{\lambda_{L}\left(\frac{2M}{n}\right)^{k}}&=1+iO\left(\frac{k}{\lambda_{L}\left(\frac{2M}{n}\right)}\frac{n}{(n-2M)}\theta\right)+O\left(\frac{k^{2}}{\lambda_{L}\left(\frac{2M}{n}\right)^{2}}\frac{n^{2}}{(n-2M)^{2}}\theta^{2}\right)\\ &=1+iO\left(\frac{n}{(n-2M)}\theta\right)+O\left(\frac{n^{2}}{(n-2M)^{2}}\theta^{2}\right).\end{split} (24)

The integrand can be bounded on [π,θ0)(θ0,π)[-\pi,-\theta_{0})\cup(\theta_{0},\pi) because

|exp(nh(2Mneiθ)nh(2Mn))|=O(eω(n)2/2).\left|\exp\left(nh\left(\frac{2M}{n}e^{i\theta}\right)-nh\left(\frac{2M}{n}\right)\right)\right|=O(e^{-\omega(n)^{2}/2}). (25)

Combining (23), (24) and (25), we have

Pr[Xn,M(L)=k]=n!((n2)M)(nM)!2Mn2πg(2Mn)exp(nh(2Mn))λL(2Mn)kk!×θ0θ0enτθ22(1+iO(n(n2M)θ)+O(n2(n2M)2θ2))×(1+in(n25nM+2M2)M6(nM)2θ3+O(nθ4))dθ(1+O(eω(n)2/2)).\begin{array}[]{ll}\mathrm{Pr}\left[X_{n,M}^{(L)}=k\right]=\frac{n!}{\binom{\binom{n}{2}}{M}(n-M)!}\frac{2^{M-n}}{2\pi}g\left(\frac{2M}{n}\right)\exp\left(nh\left(\frac{2M}{n}\right)\right)\frac{\lambda_{L}\left(\frac{2M}{n}\right)^{k}}{k!}\times\\ \int_{-\theta_{0}}^{\theta_{0}}e^{-n\tau\frac{\theta^{2}}{2}}\left(1+iO\left(\frac{n}{(n-2M)}\theta\right)+O\left(\frac{n^{2}}{(n-2M)^{2}}\theta^{2}\right)\right)\times\\ \left(1+in\frac{(n^{2}-5nM+2M^{2})M}{6(n-M)^{2}}\theta^{3}+O(n\theta^{4})\right)d\theta\left(1+O(e^{-\omega(n)^{2}/2})\right).\end{array}

We set θ=τ/nx\theta=\sqrt{{\tau}/{n}}x and the integral becomes

τnω(n)ω(n)ex22(1+iO(n(n2M)3/2x)+O(n2(n2M)3x2)).(1+iO(n(n2M)3/2x3)+O(n(n2M)3x4))dx=τnω(n)ω(n)ex22(1+O(n2(n2M)3x4))𝑑x=2πτn(1+O(n2(n2M)3)).\begin{split}\sqrt{\frac{\tau}{n}}&\int_{-\omega(n)}^{\omega(n)}e^{-\frac{x^{2}}{2}}\left(1+iO\left(\frac{n}{(n-2M)^{3/2}}x\right)+O\left(\frac{n^{2}}{(n-2M)^{3}}x^{2}\right)\right)\\ &\quad.\left(1+iO\left(\frac{n}{(n-2M)^{3/2}}x^{3}\right)+O\left(\frac{n}{(n-2M)^{3}}x^{4}\right)\right)dx\\ &=\sqrt{\frac{\tau}{n}}\int_{-\omega(n)}^{\omega(n)}e^{-\frac{x^{2}}{2}}\left(1+O\left(\frac{n^{2}}{(n-2M)^{3}}x^{4}\right)\right)dx\\ &=\sqrt{\frac{2\pi\tau}{n}}\left(1+O\left(\frac{n^{2}}{(n-2M)^{3}}\right)\right).\end{split}

After simple algebraic manipulations as in the proof of Theorem 3.1 we obtain

Pr[Xn,M(L)=k]=eλL(2Mn)λL(2Mn)kk!(1+O(μ3)).\mathrm{Pr}\left[X_{n,M}^{(L)}=k\right]=e^{-\lambda_{L}\left(\frac{2M}{n}\right)}\frac{\lambda_{L}\left(\frac{2M}{n}\right)^{k}}{k!}\left(1+O\left(\mu^{-3}\right)\right).

This proves the first part of the theorem.

We assume now that limnλL(2Mn)=\lim_{n}\lambda_{L}\left(\frac{2M}{n}\right)=\infty. Set k=λL(2Mn)+ρnλL(2Mn)k=\lambda_{L}\left(\frac{2M}{n}\right)+\rho_{n}\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)} with |ρn|=o(λL(2Mn))1/6|\rho_{n}|=o\left(\lambda_{L}\left(\frac{2M}{n}\right)\right)^{1/6}. We can apply Theorem 2.1 and obtain

Pr[Xn,M(L)=k]=12πλL(2Mn)eρn2/2(1+ρn3ρnλL(2Mn)+O(1+ρn3λL(2Mn)))=12πλL(2Mn)eρn2/2(1+o(1)).\begin{split}\Pr\left[X_{n,\,M}^{(L)}=k\right]&=\frac{1}{\sqrt{2\pi\lambda_{L}\left(\frac{2M}{n}\right)}}e^{-\rho_{n}^{2}/2}\left(1+\frac{\rho_{n}^{3}-\rho_{n}}{\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)}}+O\left(\frac{1+\rho_{n}^{3}}{\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)}}\right)\right)\,\\ &=\frac{1}{\sqrt{2\pi\lambda_{L}\left(\frac{2M}{n}\right)}}e^{-\rho_{n}^{2}/2}(1+o(1)).\end{split}

The central limit theorem for Xn,M(L)X_{n,M}^{(L)} follows, that is, for fixed real y0<y1y_{0}<y_{1} we have

Pr[y0Xn,M(L)λL(2Mn)λL(2Mn)y1]12πy0y1eu2/2𝑑u,as n.\Pr\left[y_{0}\leqslant\frac{X_{n,M}^{(L)}-\lambda_{L}\left(\frac{2M}{n}\right)}{\sqrt{\lambda_{L}\left(\frac{2M}{n}\right)}}\leqslant y_{1}\right]\to\frac{1}{\sqrt{2\pi}}\int_{y_{0}}^{y_{1}}e^{-u^{2}/2}du,\qquad\hbox{as $n\to\infty$}.

3.3. Critical regime

In this regime we have to take into account the appearance of complex components. Let pk(n,M;L,r)p_{k}(n,M;L,r) be the probability that G(n,M)G(n,M) has a total excess rr with exactly kk unicyclic components containing an LL-cycle. The following lemma gives an estimate for pk(n,M;L,r)p_{k}(n,M;L,r).

Lemma 3.3.

Let M=n2(1+μn1/3)M=\frac{n}{2}(1+\mu n^{-1/3}) with μ=O(1)\mu=O(1). Let α\alpha be the positive solution to μ=1αα\mu=\frac{1}{\alpha}-\alpha. Let

k=λL(eαn1/3)+ρλL(eαn1/3),k=\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)+\rho\sqrt{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)},

which satisfies ρ=ω(λL(eαn1/3)1/6)\rho=\omega\left(\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{1/6}\right).

Then for fixed rr we have

pk(n,M;L,r)=eλL(eαn1/3)λL(eαn1/3)kk!2πerA(3r+1/2,μ)(1+O(n1/12)),\begin{split}p_{k}(n,M;L,r)=e^{-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}\frac{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{k}}{k!}\sqrt{2\pi}e_{r}\;A(3r+1/2,\mu)\cdot\left(1+O\left(n^{-1/12}\right)\right),\end{split}

where

er=(6r)!25r32r(3r)!(2r)!,A(y,μ)=eμ3/63(y+1)/3k0(1232/3μ)kk!Γ((y+12k)/3).e_{r}=\frac{(6r)!}{2^{5r}3^{2r}(3r)!\,(2r)!},\,\,A(y,\mu)=\frac{e^{-\mu^{3}/6}}{3^{(y+1)/3}}\sum_{k\geqslant 0}\frac{(\frac{1}{2}3^{2/3}\mu)^{k}}{k!\Gamma((y+1-2k)/3)}.

Moreover, for rr large enough there exist absolute constants C>0C>0 and ε>0\varepsilon>0 such that

pk(n,M;L,r)eλL(eαn1/3)λL(eαn1/3)kk!Ceεr.p_{k}(n,M;L,r)\leqslant e^{-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}\frac{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{k}}{k!}Ce^{-\varepsilon r}. (26)
Proof.

The proof is based on analytic techniques introduced in [4] and [6]; see also [9]. The probability pk(n,M;L,r)p_{k}(n,M;L,r) is given by

pk(n,M;L,r)=n!((n2)M)[xn]W1(x)nM+r(nM+r)!Er(x)λL(T(x))kk!eW0(x)λL(T(x)),p_{k}(n,M;L,r)=\frac{n!}{\binom{\binom{n}{2}}{M}}[x^{n}]\frac{W_{-1}(x)^{n-M+r}}{(n-M+r)!}E_{r}(x)\frac{\lambda_{L}(T(x))^{k}}{k!}\,e^{W_{0}(x)-\lambda_{L}(T(x))}, (27)

where Er(x)E_{r}(x) is the EGF of complex components with total excess rr given by [6, Equation (6.8)]. As shown in [6], when r=o(n1/3)r=o(n^{1/3}), the series Er(x)E_{r}(x) can be approximated [6, Equation (6.8)] by er(1T(x))3r\frac{e_{r}}{(1-T(x))^{3r}}, where

er=(6r)!25r32r(3r)!(2r)!,e_{r}=\frac{(6r)!}{2^{5r}3^{2r}(3r)!\,(2r)!},

and the error term is of order O(r3/2n1/2)O\left(\frac{r^{3/2}}{n^{1/2}}\right). In order to evaluate (27) we have to compute the expression

St(n,M,r)2πi(1z)13renh1(z)λL(z)kk!eW0(z)λL(z)Er(z)dzz,\frac{St(n,M,r)}{2\pi i}\oint(1-z)^{1-3r}e^{nh_{1}(z)}\,\frac{\lambda_{L}(z)^{k}}{k!}\,e^{W_{0}(z)-\lambda_{L}(z)}\,E_{r}(z)\,\frac{dz}{z}, (28)

where

St(n,M,r)\displaystyle St(n,M,r) =\displaystyle= n!((n2)M)2n+Mrener(nM+r)!,\displaystyle\frac{n!}{\binom{\binom{n}{2}}{M}}\frac{2^{-n+M-r}e^{n}e_{r}}{(n-M+r)!}, (29)
h1(z)\displaystyle h_{1}(z) =\displaystyle= z1logz+(1Mn)log(2zz2).\displaystyle z-1-\log z+\left(1-\frac{M}{n}\right)\log(2z-z^{2})\,. (30)

We remark the difference between h1(z)h_{1}(z) and the function h(z)h(z) defined in Equation (13). Note also that h1(z)h_{1}(z) is exactly the same as in [6, Equation (10.12)], which satisfies h1(1)=h1(1)=0h_{1}(1)=h_{1}^{\prime}(1)=0 and also h1′′(1)=0h_{1}^{\prime\prime}(1)=0 if M=n/2M=n/2. We now follow the method of the proof of [6, Lemma 3] in order to compute our integral by choosing as path of integration

z=z(t)=eαn1/3itn1/2,z=z(t)=e^{-\alpha n^{-1/3}-itn^{-1/2}}, (31)

where α\alpha is the unique positive solution of μ=1αα\mu=\frac{1}{\alpha}-\alpha, and tt belongs to the interval

(πn1/4λL(eαn1/3)1/2,πn1/4λL(eαn1/3)1/2).\left(-\pi n^{1/4}\lambda_{L}^{\prime}(e^{-\alpha n^{-1/3}})^{-1/2},\pi n^{1/4}\lambda_{L}^{\prime}(e^{-\alpha n^{-1/3}})^{-1/2}\right).

Given that

λL(eαn1/3itn1/2)kk!eλL(eαn1/3itn1/2)=λL(eαn1/3)kk!eλL(eαn1/3)×(1+iO(kλL(eαn1/3)λL(eαn1/3)n1/2t)+O(k2λL(eαn1/3)2λL(eαn1/3)2nt2))\begin{split}&\frac{\lambda_{L}(e^{-\alpha n^{-1/3}-itn^{-1/2}})^{k}}{k!}e^{-\lambda_{L}(e^{-\alpha n^{-1/3}-itn^{-1/2}})}=\\ &\frac{\lambda_{L}(e^{-\alpha n^{-1/3}})^{k}}{k!}e^{-\lambda_{L}(e^{-\alpha n^{-1/3}})}\qquad\times\\ &\left(1+iO\left(\frac{k}{\lambda_{L}(e^{-\alpha n^{-1/3}})}\frac{\lambda_{L}^{\prime}\left(e^{-\alpha n^{-1/3}}\right)}{n^{1/2}}t\right)+O\left(\frac{k^{2}}{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{2}}\frac{\lambda_{L}^{\prime}\left(e^{-\alpha n^{-1/3}}\right)^{2}}{n}t^{2}\right)\right)\end{split} (32)

as long as k=O(λL(eαn1/3))k=O\left(\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)\right), our choice ensures that the OO terms in (32) can be moved out of the integral. By following the proof of [6, Equation (10.1) of Lemma 3] we obtain that, for fixed values of rr

pk(n,M;L,r)=eλL(eαn1/3)λL(eαn1/3)kk!2πerA(3r+1/2,μ)(1+O(n1/12)+O(μ4n1/3)).p_{k}(n,M;L,r)=e^{-\lambda_{L}(e^{-\alpha n^{-1/3}})}\frac{\lambda_{L}(e^{-\alpha n^{-1/3}})^{k}}{k!}\sqrt{2\pi}e_{r}A(3r+1/2,\mu)\left(1+O\left(n^{-1/12}\right)+O\left(\frac{\mu^{4}}{n^{1/3}}\right)\right). (33)

Since k=O(λL(eαn1/3))k=O\left(\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)\right) in the OO terms above we have

λL(eαn1/3)n1/2t=O(λL(eαn1/3)1/2n1/4)=O(1(1eαn1/3)1/21n1/4)=O(n1/12).\frac{\lambda_{L}^{\prime}\left(e^{-\alpha n^{-1/3}}\right)}{n^{1/2}}t=O\left(\frac{\lambda_{L}^{\prime}\left(e^{-\alpha n^{-1/3}}\right)^{1/2}}{n^{1/4}}\right)=O\left(\frac{1}{(1-e^{-\alpha n^{-1/3}})^{1/2}}\frac{1}{n^{1/4}}\right)=O\left(n^{-1/12}\right).

This proves the first statement of the theorem.

Next let us assume that rr\to\infty. We know that Er(z)er(1T(z))3rE_{r}(z)\preceq\frac{e_{r}}{(1-T(z))^{3r}} (see for instance [6, Lemma 4]). From (27) we have

pk(n,M;L,r)n!((n2)M)[zn]W1(x)nM+r(nM+r)!λL(T(z))kk!eW0(z)λL(T(z))er(1T(z))3r.p_{k}(n,M;L,r)\leqslant\frac{n!}{\binom{\binom{n}{2}}{M}}[z^{n}]\frac{W_{-1}(x)^{n-M+r}}{(n-M+r)!}\frac{\lambda_{L}(T(z))^{k}}{k!}\,e^{W_{0}(z)-\lambda_{L}(T(z))}\frac{e_{r}}{(1-T(z))^{3r}}.

Then, we obtain

pk(n,M;L,r)St(n,M,r)2πizr(2z)r(1z)3renh1(z)λL(z)kk!eλ(z)λL(z)dzz,p_{k}(n,M;L,r)\leqslant\frac{St(n,M,r)}{2\pi i}\oint\frac{z^{r}(2-z)^{r}}{(1-z)^{3r}}e^{nh_{1}(z)}\frac{\lambda_{L}(z)^{k}}{k!}e^{\lambda(z)-\lambda_{L}(z)}\frac{dz}{z}, (34)

with h1h_{1} is defined by (30). In this case we take as contour of integration the circle {δeiθ:θ[π,π)}\{\delta e^{i\theta}:\theta\in[-\pi,\pi)\} with δ=1r1/3n1/3<1\delta=1-\frac{r^{1/3}}{n^{1/3}}<1. On this circle, since r1r\geqslant 1, for some constant CC and function f(n)f(n) with limnf(n)=+\lim_{n}f(n)=+\infty, we have

λL(δ)kk!eλL(δ)δ2π(δ(2δ)(1δ)3)renh1(δ)(1δ)1/2ππef(n)θ2𝑑θ<Cnδr(δ(2δ)(1δ)3)renh1(δ).\begin{split}\frac{\lambda_{L}(\delta)^{k}}{k!}e^{-\lambda_{L}(\delta)}&\frac{\delta}{2\pi}\left(\frac{\delta(2-\delta)}{(1-\delta)^{3}}\right)^{r}e^{nh_{1}(\delta)}(1-\delta)^{1/2}\int_{-\pi}^{\pi}e^{-f(n)\theta^{2}}d\theta<\frac{C}{\sqrt{n}}\,\delta^{r}\left(\frac{\delta(2-\delta)}{(1-\delta)^{3}}\right)^{r}e^{nh_{1}(\delta)}.\end{split}

Note that rM=n2(1+μn1/3)r\leqslant M=\frac{n}{2}(1+\mu n^{-1/3}). Then for nn large enough

δr(2δ)r(1δ)3r<nrrr,nh1(δ)<1312r+116μr2/3.\frac{\delta^{r}(2-\delta)^{r}}{(1-\delta)^{3r}}<\frac{n^{r}}{r^{r}},\,nh_{1}(\delta)<\frac{13}{12}r+\frac{11}{6}\mu r^{2/3}. (35)

Using Stirling’s formula we find that

n!((n2)M)(nM+r)!en2n+Mr<n1/2nreμ3/6+3/42r,\frac{n!}{\binom{\binom{n}{2}}{M}(n-M+r)!}e^{n}2^{-n+M-r}<\frac{n^{1/2}}{n^{r}}e^{-\mu^{3}/6+3/4}2^{-r},

and for rr\to\infty, we have

er=(6r)!25r32r(3r)!(2r)!1r1/2(3r2e)r.e_{r}=\frac{(6r)!}{2^{5r}3^{2r}(3r)!\,(2r)!}\leqslant\frac{1}{r^{1/2}}\left(\frac{3r}{2e}\right)^{r}. (36)

Combining (35) and (36) in (34), we deduce that

pk(n,M;L,r)<c0r1/2exp(μ36+116μr2/3+(1312+log341)r),p_{k}(n,M;L,r)<\frac{c_{0}}{r^{1/2}}\exp\left(-\frac{\mu^{3}}{6}+\frac{11}{6}\mu r^{2/3}+\left(\frac{13}{12}+\log\frac{3}{4}-1\right)r\right),

for some constant c0>0c_{0}>0. Since 1312+log341<0\frac{13}{12}+\log\frac{3}{4}-1<0, when rr\to\infty we deduce

pk(n,M;L,r)eO(r).p_{k}(n,M;L,r)\leqslant e^{-O(r)}.

We can now proof the main result in the critical regime.

Theorem 3.4.

Let M=n2(1+μn1/3)M=\frac{n}{2}(1+\mu n^{-1/3}) where μ=O(1)\mu=O(1). Let α\alpha be the positive solution to μ=1αα\mu=\frac{1}{\alpha}-\alpha. Let L3L\subseteq\mathbb{N}_{\geqslant 3} and let λL(z)=kLzk2k\lambda_{L}(z)=\sum_{k\in L}\frac{z^{k}}{2k}. Then the random variable Xn,M(L)X_{n,\,M}^{(L)} equal LL-cycles in G(n,M)G(n,M) satisfies

Pr[Xn,M(L)=k]=eλL(eαn1/3)λL(eαn1/3)kk!(1+O(n1/12)).\mathrm{Pr}\left[X_{n,M}^{(L)}=k\right]=e^{-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}\frac{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{k}}{k!}\left(1+O\left(n^{-1/12}\right)\right). (37)

Moreover, assume that limnλL(eαn1/3)=+\lim_{n\to\infty}\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)=+\infty. Then, for each choice of real y0<y1y_{0}<y_{1}

Pr[y0Xn,M(L)λL(eαn1/3)λL(eαn1/3)y1]12πy0y1eu2/2𝑑u, when n.\Pr\left[y_{0}\leqslant\frac{X_{n,M}^{(L)}-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}{\sqrt{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}}\leqslant y_{1}\right]\to\frac{1}{\sqrt{2\pi}}\int_{y_{0}}^{y_{1}}e^{-u^{2}/2}du,\qquad\hbox{ when }n\to\infty. (38)
Proof.

By Lemma 3.3 and the dominated convergence theorem, Pr[Xn,M(L)=k]\Pr\left[X_{n,\,M}^{(L)}=k\right] is equal to

r0pk(n,M;L,r)=r0eλL(eαn1/3)λL(eαn1/3)kk!2πerA(3r+1/2,μ)(1+O(n1/12)).\begin{split}\sum_{r\geqslant 0}p_{k}(n,M;L,r)=\sum_{r\geqslant 0}e^{-\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)}\frac{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{k}}{k!}\sqrt{2\pi}e_{r}\;A(3r+1/2,\mu)\cdot\left(1+O\left(n^{-1/12}\right)\right).\end{split}

For μ=O(1)\mu=O(1) Janson, Knuth, Łuczak and Pittel [6, Equation (13.17) and Corollary p. 61] have shown that the probability that G(n,M)G(n,\,M) has total excess rr is asymptotically 2πerA(3r+1/2,μ)\sqrt{2\pi}e_{r}A(3r+1/2,\mu), and that the ss-th moment of the excess rr satisfies r02πerrsA(3r+1/2,μ)=O(μ3s)=O(1)\sum_{r\geqslant 0}\sqrt{2\pi}e_{r}r^{s}A(3r+1/2,\mu)=O(\mu^{3s})=O(1). Hence r02πerA(3r+1/2,μ)=1\sum_{r\geqslant 0}\sqrt{2\pi}e_{r}A(3r+1/2,\mu)=1. This shows relation (37).

In order to prove (38), we apply Theorem 2.1 by choosing

k=λL(eαn1/3)+ρnλL(eαn1/3),k=\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)+\rho_{n}\sqrt{\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)},

with λL(eαn1/3)1/6=o(ρn).\lambda_{L}\left(e^{-\alpha n^{-1/3}}\right)^{1/6}=o(\rho_{n}).

4. Acknowledgments

Part of this research was done when the J.R. was visiting IRIF – Université Paris Denis Diderot in July 2018. J.R. thanks the hospitality of the institution during his research stay. V.R. was partially supported by grants ANR 2010 BLAN 0204 (Magnum). V.R. and J.R. were supported by the Project PHC Procope ID-57134837 ‘Analytic, probabilistic and geometric Methods for Random Constrained graphs’. J.R. was partially supported by the FP7-PEOPLE-2013-CIG project CountGraph (ref. 630749). Finally, M.N. and J.R. were supported by the Spanish project MTM2017-82166-P and by the H2020-MSCA-RISE-2020 project RandNET (ref. 101007705).

References

  • [1] N. Alon and J. H. Spencer. The probabilistic method. Wiley-Interscience Series in Discrete Mathematics and Optimization. Third edition.
  • [2] H. Daudé and V. Ravelomanana. Random 2-xorsat phase transition. Algorithmica, Special issue of LATIN 2008:1–18, 2009.
  • [3] P. Erdős and A. Rényi. On the evolution of random graphs. Publ. Math. Inst. Hung. Acad. Sci, 5:17–61, 1960.
  • [4] P. Flajolet, D. E. Knuth, and B. Pittel. The first cycles in an evolving graph. Discrete Mathematics, 75(1-3):167–215, 1989.
  • [5] P. Flajolet and B. Sedgewick. Analytic Combinatorics. Cambridge University Press, 2009.
  • [6] S. Janson, D. E. Knuth, T. Łuczak, and B. Pittel. The Birth of the Giant Component. Random Structures and Algorithms, 4(3):233–358, 1993.
  • [7] S. Janson, T. Łuczak, and A. Ruciński. Random Graphs. Wiley-Interscience, 2000.
  • [8] V. F. Kolchin. Random Graphs. Cambridge University Press, 1998.
  • [9] Marc Noy, Vlady Ravelomanana, and Juanjo Rué. On the probability of planarity of a random graph near the critical point. Proc. Amer. Math. Soc., 143(3):925–936, 2015.