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

11institutetext: Claude Godrèche 22institutetext: Université Paris-Saclay, CNRS, CEA, Institut de Physique Théorique, 91191 Gif-sur-Yvette, France
22email: [email protected]

Condensation and extremes for a fluctuating number of independent random variables

Claude Godrèche
(Received: date / Accepted: date)
Abstract

We address the question of condensation and extremes for three classes of intimately related stochastic processes: (a) random allocation models and zero-range processes, (b) tied-down renewal processes, (c) free renewal processes. While for the former class the number of components of the system is fixed, for the two other classes it is a fluctuating quantity. Studies of these topics are scattered in the literature and usually dressed up in other clothing. We give a stripped-down account of the subject in the language of sums of independent random variables in order to free ourselves of the consideration of particular models and highlight the essentials. Besides giving a unified presentation of the theory, this work investigates facets so far unexplored in previous studies. Specifically, we show how the study of the class of random allocation models and zero-range processes can serve as a backdrop for the study of the two other classes of processes central to the present work—tied-down and free renewal processes. We then present new insights on the extreme value statistics of these three classes of processes which allow a deeper understanding of the mechanism of condensation and the quantitative analysis of the fluctuations of the condensate.

Keywords:
Condensationextremesrenewal processeszero-range processessubexponentiality
journal: Journal of Statistical Physics

1 Introduction

It is well known that nn independent and identically distributed (iid) positive random variables conditioned by an atypical value of their sum exhibit the phenomenon of condensation, whereby one of the summands dominates upon the others, when their common distribution is subexponential (decaying more slowly than an exponential at large values of its argument).

Let X1,X2,,XnX_{1},X_{2},\dots,X_{n} be these nn random variables, henceforth taken discrete with positive integer values, whose common distribution is denoted by f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k). Hereafter we consider the particular case of a subexponential distribution with asymptotic power-law decay111 In the further course of this work, the symbol \approx stands for asymptotic equivalence; the symbol \sim means either ‘of the order of’, or ‘with exponential accuracy’, depending on the context.

f(k)ck1+θ,f(k)\approx\frac{c}{k^{1+{\theta}}}, (1.1)

where the index θ{\theta} and the tail parameter cc are both positive. Assume—for the time being—that the first moment X\langle X\rangle is finite (hence θ>1{\theta}>1) and that the sum of these random variables,

Sn=i=1nXi,S_{n}=\sum_{i=1}^{n}X_{i}, (1.2)

is conditioned to take the atypically large value L>Sn=nXL>\langle S_{n}\rangle=n\langle X\rangle. In the present context the phenomenon of condensation has a simple pictorial representation. Let us consider the partial sums S1,S2,S_{1},S_{2},\dots as the successive positions of a random walk whose steps are the summands XiX_{i}. Such a representation is used in figure 1, which depicts six different paths of this walk, conditioned by a large, atypical value of its final position SnS_{n} after nn steps. The steps have distribution (1.1) with θ=3/2{\theta}=3/2 (see the caption for details). As can be seen on this figure, for most of the paths there is a single big step bearing the excess difference Δ=LnX\Delta=L-n\langle X\rangle. In the thermodynamic limit, LL\to\infty, nn\to\infty with ρ=L/n\rho=L/n fixed, the distribution of the size of this big step becomes narrow around Δ\Delta. This picture gives the gist of the phenomenon of condensation.

The two ingredients responsible for such a phenomenon are: (i) subexponentiality of the distribution f(k)f(k), and: (ii) conditioning by an atypically large value of the sum SnS_{n}. In contrast, keeping the same distribution (1.1), but conditioning the sum SnS_{n} to be less than or equal to nXn\langle X\rangle, yields a ‘democratic’ situation where all the steps XiX_{i} are on the same footing, sharing the—now negative—excess difference Δ\Delta. Otherwise stated, in this situation, the path responds ‘elastically’ to the conditioning.

When the distribution f(k)f(k) is decaying exponentially (e.g., a geometric distribution) the paths responds ‘elastically’ in all cases, i.e., regardless of whether the sum SnS_{n} is conditioned to take an atypical value larger or smaller than nXn\langle X\rangle. In both cases all the steps XiX_{i} are on the same footing, sharing the excess difference Δ\Delta (which is now either positive or negative).

Refer to caption
Figure 1: Pictorial illustration of the phenomenon of condensation for the random allocation models and zrp class. This figure depicts six paths of a random walk whose positions are given by the partial sums S1,S2,S_{1},S_{2},\dots. The distribution of the steps XiX_{i} is given by f(k)=3k5/2/2f(k)=3k^{-5/2}/2 (k>1k>1 is taken continuous), for which X=3\langle X\rangle=3. The random walk is conditioned to end at position L=6000L=6000 at time n=500n=500. For most of the paths one can observe the occurrence of a big step whose magnitude fluctuates around Δ=LSn=4500\Delta=L-\langle S_{n}\rangle=4500. More rarely, the excess Δ\Delta is shared by two big steps, as for the green path.

This scenario of condensation has been investigated in great detail and is basically understood. It is for example encountered in random allocation models where nn sites (or boxes) contain altogether LL particles, the XiX_{i} representing the occupations of these sites burda ; burda2 ; janson . This situation in turn accounts for the stationary state of dynamical urn models such as zero range processes (zrp) or variants spitzer ; andjel ; camia ; evans2000 ; jeon ; gl2002 ; cg2003 ; gross ; hanney ; gl2005 ; lux ; maj2 ; ferrari ; maj3 ; armendariz2009 ; armendariz2011 ; armendariz2013 ; armendariz2017 ; cg2019 . For short we shall refer to this class of models as the class of random allocation models and zrp. Condensation means that one of the sites contains a macroscopic fraction of all the particles.

Refer to caption
Figure 2: In a free renewal process, the number NLN_{L} of intervals before LL is a random variable such that SNL<L<SNL+1S_{N_{L}}<L<S_{N_{L+1}}, where SNLS_{N_{L}} is the sum of the NLN_{L} intervals X1,X2,,XNLX_{1},X_{2},\dots,X_{N_{L}}. The last interval BL=LSNLB_{L}=L-S_{N_{L}} is unfinished. For a tied-down renewal process BL=0B_{L}=0, SNL=LS_{N_{L}}=L.

Here we shall be concerned by a different situation where the number of random variables X1,X2,,X_{1},X_{2},\dots, is itself a random variable, henceforth denoted by NLN_{L} and defined by conditioning the sum,

SNL=i=1NLXi,S_{N_{L}}=\sum_{i=1}^{N_{L}}X_{i}, (1.3)

to satisfy either the inequality

SNL<L<SNL+1,S_{N_{L}}<L<S_{N_{L}+1}, (1.4)

or the equality

SNL=L,S_{N_{L}}=L, (1.5)

where LL is a given positive integer number, see figure 2. These conditions are imposed irrespectively of whether the mean X\langle X\rangle is finite or not.

The process conditioned by the inequality (1.4) defines a free renewal process feller ; doob ; smith ; cox , the process conditioned by the equality (1.5) defines a tied-down renewal process wendel ; wendel1 ; wendel2 . In the former case the process is pinned at the origin, in the latter case it is also pinned at the end point. For both, the random variables XiX_{i} are the sizes of the iid (spatial or temporal) intervals between two renewals. Using the temporal language, the sum SNLS_{N_{L}} is the time of occurrence of the last renewal before or at time LL. The last, unfinished, interval BL=LSNLB_{L}=L-S_{N_{L}} is known as the backward recurrence time in renewal theory. Tied-down renewal processes (tdrp) are special because the pinning condition (1.5) imposes BL=0B_{L}=0.

A simple implementation of a tdrp is provided by the Bernoulli bridge, or tied-down random walk, made of ±1\pm 1 steps, starting from the origin and ending at the origin at time 2L2L wendel ; wendel1 ; labarbe . The sizes of the intervals between the successive passages by the origin of the walk (where each tick mark on the xx-axis represents two units of time) represent the random variables XiX_{i}, as depicted in figure 3. The continuum limit of the tied-down random walk is the Brownian bridge, also known as tied-down Brownian motion or else pinned Brownian motion verv .

Refer to caption
Figure 3: A tied-down random walk, or Bernoulli bridge, is a simple random walk, with steps ±1\pm 1, starting and ending at the origin. Time is along the xx-axis, space along the yy-axis. The tick marks on the xx-axis correspond to two time-steps. In this example the walk is made of L=15L=15 tick marks, with N15=5N_{15}=5 intervals between zeros, X1,,X5X_{1},\dots,X_{5}, taking the values 1,3,9,1,11,3,9,1,1 tick marks, respectively. The distribution of the sizes of the intervals, f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k), is given by (5.1).

For renewal processes (both free or tied-down) with a subexponential distribution f(k)f(k), we shall show that, by weighting the configurations according to the number NLN_{L} of summands, a phase transition occurs, as LL\to\infty, when the positive weight parameter ww conjugate to NLN_{L}, varies from larger values, favouring configurations with a large number of summands, to smaller ones, favouring atypical configurations with a smaller number of summands. In the former case the weight parameter ww is to be interpreted as a reward, in the latter case as a penalty.

Characterising this transition is the aim of the present work, with main focus on the quantitative analysis of the fluctuations of the condensate. Again, the occurrence of the phenomenon of condensation is due to: (i) subexponentiality of the distribution f(k)f(k), and: (ii) atypicality of the configurations.

Tied-down renewal processes fall into the class of linear systems considered by Fisher fisher . The latter are defined as one-dimensional chains of total length LL, made up, e.g., of alternating intervals of two kinds, AA and BB. This class encompasses the Poland-Scheraga model poland ; poland2 , consisting of an alternating sequence of straight paths AA and loops BB, wetting models, where AA and BB represent two phases, etc. If the direction of the chain is taken as a time axis, the loops in the perpendicular direction can be seen as random walks. The Bernoulli bridge or tied-down random walk of figure 3 is a natural implementation of this situation, where there is only one kind of intervals, say the loops BB, representing the intervals between two passages at the origin of the walk. In the same vein, a variant of the random allocation model defined in burda considers the case where the number of sites is varying burda3 , with an occupation variable XX starting at k=1k=1. This model, as well as the spin domain model considered in bar2 ; bar3 ; barma are examples of linear systems with only one kind of intervals. Both models are actually equivalent and are just particular instances of the tdrp considered in the present work. Let us finally mention the random walk (or polymer) models considered in gia1 ; gia2 which are free or tied-down renewal processes with a penalty (or reward) at each renewal events, as in the present work. In these models the condensation transition is interpreted as the transition between a localised phase and a delocalised one gia1 ; gia2 . For the example of figure 3, a localised configuration corresponds to many contacts of the walk with the origin, while a delocalised one corresponds to the presence of a macroscopic excursion.

Renewal theory is a classic in probability studies. It is also ubiquitous in statistical physics and has a wide range of applicability (see examples in gl2001 ; gms2015 ; cohen ; barkai2003 ). Yet, besides gia1 ; gia2 , which give rigorous mathematical results, studies of weighted free renewal processes are scarce. In particular, there are no available detailed characterisation of the condensation phenomenon for these processes in the existing literature, nor considerations on the statistics of extremes in the condensed phase.

We now describe the organisation of the paper, with further details on the literature on the subject.

The text is composed of three parts, which are respectively sections 2 and 3, dealing with the class of random allocation models and zrp, sections 4 to 8, dealing with tdrp, and finally, sections 9 to 11, dealing with free renewal processes. These three parts are both conceptually and analytically related.

Section 2 is a short presentation of how condensation arises for random allocation models and zrp. Subsection 2.1 gives the basic formalism. Equation (2.20) introduces a simple remark on the distribution of the maximum in the second half (k>L/2k>L/2), which will turn out to be instrumental in section 3, and which will be generalised in the later sections on renewal processes. Subsection 2.2 summarises the main features of the phenomenon of condensation in the thermodynamic limit. This classical topic has been much investigated in the past, both in statistical physics burda ; burda2 ; camia ; evans2000 ; gl2002 ; cg2003 ; gross ; hanney ; gl2005 ; lux ; maj2 ; maj3 and in mathematics jeon ; ferrari ; armendariz2009 ; armendariz2011 ; armendariz2013 ; armendariz2017 ; landim . The summary given in this section relies on the short review cg2019 to which we refer the reader for further bibliographical references. A mathematical review of a number of aspects of the subject and related matters can be found in janson .

Section 3, which will serve as a backdrop for the study contained in the two other parts, contains novel aspects of the phenomenon of condensation for the class of models at hand. Namely, instead of considering the thermodynamic limit LL\to\infty, nn\to\infty with fixed ratio ρ=L/n\rho=L/n, we investigate the situation where the number of summands nn is kept fixed and the value of the sum Sn=LS_{n}=L increases to infinity. Such a framework is precisely that considered in ferrari . In this reference it is shown that for nn fixed, LL\to\infty, assuming that ρc=X\rho_{c}=\langle X\rangle exists, i.e., θ>1{\theta}>1, if the largest summand is removed, the measure on the remaining summands converges to the product measure with density ρc\rho_{c}, a feature which is apparent on figure 1. Therefore the largest summand is the unique condensate, with size L(n1)XL-(n-1)\langle X\rangle. As emphasised in ferrari this phenomenon is a combinatorial fact that can be observed without making the number of sites grow to infinity’. Simply stated, there is total condensation in this case, in the sense that the condensate essentially bears the totality of the LL particles. We shall prove this result by elementary means (see (3.8)) and extend it to the case where θ<1{\theta}<1, thus proving that even though the first moment ρc=X\rho_{c}=\langle X\rangle does not exist, there is however (total) condensation222The fact that condensation also occurs when θ<1{\theta}<1 in the present context has been previously mentioned in landim . I am indebted to S Grosskinsky for pointing this reference to me.. If θ<1{\theta}<1, the correction of the mean largest summand to LL—in other words the fluctuations of the condensate—scales as L1θL^{1-{\theta}}, with a known amplitude, given in (3.6).

It turns out that this scenario of (total) condensation is precisely that prevailing for the two other processes studied in the present work, namely, tied-down and free renewal processes. This scenario will be the red thread for the rest of the paper, with all the complications introduced by a now fluctuating number of summands. This red thread in particular links the three figures 5, 9 and 11, and equations (3.6), (3.8), (8.31), (8.32), (11.27) and (11.28), which are the central results of the present study.

Sections 4 to 8 are devoted to tdrp with a power-law distribution of summands (1.1). Section 4 gives a systematic presentation of the formalism, valid for any arbitrary distribution f(k)f(k), followed, in sections 6 to 8 by the analysis of the behaviour of the system in the different phases, when f(k)f(k) is a subexponential distribution of the form (1.1). As said in the abstract, these topics are scattered in the literature burda3 ; bar2 ; bar3 ; barma . The analyses presented in sections 6 to 8 are comprehensive and go deeper than previous studies, especially in the description of the condensation phenomenon and in the analysis of the statistics of extremes, as detailed in section 8.

Likewise, sections 9 to 11, devoted to free renewal processes with power-law distribution of summands (1.1), give a thorough analysis both of the formalism and of the phenomenon of condensation. It turns out that this case, which is more generic than that of tdrp since the process is not pinned at the end point LL, is yet more complicated to analyse, the reason being that besides the intervals XiX_{i}, the last interval BLB_{L}, depicted in figure 2, enters the analysis.

Section 12, together with tables 1 and 2, gives a summary of the present study.

2 Condensation for random allocation models and ZRP

The key quantity for the study of the processes described above (random allocation models and zrp, free and tied-down renewal processes) is the statistical weight of a configuration, or in the language of the random walk of figure 1, the statistical weight of a path.

The choice of conventions on the initial values of nn, LL, kk is a matter of convenience which depends on the kind of reality that we wish to describe, as will appear shortly.

We start by giving, in §2.1, some elements of the formalism for random allocation models and zrp, where f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k) is any arbitrary distribution of the positive random variable XX. In the rest of section 2, f(k)f(k) has a power-law tail (1.1).

2.1 General formalism

2.1.1 Statistical weight of a configuration

Let X1,X2,,XnX_{1},X_{2},\dots,X_{n} be nn positive iid integer random variables with sum SnS_{n} conditioned to be equal to LL. The joint conditional probability associated to a configuration {X1=k1,,Xn=kn}\{X_{1}=k_{1},\dots,X_{n}=k_{n}\}, with Sn=LS_{n}=L given, reads333For the sake of simplicity we restrict the study to the case where f(k)f(k) is a normalisable probability distribution.

p(k1,,kn|L)\displaystyle p(k_{1},\dots,k_{n}|L) =\displaystyle= Prob(X1=k1,,Xn=kn|Sn=L)\displaystyle\mathop{\rm Prob}\nolimits(X_{1}=k_{1},\dots,X_{n}=k_{n}|S_{n}=L) (2.1)
=\displaystyle= 1Zn(L)f(k1)f(kn)δ(i=1nki,L),\displaystyle\frac{1}{Z_{n}(L)}f(k_{1})\ldots f(k_{n})\,\delta\Big{(}\sum_{i=1}^{n}k_{i},L\Big{)},

where δ(.,.)\delta(.,.) is the Kronecker delta, and where the denominator, whose presence stems from the constraint Sn=LS_{n}=L, is the partition function

Zn(L)\displaystyle Z_{n}(L) =\displaystyle= {ki}f(k1)f(kn)δ(i=1nki,L)=(f)n(L)\displaystyle\sum_{\{k_{i}\}}f(k_{1})\dots f(k_{n})\delta\Big{(}\sum_{i=1}^{n}k_{i},L\Big{)}=(f\star)^{n}(L) (2.2)
=\displaystyle= δ(Sn,L)=Prob(Sn=L).\displaystyle\langle\delta(S_{n},L)\rangle=\mathop{\rm Prob}\nolimits(S_{n}=L).

So Zn(L)Z_{n}(L) is another notation for the distribution of the sum SnS_{n}.

The physical picture associated to these definitions correspond to a system of nn sites (or boxes), LL particles in total, and where the summands XiX_{i} are the occupation numbers of these sites, i.e., the number of particles on each of them. Since these sites can be empty, the occupation probability f(0)f(0) is non zero in general. It is therefore natural to initialise LL to 0. In particular, the probability that the sum of occupations be zero is that all sites are empty, i.e.,

Zn(0)=f(0)n.Z_{n}(0)=f(0)^{n}. (2.3)

It also turns out to be convenient to start nn at 0, and set

Z0(L)=δ(L,0),Z_{0}(L)=\delta(L,0), (2.4)

which serves as an initial condition for the recursion

Zn(L)=k=0Lf(k)Zn1(Lk).Z_{n}(L)=\sum_{k=0}^{L}f(k)Z_{n-1}(L-k). (2.5)

As can be seen either from (2.2) or (2.5), the generating function of Zn(L)Z_{n}(L) with respect to LL yields

Z~n(z)=L0zLZn(L)=f~(z)n,\tilde{Z}_{n}(z)=\sum_{L\geq 0}z^{L}Z_{n}(L)=\tilde{f}(z)^{n}, (2.6)

where the generating function of f(k)f(k) with respect to kk is

f~(z)=k0zkf(k).\tilde{f}(z)=\sum_{k\geq 0}z^{k}f(k). (2.7)

The marginal distribution of the occupation of a generic site, say site 11, is

πn(k|L)=Prob(X1=k|Sn=L)=δ(X1,k)=f(k)Zn1(Lk)Zn(L),\pi_{n}(k|L)=\mathop{\rm Prob}\nolimits(X_{1}=k|S_{n}=L)=\langle\delta(X_{1},k)\rangle=\frac{f(k)Z_{n-1}(L-k)}{Z_{n}(L)}, (2.8)

where \langle\cdot\rangle is the average with respect to (2.1). The mean conditional occupation is

X|L=k0kπn(k|L)=Ln=ρ.\langle X|L\rangle=\sum_{k\geq 0}k\pi_{n}(k|L)=\frac{L}{n}=\rho. (2.9)

2.1.2 Distribution of the largest occupation

Condensation corresponds to the presence of a site with a macroscopic occupation. We are therefore led to investigate the statistics of the largest occupation. This topic has been discussed in janson ; jeon ; gross ; gl2005 ; maj3 ; armendariz2009 ; armendariz2011 ; cg2019 . We briefly revisit this topic and supplement it with equation (2.20) at the end of this subsection which sheds some new light on the subject and lay the ground for the parallel study of renewal processes.

Let XmaxX_{\rm max} be the largest summand (or occupation) under the conditioning Sn=LS_{n}=L,

Xmax=max(X1,,Xn).X_{\rm max}={\rm max}(X_{1},\dots,X_{n}). (2.10)

The distribution function of this variable is

Fn(k|L)=Prob(Xmaxk|Sn=L)=Prob(Xmaxk,Sn=L)Zn(L),F_{n}(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}\leq k|S_{n}=L)=\frac{\mathop{\rm Prob}\nolimits(X_{{\rm max}}\leq k,S_{n}=L)}{Z_{n}(L)}, (2.11)

whose numerator is

Fn(k|L)|num=Prob(Xmaxk,Sn=L)=k1=0kf(k1)kn=0kf(kn)δ(i=1nki,L).F_{n}(k|L)_{|{\rm num}}=\mathop{\rm Prob}\nolimits(X_{{\rm max}}\leq k,S_{n}=L)=\sum_{k_{1}=0}^{k}f(k_{1})\dots\sum_{k_{n}=0}^{k}f(k_{n})\delta\Big{(}\sum_{i=1}^{n}k_{i},L\Big{)}. (2.12)

The generating function of the latter reads

L0zLFn(k|L)|num=i=1n(ki=0kf(ki)zki)=f~(z,k)n,\sum_{L\geq 0}z^{L}F_{n}(k|L)_{|{\rm num}}=\prod_{i=1}^{n}\Big{(}\sum_{k_{i}=0}^{k}f(k_{i})z^{k_{i}}\Big{)}=\tilde{f}(z,k)^{n}, (2.13)

where

f~(z,k)=j=0kzjf(j).\tilde{f}(z,k)=\sum_{j=0}^{k}z^{j}f(j). (2.14)

The distribution of the largest occupation is thus given by the difference

pn(k|L)=Prob(Xmax=k|Sn=L)=Fn(k|L)Fn(k1|L),p_{n}(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}=k|S_{n}=L)=F_{n}(k|L)-F_{n}(k-1|L), (2.15)

where

Fn(0|L)|num=f(0)nδ(L,0).F_{n}(0|L)_{|{\rm num}}=f(0)^{n}\delta(L,0). (2.16)

Its generating function is

L0zLpn(k|L)|num=f~(z,k)nf~(z,k1)n.\sum_{L\geq 0}z^{L}p_{n}(k|L)_{|{\rm num}}=\tilde{f}(z,k)^{n}-\tilde{f}(z,k-1)^{n}. (2.17)

The numerator (2.12) obeys the recursion

Fn(k|L)|num=j=0min(k,L)f(j)Fn1(k|Lj)|num,F_{n}(k|L)_{|{\rm num}}=\sum_{j=0}^{\min(k,L)}f(j)\,F_{n-1}(k|L-j)_{|{\rm num}}, (2.18)

with initial condition

F0(k|L)|num=δ(L,0).F_{0}(k|L)_{|{\rm num}}=\delta(L,0). (2.19)

Let us note that if the occupation number X1X_{1} is larger than L/2L/2, then it is necessarily the largest one, XmaxX_{\rm max}. If so, the probability distribution of the latter, pn(k|L)p_{n}(k|L), is identical to nπn(k|L)n\pi_{n}(k|L), since there are nn possible choices of the generic summand X1X_{1}. Denoting the restriction of pn(k|L)p_{n}(k|L) to the range k>L/2k>L/2 by qn(k|L)q_{n}(k|L), we thus have

qn(k|L)=nπn(k|L)=nf(k)Zn1(Lk)Zn(L).q_{n}(k|L)=n\pi_{n}(k|L)=\frac{nf(k)Z_{n-1}(L-k)}{Z_{n}(L)}. (2.20)

We shall see later that this relation, as simple as it may seem, is instrumental for the analysis of the fluctuations of the condensate and extends naturally to the case of tdrp or free renewal processes. When k>L/2k>L/2 we note that

Fn(k|Lk)|num=Zn(Lk),F_{n}(k|L-k)_{|{\rm num}}=Z_{n}(L-k), (2.21)

since XmaxX_{{\rm max}} is necessarily less than kk (with k>L/2k>L/2) when Sn=LkS_{n}=L-k.

2.2 Phenomenology of condensation in the thermodynamic limit

This subsection is a reminder of well-known facts on the phenomenon of condensation for a thermodynamic system with a large number nn of sites and large total occupation LL, at fixed density ρ=L/n\rho=L/n. More detailed accounts or complements on this topic can be found, e.g., in burda ; burda2 ; janson ; gross ; hanney ; lux ; maj2 ; maj3 ; armendariz2009 ; armendariz2011 ; cg2019 . with further bibliographical references contained in the last reference.

This reminder will help emphasising the differences between the scenario of condensation in the thermodynamic limit, described in the present subsection, with another scenario of condensation, to be described in section 3, where LL is still large, but nn (the number of summands or sites) is kept fixed. In the latter regime, condensation will turn out to be total, with a condensed fraction Xmax/LX_{\rm max}/L asymptotically equal to unity.

For the time being, we consider the situation where nn and LL are both large, with ρ=L/n\rho=L/n kept fixed, assuming that f(k)f(k) has a power-law tail (1.1) and that X=ρc\langle X\rangle=\rho_{c} is finite (θ>1{\theta}>1).

2.2.1 Regimes for the single occupation distribution

Evidence for the existence of a condensate, i.e., a site with a macroscopic occupation, is demonstrated by the behaviour of the single occupation distribution (2.8). There are three regimes to consider, according to the respective values of ρ\rho and ρc\rho_{c}.

1. Subcritical regime (ρ<ρc\rho<\rho_{c})
The asymptotic estimate of the partition function Zn(L)Z_{n}(L) is given by the saddle-point method

Zn(L)=dz2πizL+1f~(z)nf~(z0)nz0L,Z_{n}(L)=\oint\frac{{\rm d}z}{2\pi{\rm i}z^{L+1}}\,\tilde{f}(z)^{n}\sim\frac{\tilde{f}(z_{0})^{n}}{z_{0}^{L}}, (2.22)

where z0z_{0} obeys the saddle-point (sp) equation

z0f~(z0)f~(z0)=ρ.\frac{z_{0}\tilde{f}^{\prime}(z_{0})}{\tilde{f}(z_{0})}=\rho. (2.23)

This equation has a solution z0(ρ)z_{0}(\rho) for any ρ<ρc\rho<\rho_{c}. It follows that

πn(k|L)spz0kf(k)f~(z0),\pi_{n}(k|L)_{\rm sp}\approx\frac{z_{0}^{k}f(k)}{\tilde{f}(z_{0})}, (2.24)

which is no longer dependent on LL and nn separately and only depends on their ratio ρ\rho. Note that (2.23) and (2.24) entail that

X|Lsp=k0kπn(k|L)spz0f~(z0)f~(z0)=ρ,\langle X|L\rangle_{\rm sp}=\sum_{k\geq 0}k\pi_{n}(k|L)_{\rm sp}\approx\frac{z_{0}\tilde{f}^{\prime}(z_{0})}{\tilde{f}(z_{0})}=\rho, (2.25)

consistently with (2.9). In this regime, the system is made of a fluid of independent particles with common distribution (2.24).

2. Critical regime (ρ=ρc\rho=\rho_{c})
A phase transition occurs when the saddle-point value z0z_{0} reaches the maximum value of zz, equal to one, where f~(z)\tilde{f}(z) is singular, with a branch cut chosen to be on the negative axis. The bulk of the partition function is given by the generalised central limit theorem and

πn(k|L)f(k),\pi_{n}(k|L)\approx f(k), (2.26)

up to finite-size corrections. At criticality the equality X|L=X\langle X|L\rangle=\langle X\rangle holds identically thanks to (2.9). In this regime, the system is made of a critical fluid of independent particles with common distribution (2.26).

3. Supercritical regime (ρ>ρc\rho>\rho_{c})
In this regime the saddle-point equation (2.23) can no longer be satisfied because z0z_{0} sticks to the head of the cut of f~(z)\tilde{f}(z). The excess difference,

Δ=LnX=n(ρρc),\Delta=L-n\langle X\rangle=n(\rho-\rho_{c}), (2.27)

instead of being equally shared by all the sites, is, with high probability, accommodated by a single site, the condensate. The partition function Zn(L)Z_{n}(L) is asymptotically given by its right tail (see cg2019 for more details),

Zn(L)nL=nρncΔ1+θ.Z_{n}(L)\mathrel{\mathop{\approx}\limits_{n\to\infty}^{L=n\rho}}\frac{nc}{\Delta^{1+{\theta}}}. (2.28)

In the supercritical regime, the marginal distribution πn(k|L)\pi_{n}(k|L) has different behaviours in the three regions of values of the occupation variable.

(a) The critical background corresponds to values of kk finite, for which (2.26) holds again. The main contribution to the total weight comes from this region.

(b) The condensate is located in the region kΔk\approx\Delta (i.e., the difference Δk\Delta-k is subextensive). The ratio of f(k)c/Δ1+θf(k)\approx c/\Delta^{1+\theta} to Zn(L)Z_{n}(L), given by (2.28), is asymptotically equal to

f(k)Zn(L)c/Δ1/θnc/Δ1/θ=1n.\frac{f(k)}{Z_{n}(L)}\approx\frac{c/\Delta^{1/\theta}}{nc/\Delta^{1/\theta}}=\frac{1}{n}. (2.29)

On the other hand, Zn1(Lk)Z_{n-1}(L-k), is given by its bulk since LknρcL-k\approx n\rho_{c}. Hence, if 1<θ<21<\theta<2,

πn(k|L)|cond1nZn1(Lk)1n1n1/θθ,c(Δkn1/θ),\pi_{n}(k|L)_{|{\rm cond}}\approx\frac{1}{n}Z_{n-1}(L-k)\approx\frac{1}{n}\frac{1}{n^{1/\theta}}\mathcal{L}_{\theta,c}\left(\frac{\Delta-k}{n^{1/\theta}}\right), (2.30)

where θ,c\mathcal{L}_{\theta,c} is the stable Lévy distribution of index θ{\theta}, asymmetry parameter β=1\beta=1, and tail parameter cc gnedenkoK , while, if θ>2\theta>2,

πn(k|L)|cond1nZn1(Lk)1n1n1/2𝒢(Δkn1/2),\pi_{n}(k|L)_{|{\rm cond}}\approx\frac{1}{n}Z_{n-1}(L-k)\approx\frac{1}{n}\frac{1}{n^{1/2}}\mathcal{G}\left(\frac{\Delta-k}{n^{1/2}}\right), (2.31)

where 𝒢\mathcal{G} is the Gaussian distribution gnedenkoK . These expressions describe the bulk of the fluctuating condensate which manifests itself by a hump in the marginal distribution πn(k|L)\pi_{n}(k|L), in the neighbourhood of kΔk\approx\Delta, visible on figure 4. The weight of this region is obtained from (2.30) or (2.31), according to the value of θ{\theta}, as

Prob(X1cond)=khumpπn(k|L)|cond1n,\mathop{\rm Prob}\nolimits(X_{1}\in\mathrm{cond})=\sum_{k\in\mathrm{hump}}\pi_{n}(k|L)_{|{\rm cond}}\approx\frac{1}{n}, (2.32)

which demonstrates that the excess difference Δ\Delta is typically borne by only one summand.

This hump becomes peaked in the thermodynamic limit. For a finite system, most often there is a single condensate, i.e., a site with a macroscopic occupation, while more rarely there are two sites with macroscopic occupations, both of order LL. This situation corresponds to the dip region, described next.

(c) The range of values of kk such that kk and Δk\Delta-k are large and comparable, interpolates between the critical part of πn(k|L)\pi_{n}(k|L), for kk or order 11, and the condensate, for kk close to Δ\Delta. It corresponds to the dip region on figure 4. In this region, Zn1(Lk)Z_{n-1}(L-k) is given by its right tail (2.28). So, for any θ>1\theta>1,

πn(k|L)|dipc[Δk(Δk)]1+θf(k)f(Δk)f(Δ).\pi_{n}(k|L)_{|{\rm dip}}\approx c\left[\frac{\Delta}{k(\Delta-k)}\right]^{1+\theta}\approx\frac{f(k)f(\Delta-k)}{f(\Delta)}. (2.33)

The interpretation of this result is that in the dip region typical configurations where one summand takes the value kk are such that the remaining Δk\Delta-k excess difference is borne by a single other summand. The dip region is therefore dominated by rare configurations where the excess difference is shared by two summands gl2005 . An example of such a configuration is the green path in figure 1.

Setting k=λΔk={\lambda}\Delta in (2.33) and introducing a cutoff Λ=ϵΔ\Lambda={\epsilon}\Delta, the weight of these configurations can be estimated as

Prob(X1dip)=k=ϵΔ(1ϵ)Δπn(k|L)|dipΔθnθ.\mathop{\rm Prob}\nolimits(X_{1}\in\mathrm{dip})=\sum_{k={\epsilon}\Delta}^{(1-{\epsilon})\Delta}\pi_{n}(k|L)_{|{\rm dip}}\sim\Delta^{-\theta}\sim n^{-\theta}. (2.34)

The relative weights of the dip (2.34) and condensate (2.32) regions is therefore of order n(θ1)n^{-(\theta-1)}, i.e., the weight of events where the condensate is broken into two pieces of order nn is subleading with respect to events with a single big summand.

2.2.2 Statistics of the largest summand in the condensed phase

In view of (2.32) and the following equations (2.33) and (2.34), we infer that, for kk larger than Δ/2\Delta/2, which is the centre of the dip, the excess difference Δ\Delta is typically borne by only one summand—namely the condensate XmaxX_{\rm max}—thus

pn(k|L)k>Δ/2nπn(k|L).p_{n}(k|L)\mathrel{\mathop{\approx}\limits_{k>\Delta/2}}n\pi_{n}(k|L). (2.35)

According to (2.20) we know that the two sides of this equation are actually identical for k>L/2k>L/2, for any finite values of nn and LL. Note however that, while L/2L/2 is always larger than Δ/2\Delta/2, it can be smaller or larger than Δ\Delta depending on whether ρ\rho is larger or smaller than 2ρc2\rho_{c}. The significance of (2.35) is that the equality (2.20), valid for k>L/2k>L/2, extends asymptotically to the entire region k>Δ/2k>\Delta/2.

More precise statements have been given on the asymptotic distribution of the largest summand XmaxX_{\rm max} in maj3 ; armendariz2009 ; armendariz2011 ; janson . The result is that, if L=nρL=n\rho, ρ>ρc\rho>\rho_{c}, nn\to\infty, the rescaled variable n1/α(ΔXmax)n^{-1/\alpha}(\Delta-X_{\rm max}) converges to a stable law of index α\alpha, with α=θ\alpha=\theta if θ<2\theta<2, or α=2\alpha=2 if θ>2\theta>2. This means that, asymptotically, the probability distribution of XmaxX_{\rm max} coincides, up to a factor nn, with the estimates of the marginal density in the condensate region (Δkn1/α\Delta-k\sim n^{1/\alpha}), that is with (2.30) or (2.31) according to the value of θ\theta, which is precisely the content of (2.35).

On the other hand, denoting the rr-th largest summand by X(r)X^{(r)} (r=1,,nr=1,\dots,n), with X(1)XmaxX^{(1)}\equiv X_{\rm max}, the distributions of these ranked summands, denoted by pn(r)(k|L)p_{n}^{(r)}(k|L) with pn(1)(k|L)pn(k|L)p_{n}^{(1)}(k|L)\equiv p_{n}(k|L), sum up exactly to

pn(k|L)+r=2npn(r)(k|L)=nπn(k|L).p_{n}(k|L)+\sum_{r=2}^{n}p_{n}^{(r)}(k|L)=n\pi_{n}(k|L). (2.36)

Thus according to (2.35) the sum upon r2r\geq 2 in the left side of (2.36) is negligible for k>Δ/2k>\Delta/2.

As shown in armendariz2009 ; armendariz2011 ; janson , the distribution of the second largest summand, X(2)X^{(2)}, is asymptotically Fréchet, and the subsequent ones, X(r)X^{(r)} (r2)(r\geq 2), are the order statistics of n1n-1 iid random variables XiX_{i} with distribution f(k)f(k), which amounts to saying that, in the supercritical regime, the dependency between the summands XiX_{i} introduced by the conditioning goes asymptotically in the condensate XmaxX_{\rm max}.

Since XmaxX_{\rm max} typically scales as nn, while X(2),X(3),X^{(2)},X^{(3)},\dots typically scale as n1/θn^{1/\theta}, the condensate is increasingly separated from the background as nn increases, leaving space to the dip region (kk and Δk\Delta-k large and comparable). We know from the analysis made above (see discussion following (2.33)) that this region is dominated by configurations where the excess difference is shared by two summands, namely XmaxX_{\rm max} and X(2)X^{(2)}, so

pn(k|L)+pn(2)(k|L)kdipnπn(k|L)|dip,p_{n}(k|L)+p_{n}^{(2)}(k|L)\mathrel{\mathop{\approx}\limits_{k\in\mathrm{dip}}}n\pi_{n}(k|L)|_{\rm dip}, (2.37)

and that the contributions of these events to nπn(k|L)n\pi_{n}(k|L) are of relative order n(θ1)n^{-(\theta-1)}. To the right of Δ/2\Delta/2 the predominant contribution to the sum on the left side of (2.37) comes from pn(k|L)p_{n}(k|L), to the left of Δ/2\Delta/2 it comes from pn(2)(k|L)p_{n}^{(2)}(k|L).

An illustration

Refer to caption
Figure 4: Random allocations models and zrp: comparison of nπn(k|L)n\pi_{n}(k|L) (where πn(k|L)\pi_{n}(k|L) is the single occupation distribution) with pn(k|L)p_{n}(k|L) (distribution of the maximum) for the example (2.38), with θ=3{\theta}=3, ρc0.1106\rho_{c}\approx 0.1106, n=600n=600, L=168L=168, Δ102\Delta\approx 102. The vertical dotted line is at L/2L/2. There is an exact identity between nπn(k|L)n\pi_{n}(k|L) and pn(k|L)p_{n}(k|L), for k>L/2k>L/2, as explained in §2.1.2. Moreover, there is already excellent numerical coincidence between the two curves as soon as k>58Δ/2k>58\gtrsim\Delta/2.

Figure 4 depicts a comparison between nπn(k|L)n\pi_{n}(k|L), obtained from (2.8), and pn(k|L)p_{n}(k|L) obtained from (2.17), on the following example, defined by the normalised distribution

f(k)=1ζ(1+θ)1(k+1)1+θ,(k0),f(k)=\frac{1}{\zeta(1+{\theta})}\frac{1}{(k+1)^{1+{\theta}}},\quad(k\geq 0), (2.38)

where ζ(s)=s11/ns\zeta(s)=\sum_{s\geq 1}1/n^{s} is the Riemann zeta function. This model has been introduced in burda , then further investigated in camia ; burda2 ; glzeta .

In this figure, θ=3{\theta}=3, ρc0.1106\rho_{c}\approx 0.1106, n=600n=600, L=168L=168. This choice of parameters corresponds to a density ρ=0.28\rho=0.28 slightly larger than 2ρc2\rho_{c}, where Δ\Delta and L/2L/2 coincide. The top of the hump is approximately located at Δ102\Delta\approx 102 and the minimum of the dip is a bit less than Δ/2\Delta/2. These curves are practically indiscernible as soon as k58k\approx 58, which is less than L/2=84L/2=84, indicated by the vertical dotted line on the figure, from which the identity becomes exact.

Refer to caption
Figure 5: Random allocations models and zrp: comparison of nπn(k|L)n\pi_{n}(k|L) (where πn(k|L)\pi_{n}(k|L) is the single occupation distribution) with pn(k|L)p_{n}(k|L) (distribution of the maximum) for the example (3.1) corresponding to a tail index θ=1/2{\theta}=1/2, with L=60L=60, n=4n=4. The inset highlights the cusp at L/2L/2.

Finally, let us compute the mean condensed fraction Xmax/LX_{\rm max}/L. For θ>2{\theta}>2, using (2.31), it can be estimated as

XmaxL\displaystyle\frac{\langle X_{\rm max}\rangle}{L} =\displaystyle= 1Lk=0Lkpn(k|L)1Lk=Δ/2Lknπn(k|L)\displaystyle\frac{1}{L}\sum_{k=0}^{L}k\,p_{n}(k|L)\approx\frac{1}{L}\sum_{k=\Delta/2}^{L}k\,n\,\pi_{n}(k|L) (2.39)
\displaystyle\approx 1L1n1/22πdkke12(Δkn1/2)2=1ρcρ=ΔL.\displaystyle\frac{1}{L}\frac{1}{n^{1/2}\sqrt{2\pi}}\int_{-\infty}^{\infty}{\rm d}k\,k\,{\rm e}^{-\frac{1}{2}\left(\frac{\Delta-k}{n^{1/2}}\right)^{2}}=1-\frac{\rho_{c}}{\rho}=\frac{\Delta}{L}.

The same result holds if 1<θ<21<{\theta}<2, using (2.30).

As ρ\rho increases, the peak of the condensate moves towards the right end LL, hence if ρρc\rho\gg\rho_{c} the condensed fraction tends to unity, corresponding to total condensation. As detailed in section 3, this scenario still holds when nn is kept fixed.

3 Phenomenon of total condensation when nn is kept fixed and LL\to\infty

As seen above, if ρρc\rho\gg\rho_{c}, condensation becomes total, and the peak of the condensate is asymptotically located at LL. As we now show, this still holds true if the number of summands nn is kept fixed, and LL is large, irrespective of the existence of a first moment ρc=X\rho_{c}=\langle X\rangle, or in other words, irrespective of whether θ{\theta} is smaller or larger than one. Existence of condensation in such a situation has been pointed out in ferrari ; landim . The quantitative characterisation of this phenomenon is the aim of this section.

3.1 An illustration

We start by giving an illustration of the phenomenon on the following example, corresponding to a tail index θ=1/2{\theta}=1/2 for the decay of f(k)f(k),

f~(z)=11zz,\tilde{f}(z)=\frac{1-\sqrt{1-z}}{z}, (3.1)

which entails that

f(k)=122k+1(2k)!k!(k+1)!kck3/2,f(k)=\frac{1}{2^{2k+1}}\frac{(2k)!}{k!(k+1)!}\mathrel{\mathop{\approx}\limits_{k\to\infty}}\frac{c}{k^{3/2}}, (3.2)

with c=1/(2π)c=1/(2\sqrt{\pi}), and

Zn(L)=n22L+n(2L+n)(2L+nL)LnfixedncL3/2,Z_{n}(L)=\frac{n}{2^{2L+n}(2L+n)}\left(2L+n\atop L\right)\mathrel{\mathop{\approx}\limits_{L\to\infty}^{n\ \mathrm{fixed}}}\frac{nc}{L^{3/2}}, (3.3)

for nn kept fixed, LL large, which is a different regime from that leading to (2.28). In the general case, whenever f(k)f(k) obeys (1.1), the asymptotic estimate (3.3) becomes

Zn(L)LnfixedncL1+θ,Z_{n}(L)\mathrel{\mathop{\approx}\limits_{L\to\infty}^{n\ \mathrm{fixed}}}\frac{nc}{L^{1+{\theta}}}, (3.4)

as can be deduced from (2.6).

Figure 5 depicts a comparison of nπn(k|L)n\pi_{n}(k|L) with pn(k|L)p_{n}(k|L) for this example with L=60L=60, n=4n=4. These two quantities are identical for k>L/2k>L/2, The inset highlights the existence of a cusp at L/2L/2. The distribution of the maximum pn(k|L)p_{n}(k|L) is significantly depressed for k<L/2k<L/2. It vanishes identically for k<L/4k<L/4 (more generally L/nL/n). These features are analysed in the two subsections below.

3.2 Fluctuations of the condensate

Let us turn to the general case. A measure of the fluctuations of the condensate is provided by the width of the peak of the maximum, i.e., the mass outside the condensate. It can be estimated by the sum (see §3.3 below for details)

LXmax==0L1pn(L|L)=0L/21qn(L|L)=0L/21nπn(L|L),L-\langle X_{\rm max}\rangle=\sum_{\ell=0}^{L-1}\ell\,p_{n}(L-\ell|L)\approx\sum_{\ell=0}^{L/2-1}\ell\,q_{n}(L-\ell|L)\approx\sum_{\ell=0}^{L/2-1}\ell\,n\pi_{n}(L-\ell|L), (3.5)

where =Lk\ell=L-k. The dominant contribution to this sum depends on whether θ{\theta} is smaller or larger than one.

\bullet If θ<1{\theta}<1, the dominant contribution comes from values of =Lk\ell=L-k comparable to LL. Setting =λL\ell=\lambda L in (3.5), we have (see §3.3 below for details)

LXmax(n1)cL1θ01/2dλλ[(λ(1λ)]1+θ(n1)cB12(1θ,θ)L1θ,L-\langle X_{\rm max}\rangle\approx(n-1)cL^{1-{\theta}}\int_{0}^{1/2}{\rm d}{\lambda}\frac{{\lambda}}{[({\lambda}(1-{\lambda})]^{1+{\theta}}}\approx(n-1)c\,\mathrm{B}_{\frac{1}{2}}\Big{(}1-{\theta},-{\theta}\Big{)}L^{1-{\theta}}, (3.6)

where the incomplete beta function is defined as

Bx(a,b)=0xdtta1(1t)b1.\mathrm{B}_{x}(a,b)=\int_{0}^{x}{\rm d}t\,t^{a-1}(1-t)^{b-1}. (3.7)

For example, if θ=1/2{\theta}=1/2, we have B12(1θ,θ)=2\mathrm{B}_{\frac{1}{2}}\Big{(}1-{\theta},-{\theta}\Big{)}=2.

\bullet If θ>1{\theta}>1, the main contribution comes from finite values of \ell,

LXmax\displaystyle L-\langle X_{\rm max}\rangle \displaystyle\approx n=0L/2f(L)Zn1()Zn(L)=0L/2Zn1()\displaystyle n\sum_{\ell=0}^{L/2}\ell\,\frac{f(L-\ell)Z_{n-1}(\ell)}{Z_{n}(L)}\approx\sum_{\ell=0}^{L/2}\ell\,Z_{n-1}(\ell) (3.8)
\displaystyle\approx (n1)=0L/2f()(n1)=0f()=(n1)X.\displaystyle(n-1)\sum_{\ell=0}^{L/2}\ell\,f(\ell)\to(n-1)\sum_{\ell=0}^{\infty}\ell f(\ell)=(n-1)\langle X\rangle.

This last result (3.8) has a simple interpretation. It says that the correction LXmaxL-\langle X_{\rm max}\rangle comes from the n1n-1 sites of the fluid, each with mean occupation X=ρc\langle X\rangle=\rho_{c}, in accordance with the prediction made in ferrari and recalled in the introduction.

3.3 A finer analysis

Let us now add some more details on the derivations made above. The aim of this subsection is to give a detailed analysis of the distributions in the various regimes, in order to eventually compute the corrections to the scaling expressions predicted in (3.6) and (3.8) above.

We start with the discussion of the regimes for the single occupation distribution Prob(X1=k|Sn=L)=πn(k|L)\mathop{\rm Prob}\nolimits(X_{1}=k|S_{n}=L)=\pi_{n}(k|L). There are such three regimes to consider (see figure 5):

  1. 1.

    Downhill region. For X1=kX_{1}=k finite, using (3.4), we have

    πn(k|L)Ln1nf(k),\pi_{n}(k|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{n-1}{n}f(k), (3.9)

    reflecting the fact that, with probability (n1)/n(n-1)/n, a randomly chosen site belongs to the fluid.

    Introducing a cutoff Λ\Lambda, such that 1ΛL1\ll\Lambda\ll L, the weight of this region can be estimated by the sum

    k=0Λπn(k|L)k=0Λn1nf(k)Lk=0n1nf(k)=11n.\sum_{k=0}^{\Lambda}\pi_{n}(k|L)\approx\sum_{k=0}^{\Lambda}\frac{n-1}{n}f(k)\mathrel{\mathop{\to}\limits_{L\to\infty}}\sum_{k=0}^{\infty}\frac{n-1}{n}f(k)=1-\frac{1}{n}. (3.10)
  2. 2.

    Dip region. In the dip region, where kk and LkL-k are simultaneously large and comparable, setting k=λLk={\lambda}L in (2.8) where 0<λ<10<{\lambda}<1, and using (3.4), yields the estimate

    πn(k|L)Ln1nf(k)f(Lk)f(L)n1ncL1+θ1[λ(1λ)]1+θ,\pi_{n}(k|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{n-1}{n}\frac{f(k)f(L-k)}{f(L)}\approx\frac{n-1}{n}\frac{c}{L^{1+{\theta}}}\frac{1}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}, (3.11)

    In this region the distribution is therefore U-shaped: the most probable configurations are those where almost all the particles are located on one of two sites. The dip centred around k=L/2k=L/2 becomes deeper and deeper with LL.

    The weight of this region reads, choosing Λ=ϵL\Lambda={\epsilon}L,

    k=ΛLΛπn(k|L)n1ncLθϵ1ϵdλ[λ(1λ)]1+θ.\sum_{k=\Lambda}^{L-\Lambda}\pi_{n}(k|L)\approx\frac{n-1}{n}cL^{-{\theta}}\int_{\epsilon}^{1-{\epsilon}}\frac{{\rm d}{\lambda}}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}. (3.12)
  3. 3.

    Uphill region. The condensate region corresponds to =Lk\ell=L-k finite, where (2.8) simplifies into

    πn(L|L)L1nZn1(),\pi_{n}(L-\ell|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{1}{n}Z_{n-1}(\ell), (3.13)

    as in (2.30) or (2.31). The weight of this uphill region can be estimated as

    =0Λπn(L|L)1n=0ΛZn1()1n=0Zn1()=1n,\sum_{\ell=0}^{\Lambda}\pi_{n}(L-\ell|L)\approx\frac{1}{n}\sum_{\ell=0}^{\Lambda}Z_{n-1}(\ell)\to\frac{1}{n}\sum_{\ell=0}^{\infty}Z_{n-1}(\ell)=\frac{1}{n}, (3.14)

    where Λ\Lambda is yet another cutoff, and where the last step is obtained by setting z=1z=1 in the expression of the generating function (2.6). Thus, as seen in (3.10) and (3.14) the weights of the downhill and uphill regions add up to one, in line with the fact that the contribution of the dip region is subdominant, as shown in (3.12) above.

We now proceed to the discussion of the regimes for the distribution of the maximum, pn(k|L)p_{n}(k|L). There are again three regimes to consider, that we describe in turn, from right to left in figure 5. In the uphill and dip regions, such that Xmax=k>L/2X_{\rm max}=k>L/2, pn(k|L)p_{n}(k|L) is denoted by qn(k|L)=nπn(k|L)q_{n}(k|L)=n\pi_{n}(k|L) (see (2.20)), whose estimates follow from those of πn(k|L)\pi_{n}(k|L) seen above.

  1. 1.

    Uphill region. For \ell finite, using (3.13), we have

    qn(L|L)Zn1(),q_{n}(L-\ell|L)\approx Z_{n-1}(\ell), (3.15)

    with weight equal to 1 up to the subleading corrections detailed below. The interpretation of (3.15) is that, asymptotically, the difference between LL and XmaxX_{\rm max} has the same distribution as the sum of n1n-1 iid random variables, the latter composing the fluid,

    LXmaxLi=1n1Xifluid.L-X_{\rm max}\mathrel{\mathop{\approx}\limits_{L\to\infty}}\ \mathrel{\mathop{\underbrace{\sum_{i=1}^{n-1}X_{i}}}\limits_{\rm fluid}}. (3.16)
  2. 2.

    Dip region. For L/2<kLkL/2<k\sim L-k, we have, according to (3.11),

    qn(k|L)(n1)f(k)f(Lk)f(L)(n1)cL1+θ1[λ(1λ)]1+θ.q_{n}(k|L)\approx\frac{(n-1)f(k)f(L-k)}{f(L)}\approx\frac{(n-1)\,c}{L^{1+{\theta}}}\frac{1}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}. (3.17)

    The weight of this region therefore scales as LθL^{-{\theta}}, as seen in (3.12).

  3. 3.

    Left region. For kL/2k\leq L/2, the weight of this region is subdominant with respect to that of the two previous ones. A simple argument shows that

    Prob(XmaxL/2|Sn=L)=Fn(L/2|L)Lβ,β={2θif θ11+θif θ>1,\mathop{\rm Prob}\nolimits(X_{\rm max}\leq L/2|S_{n}=L)=F_{n}(L/2|L)\sim L^{-\beta},\qquad\beta=\left\{\begin{array}[]{ll}2{\theta}&\textrm{if }{\theta}\leq 1\vspace{4pt}\\ 1+{\theta}&\textrm{if }{\theta}>1,\end{array}\right. (3.18)

    where FnF_{n} is defined in (2.11).

The argument leading to (3.18) and the prediction of the amplitude limLFn(L/2|L)Lβ\lim_{L\to\infty}F_{n}(L/2|L)L^{\beta} are given in appendix A. In appendix B, we give an exact calculation of the weight of the left region when f(k)f(k) is the continuous Lévy 12\frac{1}{2} stable density.

Equation (3.18) eventually justifies the approximation made in (3.5), where the contribution of the left region to the sum =0L1pn(L|L)\sum_{\ell=0}^{L-1}\ell\,p_{n}(L-\ell|L) was neglected. In view of (3.18), this contribution is O(L1β)O(L^{1-\beta}), thus subdominant by a factor LθL^{-{\theta}} with respect to the first correction—respectively O(L1θ)O(L^{1-{\theta}}) if θ<1{\theta}<1 (see (3.6)), or O(1)O(1) for θ>1{\theta}>1 (see (3.8))—whether θ{\theta} is smaller of larger than unity.

There is actually a hierarchy of weights for the distribution of the maximum in the successive regions (L/3,L/2)(L/3,L/2), (L/4,L/3)(L/4,L/3) and so on. This can be intuitively grasped as follows.

  1. 1.

    If L/3<XmaxL/2L/3<X_{\rm max}\leq L/2, the total ‘mass’ LL is dominantly shared by two summands. The weight of this rare event scales as in (3.18).

  2. 2.

    Then if L/4<XmaxL/3L/4<X_{\rm max}\leq L/3, the total ‘mass’ LL is dominantly shared by three summands, which is a still rarer event, and so on.

  3. 3.

    Finally, if Xmax<L/nX_{\rm max}<L/n, the probability pn(k|L)p_{n}(k|L) vanishes since it is no longer possible to divide the ‘mass’ LL into nn pieces, all less than L/nL/n.

This hierarchy is reflected by the presence of cusps in the distribution of the maximum at L/2,L/3L/2,L/3, L/4L/4\dots (see appendix A).

All the discussion given in the present section is a preparation for subsections 8.3 and 11.4, where figures 9 and 11 are to be compared to figure 5, and equations (8.31), (8.32), (11.27) and (11.28) are to be compared to equations (3.6) and (3.8).

4 General statements on tied-down renewal processes

The random variables XiX_{i} now represent the sizes of (spatial or temporal) intervals, that we take strictly positive, hence

f(0)=0.f(0)=0. (4.1)

In the temporal language LL is the total duration of the process, in the spatial language it is the length of the system. To each interval (equivalently, to each renewal event) is associated a positive weight ww, to be interpreted as a reward if w>1w>1 or a penalty if w<1w<1. In the models considered in burda3 ; bar2 ; bar3 ; barma , ww has the interpretation of the ratio y/ycy/y_{c}, where yy is a fugacity, and ycy_{c} its value at criticality.

In the present section, f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k) is any arbitrary distribution of the positive random variable XX. Later on, in sections 5, 6, 7 and 8, f(k)f(k) will obey the form (1.1).

4.1 Joint distribution

The probability of the configuration {X1=k1,,XNL=kn,NL=n}\{X_{1}=k_{1},\dots,X_{N_{L}}=k_{n},N_{L}=n\}, given that SNL=LS_{N_{L}}=L, reads

p(k1,,kn,n|L)\displaystyle p(k_{1},\dots,k_{n},n|L) =\displaystyle= Prob(X1=k1,,XNL=kn,NL=n|SNL=L)\displaystyle\mathop{\rm Prob}\nolimits(X_{1}=k_{1},\dots,X_{N_{L}}=k_{n},N_{L}=n|S_{N_{L}}=L) (4.2)
=\displaystyle= 1Ztd(w,L)wnf(k1)f(kn)δ(i=1nki,L),\displaystyle\frac{1}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}w^{n}f(k_{1})\dots f(k_{n})\delta\Big{(}\sum_{i=1}^{n}k_{i},L\Big{)},

where the denominator is the tied-down partition function

Ztd(w,L)\displaystyle Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) =\displaystyle= n0wn{ki}f(k1)f(kn)δ(i=1nki,L)\displaystyle\sum_{n\geq 0}w^{n}\sum_{\{k_{i}\}}f(k_{1})\dots f(k_{n})\delta\Big{(}\sum_{i=1}^{n}k_{i},L\Big{)} (4.3)
=\displaystyle= n0wnZn(L)=δ(L,0)+n1wn(f)n(L).\displaystyle\sum_{n\geq 0}w^{n}Z_{n}(L)=\delta(L,0)+\sum_{n\geq 1}w^{n}(f\star)^{n}(L).

The probability Zn(L)Z_{n}(L) is still defined as in (2.2), except for the change of the initial value (4.1) of f(k)f(k), which entails that Zn(L)Z_{n}(L) is only defined for nLn\leq L. The first term δ(L,0)\delta(L,0) follows from (2.4). The first values of Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) are

Ztd(w,0)=1,Ztd(w,1)=wf(1),Ztd(w,2)=wf(2)+w2f(1)2,\displaystyle Z^{\mathop{\scriptstyle\mathrm{td}}}(w,0)=1,\quad Z^{\mathop{\scriptstyle\mathrm{td}}}(w,1)=wf(1),\quad Z^{\mathop{\scriptstyle\mathrm{td}}}(w,2)=wf(2)+w^{2}f(1)^{2},
Ztd(w,3)=wf(3)+2w2f(1)f(2)+w3f(1)3,\displaystyle Z^{\mathop{\scriptstyle\mathrm{td}}}(w,3)=wf(3)+2w^{2}f(1)f(2)+w^{3}f(1)^{3}, (4.4)

and so on. The generating function of Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) with respect to LL is

Z~td(w,z)=L0zLZtd(w,L)=n0wnf~(z)n=11wf~(z).\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)=\sum_{L\geq 0}z^{L}Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)=\sum_{n\geq 0}w^{n}\tilde{f}(z)^{n}=\frac{1}{1-w\tilde{f}(z)}. (4.5)

Note that Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) can be seen as the grand canonical partition function of the system with respect to NLN_{L}.

For w=1w=1, the tied-down partition function,

Ztd(1,L)=n0Prob(Sn=L)=Prob(SNL=L)=δ(SNL,L),Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)=\sum_{n\geq 0}\mathop{\rm Prob}\nolimits(S_{n}=L)=\mathop{\rm Prob}\nolimits(S_{N_{L}}=L)=\langle\delta(S_{N_{L}},L)\rangle, (4.6)

is the probability that a renewal occurs at LL.

Finally, we note that

p(k1,,kn,n|L)|num=wnpn(k1,,kn|L)|num,p(k_{1},\dots,k_{n},n|L)_{|{\rm num}}=w^{n}p_{n}(k_{1},\dots,k_{n}|L)_{|{\rm num}}, (4.7)

hence

p(k1,,kn,n|L)=pn(k1,,kn|L)pn(L),p(k_{1},\dots,k_{n},n|L)=p_{n}(k_{1},\dots,k_{n}|L)\mathrm{p}_{n}(L), (4.8)

where pn({ki}|L)p_{n}(\{k_{i}\}|L) and pn(L)\mathrm{p}_{n}(L) are respectively defined in (2.1) and (4.9).

4.2 Distribution of the number of intervals

The distribution of the number of intervals is obtained by summing the distribution (4.2) upon all variables except nn,

pn(L)=Prob(NL=n)=wnZn(L)Ztd(w,L).\mathrm{p}_{n}(L)=\mathop{\rm Prob}\nolimits(N_{L}=n)=\frac{w^{n}Z_{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}. (4.9)

For instance, taking the successive terms of (4.3) divided by Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) yields

p0(L)=δ(L,0)Ztd(w,L),p1(L)=wf(L)Ztd(w,L),p2(L)=w2k1f(k1)f(Lk1)Ztd(w,L),\mathrm{p}_{0}(L)=\frac{\delta(L,0)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)},\quad\mathrm{p}_{1}(L)=\frac{wf(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)},\quad\mathrm{p}_{2}(L)=\frac{w^{2}\sum_{k_{1}}f(k_{1})f(L-k_{1})}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)},\dots (4.10)

and more generally, for n1n\geq 1,

pn(L)=wn(f)n(L)Ztd(w,L)=wn[f~(z)n]LZtd(w,L),\mathrm{p}_{n}(L)=\frac{w^{n}(f\star)^{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}=\frac{w^{n}\left[\tilde{f}(z)^{n}\right]_{L}}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}, (4.11)

where the notation []L[\cdot]_{L} stands for the LL-th coefficient of the series inside the brackets. Hence the generating function with respect to LL of the numerator of (4.9) reads

L0zLpn(L)|num=wnf~(z)n.\sum_{L\geq 0}z^{L}\mathrm{p}_{n}(L)_{|{\rm num}}=w^{n}\tilde{f}(z)^{n}. (4.12)

Taking the sum of the right side upon n0n\geq 0 yields back Z~td(w,z)\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z) given in (4.5).

The first moment of this distribution is by definition

NL=n1npn(L).\langle N_{L}\rangle=\sum_{n\geq 1}n\,\mathrm{p}_{n}(L). (4.13)

The generating function of its numerator reads, using (4.12)

L0zLNL|num=n1n(wf~(z))n=wf~(z)(1wf~(z))2=wdZ~td(w,z)dw,\sum_{L\geq 0}z^{L}\langle N_{L}\rangle_{|{\rm num}}=\sum_{n\geq 1}n(w\tilde{f}(z))^{n}=\frac{w\tilde{f}(z)}{\big{(}1-w\tilde{f}(z)\big{)}^{2}}=w\frac{{\rm d}\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)}{{\rm d}w}, (4.14)

hence

NL=wdlnZtd(w,L)dw,\langle N_{L}\rangle=w\,\frac{{\rm d}\ln Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}{{\rm d}w}, (4.15)

as expected in the grand canonical ensemble with respect to NLN_{L}. Alternatively, since

L0zLNL|num=Z~td(w,z)2Z~td(w,z),\sum_{L\geq 0}z^{L}\langle N_{L}\rangle_{|{\rm num}}=\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)^{2}-\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z), (4.16)

we have

NL|num=(ZtdZtd)(w,L)Ztd(w,L).\langle N_{L}\rangle_{|{\rm num}}=(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)-Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L). (4.17)

More generally, the generating function of the moments of NLN_{L} is given by

vNL=n0vnpn(L).\langle v^{N_{L}}\rangle=\sum_{n\geq 0}v^{n}\mathrm{p}_{n}(L). (4.18)

Taking the generating function with respect to LL of the numerator of this expression, using (4.12),

L0zLvNL|num=n0vn(wf~(z))n=11vwf~(z)=Z~td(vw,z),\sum_{L\geq 0}z^{L}\langle v^{N_{L}}\rangle_{|{\rm num}}=\sum_{n\geq 0}v^{n}(w\tilde{f}(z))^{n}=\frac{1}{1-vw\tilde{f}(z)}=\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(vw,z), (4.19)

yields

vNL=Ztd(vw,L)Ztd(w,L).\langle v^{N_{L}}\rangle=\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(vw,L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}. (4.20)

Likewise, the inverse moment 1/NL\langle 1/N_{L}\rangle is

1NL=n1pn(L)n=1Ztd(w,L)n1[[wf~(z)]n]Ln=1Ztd(w,L)[ln(1wf~(z))]L.\Big{\langle}\frac{1}{N_{L}}\Big{\rangle}=\sum_{n\geq 1}\frac{\mathrm{p}_{n}(L)}{n}=\frac{1}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\sum_{n\geq 1}\frac{\left[[w\tilde{f}(z)]^{n}\right]_{L}}{n}=\frac{1}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\left[-\ln(1-w\tilde{f}(z))\right]_{L}. (4.21)

4.3 Single interval distribution

The marginal distribution of one of the summands, say X1X_{1}, is by definition

π(k|L)=Prob(X1=k|SNL=L)=δ(X1,k),\pi(k|L)=\mathop{\rm Prob}\nolimits(X_{1}=k|S_{N_{L}}=L)=\langle\delta(X_{1},k)\rangle, (4.22)

where \langle\cdot\rangle is the average with respect to (4.2), with a summation upon the variables k1,,knk_{1},\dots,k_{n} (with 1kL1\leq k\leq L) and n1n\geq 1, resulting in

π(k|L)|num\displaystyle\pi(k|L)_{|{\rm num}} =\displaystyle= n1k1δ(k1,k)wf(k1)k2,wn1f(k2)f(kn)δ(k1+i=2nki,L)\displaystyle\sum_{n\geq 1}\sum_{k_{1}}\delta(k_{1},k)wf(k_{1})\sum_{{k_{2},\dots}}w^{n-1}f(k_{2})\dots f(k_{n})\delta\Big{(}k_{1}+\sum_{i=2}^{n}k_{i},L\Big{)} (4.23)
=\displaystyle= k1δ(k1,k)wf(k1)δ(k1,L)+k1,k2δ(k1,k)w2f(k1)f(k2)δ(k1+k2,L)+\displaystyle\sum_{k_{1}}\delta(k_{1},k)wf(k_{1})\delta(k_{1},L)+\sum_{k_{1},k_{2}}\delta(k_{1},k)w^{2}f(k_{1})f(k_{2})\delta(k_{1}+k_{2},L)+\cdots
=\displaystyle= wf(k)δ(k,L)+wf(k)Ztd(w,Lk)(1δ(k,L)).\displaystyle wf(k)\delta(k,L)+wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)\big{(}1-\delta(k,L)\big{)}.

Finally

π(k|L)=wf(k)Ztd(w,L)δ(k,L)p(k,1|L)+wf(k)Ztd(w,Lk)Ztd(w,L)(1δ(k,L)),\pi(k|L)=\mathrel{\mathop{\underbrace{\frac{wf(k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\delta(k,L)}}\limits_{p(k,1|L)}}+wf(k)\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\big{(}1-\delta(k,L)\big{)}, (4.24)

where the first term corresponds to n=1n=1, i.e.,

π(L|L)=wf(L)Ztd(w,L)=Prob(NL=1).\pi(L|L)=\frac{wf(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}=\mathop{\rm Prob}\nolimits(N_{L}=1). (4.25)

Also, since Ztd(w,0)=1Z^{\mathop{\scriptstyle\mathrm{td}}}(w,0)=1, (4.24) can be more compactly written as

π(k|L)=wf(k)Ztd(w,Lk)Ztd(w,L).\pi(k|L)=\frac{wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}. (4.26)

The generating function of the numerator of (4.26) with respect to LL yields

LkzLπ(k|L)|num=wzkf(k)Z~td(w,z)=wzkf(k)1wf~(z).\sum_{L\geq k}z^{L}\pi(k|L)_{|{\rm num}}=wz^{k}f(k)\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)=\frac{wz^{k}f(k)}{1-w\tilde{f}(z)}. (4.27)

Summing (4.26) upon kk we obtain

Ztd(w,L)=k=1Lwf(k)Ztd(w,Lk),L1,Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)=\sum_{k=1}^{L}wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k),\quad L\geq 1, (4.28)

which can also be obtained by multiplying the recursion (2.5) for Zn(L)Z_{n}(L) by wnw^{n} and summing on nn.

Remarks

1. An alternative route to (4.26) is as follows. We have

π(k|L)|num\displaystyle\pi(k|L)_{|{\rm num}} =\displaystyle= n1wn{ki}δ(k1,k)pn({ki}|L)|num=n1wnπn(k|L)|num\displaystyle\sum_{n\geq 1}w^{n}\sum_{\{k_{i}\}}\delta(k_{1},k)p_{n}(\{k_{i}\}|L)_{|{\rm num}}=\sum_{n\geq 1}w^{n}\pi_{n}(k|L)_{|{\rm num}} (4.29)
=\displaystyle= n1wnf(k)Zn1(Lk)=wf(k)n0wnZn(Lk)\displaystyle\sum_{n\geq 1}w^{n}f(k)Z_{n-1}(L-k)=wf(k)\sum_{n\geq 0}w^{n}Z_{n}(L-k)
=\displaystyle= wf(k)Ztd(w,Lk)|num.\displaystyle wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)_{|{\rm num}}.

2. We also note that

π(k|L)=n1πn(k|L)pn(L),\pi(k|L)=\sum_{n\geq 1}\pi_{n}(k|L)\mathrm{p}_{n}(L), (4.30)

which is a simple consequence of (4.8).

4.4 Mean interval X|L\langle X|L\rangle

This is, by definition,

X|L=k1kπ(k|L).\langle X|L\rangle=\sum_{k\geq 1}k\pi(k|L). (4.31)

Multiplying (4.27) by kk and summing upon kk yields

L1zLX|L|num=wzf~(z)1wf~(z),\sum_{L\geq 1}z^{L}\langle X|L\rangle_{|{\rm num}}=\frac{wz\tilde{f}^{\prime}(z)}{1-w\tilde{f}(z)}, (4.32)

which can also be obtained by taking the derivative with respect to zz of the expression for the inverse moment 1/NL\langle 1/N_{L}\rangle. Indeed,

X|L|num=LNL|num=L[ln(1wf~(z))]L,\langle X|L\rangle_{|{\rm num}}=\Big{\langle}\frac{L}{N_{L}}\Big{\rangle}_{|{\rm num}}=L\left[-\ln(1-w\tilde{f}(z))\right]_{L}, (4.33)

then, taking the generating function of the right side gives

L0zLX|L|num\displaystyle\sum_{L\geq 0}z^{L}\langle X|L\rangle_{|{\rm num}} =\displaystyle= L1zLL[ln(1wf~(z))]L\displaystyle\sum_{L\geq 1}z^{L}L\left[-\ln(1-w\tilde{f}(z))\right]_{L} (4.34)
=\displaystyle= zddz(ln(1wf~(z)))=wzf~(z)1wf~(z).\displaystyle z\frac{{\rm d}}{{\rm d}z}\left(-\ln(1-w\tilde{f}(z))\right)=\frac{wz\tilde{f}^{\prime}(z)}{1-w\tilde{f}(z)}.

4.5 The longest interval

By definition, the longest interval is

Xmax=max(X1,,XNL).X_{{\rm max}}={\rm max}(X_{1},\dots,X_{N_{L}}). (4.35)

Its distribution function is defined as

F(k|L)=Prob(Xmaxk|L)=n0k1=1kkn=1kp({ki},n|L)=F(k|L)|numZtd(w,L),F(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}\leq k|L)=\sum_{n\geq 0}\sum_{k_{1}=1}^{k}\dots\sum_{k_{n}=1}^{k}p(\{k_{i}\},n|L)=\frac{F(k|L)_{|{\rm num}}}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}, (4.36)

with initial value

F(k|0)|num=1.F(k|0)_{|{\rm num}}=1. (4.37)

The numerator in (4.36) reads

F(k|L)|num\displaystyle F(k|L)_{|{\rm num}} =\displaystyle= n0wnk1=1kkn=1kpn({ki}|L)|num\displaystyle\sum_{n\geq 0}w^{n}\sum_{k_{1}=1}^{k}\dots\sum_{k_{n}=1}^{k}p_{n}(\{k_{i}\}|L)_{|{\rm num}} (4.38)
=\displaystyle= n0wnFn(k|L)|num,\displaystyle\sum_{n\geq 0}w^{n}F_{n}(k|L)_{|{\rm num}},

where Fn(k|L)|numF_{n}(k|L)_{|{\rm num}} is defined in (2.12). Note that F(L|L)|num=Ztd(w,L)F(L|L)_{|{\rm num}}=Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L), hence F(L|L)=1F(L|L)=1. The generating function of the numerator is

L0zLF(k|L)|num\displaystyle\sum_{L\geq 0}z^{L}F(k|L)_{|{\rm num}} =\displaystyle= 1+n1i=1n(ki=1kwf(ki)zki)\displaystyle 1+\sum_{n\geq 1}\prod_{i=1}^{n}\Big{(}\sum_{k_{i}=1}^{k}wf(k_{i})z^{k_{i}}\Big{)} (4.39)
=\displaystyle= 1+n1(wf~(z,k))n=11wf~(z,k),\displaystyle 1+\sum_{n\geq 1}\Big{(}w\tilde{f}(z,k)\Big{)}^{n}=\frac{1}{1-w\tilde{f}(z,k)},

where

f~(z,k)=j=1kzjf(j).\tilde{f}(z,k)=\sum_{j=1}^{k}z^{j}f(j). (4.40)

The numerator obeys the recursion (renewal) equation, which generalises the Rosén-Wendel’s result (2.4) of wendel1 ,

F(k|L)|num=j=1min(k,L)wf(j)F(k|Lj)|num,F(k|L)_{|{\rm num}}=\sum_{j=1}^{\min(k,L)}wf(j)\,F(k|L-j)_{|{\rm num}}, (4.41)

with initial condition (4.37).

The distribution of XmaxX_{\rm max} is given by the difference

p(k|L)=Prob(Xmax=k)=F(k|L)F(k1|L),p(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}=k)=F(k|L)-F(k-1|L), (4.42)

where F(0|L)=δ(L,0)F(0|L)=\delta(L,0), with generating function

L0zLp(k|L)|num\displaystyle\sum_{L\geq 0}z^{L}p(k|L)_{|{\rm num}} =\displaystyle= 11wf~(z,k)11wf~(z,k1)\displaystyle\frac{1}{1-w\tilde{f}(z,k)}-\frac{1}{1-w\tilde{f}(z,k-1)} (4.43)
=\displaystyle= wzkf(k)[1wf~(z,k)][1wf~(z,k1)].\displaystyle\frac{wz^{k}f(k)}{[1-w\tilde{f}(z,k)][1-w\tilde{f}(z,k-1)]}.

Its end point value is the same as π(L|L)\pi(L|L) (4.25), i.e.,

p(L|L)=Prob(NL=1)=wf(L)Ztd(w,L).p(L|L)=\mathop{\rm Prob}\nolimits(N_{L}=1)=\frac{wf(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}. (4.44)

Note that (4.43) can be obtained by multiplying (2.17) by wnw^{n} and summing on nn. In other words,

p(k|L)|num=n1wnpn(k|L)|num,p(k|L)_{|{\rm num}}=\sum_{n\geq 1}w^{n}p_{n}(k|L)_{|{\rm num}}, (4.45)

as can also be inferred from (4.38). And therefore (see (4.8))

p(k|L)=n1pn(k|L)pn(L).p(k|L)=\sum_{n\geq 1}p_{n}(k|L)\mathrm{p}_{n}(L). (4.46)

When Xmax=k>L/2X_{\rm max}=k>L/2, the longest interval is unique. Denoting the restriction of p(k|L)p(k|L) to the range k>L/2k>L/2 by q(k|L)q(k|L), and generalising the reasoning made in wendel ; wendel1 we can decompose a configuration into three contributions to obtain

q(k|L)|num\displaystyle q(k|L)_{|{\rm num}} =\displaystyle= i=0LkZtd(w,i)wf(k)Ztd(w,Lki)\displaystyle\sum_{i=0}^{L-k}Z^{\mathop{\scriptstyle\mathrm{td}}}(w,i)wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k-i) (4.47)
=\displaystyle= wf(k)(ZtdZtd)(w,Lk),\displaystyle wf(k)(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L-k),

which, using (4.17), can be alternatively written as

q(k|L)=wf(k)Ztd(w,Lk)Ztd(w,L)(1+NLk),q(k|L)=\frac{wf(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}(1+\langle N_{L-k}\rangle), (4.48)

that is (see (4.26)),

q(k|L)=π(k|L)(1+NLk).q(k|L)=\pi(k|L)(1+\langle N_{L-k}\rangle). (4.49)

For k=Lk=L, (4.44) with (4.25) is recovered.

Remarks

1. An alternative route to (4.49) can be inferred from (4.45) as follows. We have

q(k|L)|num=n1wnqn(k|L)|num=n1wnnπn(k|L)|num,q(k|L)_{|{\rm num}}=\sum_{n\geq 1}w^{n}q_{n}(k|L)_{|{\rm num}}=\sum_{n\geq 1}w^{n}n\pi_{n}(k|L)_{|{\rm num}}, (4.50)

where πn(k|L)\pi_{n}(k|L) is given in (2.8). So

q(k|L)|num\displaystyle q(k|L)_{|{\rm num}} =\displaystyle= wf(k)n1nwn1Zn1(Lk)\displaystyle wf(k)\sum_{n\geq 1}nw^{n-1}Z_{n-1}(L-k) (4.51)
=\displaystyle= wf(k)n0(wnZn(Lk)+nwnZn(Lk))\displaystyle wf(k)\sum_{n\geq 0}(w^{n}Z_{n}(L-k)+nw^{n}Z_{n}(L-k))
=\displaystyle= wf(k)(Ztd(w,Lk)+NLk|num),\displaystyle wf(k)(Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)+\langle N_{L-k}\rangle_{|{\rm num}}),

which, after division by Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L), yields (4.49).

2. Comparison between (4.51) and (4.47) entails the equality

(ZtdZtd)(w,L)=n1nwn1Zn1(L),(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)=\sum_{n\geq 1}nw^{n-1}Z_{n-1}(L), (4.52)

which can also be checked directly by taking the generating functions of both sides,

Z~td(w,z)2=n1nwn1f~(z)n1=1(1w~f(z))2.\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)^{2}=\sum_{n\geq 1}nw^{n-1}\tilde{f}(z)^{n-1}=\frac{1}{(1-\tilde{w}f(z))^{2}}. (4.53)

3. From (2.21) we infer that, if k>L/2k>L/2,

Ftd(k|Lk)|num=Ztd(w,Lk).F^{\mathop{\scriptstyle\mathrm{td}}}(k|L-k)_{|{\rm num}}=Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k). (4.54)

5 Phase transition for tied-down renewal processes

In this section and in the following sections 6, 7 and 8 the distribution f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k) is taken subexponential with asymptotic power-law decay (1.1). Before discussing the phase diagram of the process, we give some illustrative examples of such distributions.

5.1 Illustrative examples

In the sequel, we shall illustrate the general results derived for tdrp in the current section, and for free renewal processes in section 9, on the following examples.

Example 1. This first example corresponds to the tied-down random walk of figure 3 on which we come back in more detail. The distribution of the size of intervals, f(k)f(k), representing the probability of first return at the origin of the walk after 2k2k steps, or equivalently after kk tick marks on figure 3, reads

f(k)=122k1(2k2)!(k1)!k!=Γ(k1/2)2πΓ(k+1)12πk3/2,f(k)=\frac{1}{2^{2k-1}}\frac{(2k-2)!}{(k-1)!k!}=\frac{\Gamma(k-1/2)}{2\sqrt{\pi}\Gamma(k+1)}\approx\frac{1}{2\sqrt{\pi}k^{3/2}}, (5.1)

since the number of such walks is equal to (2k2)!/[(k1)!k!](2k-2)!/[(k-1)!k!]. Its generating function reads

f~(z)=11z.\tilde{f}(z)=1-\sqrt{1-z}. (5.2)

The partition function (4.6) for w=1w=1 represents the probability that the walk returns at the origin after 2L2L steps, or equivalently after LL tick marks,

Ztd(1,L)=122L(2LL)1πL,Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)=\frac{1}{2^{2L}}\left(2L\atop L\right)\approx\frac{1}{\sqrt{\pi L}}, (5.3)

since the number of such walks is equal to (2L)!/(L!)2(2L)!/(L!)^{2}. Its generating function reads

Z~td(1,z)=11z.\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(1,z)=\frac{1}{\sqrt{1-z}}. (5.4)

Note that

f(L)=Ztd(1,L1)Ztd(1,L).f(L)=Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L-1)-Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L). (5.5)

The partition function Zn(L)Z_{n}(L) is explicit for this case,

Zn(L)=n22Ln(2Ln1)!L!(Ln)!n2πL3/2,Z_{n}(L)=\frac{n}{2^{2L-n}}\frac{(2L-n-1)!}{L!(L-n)!}\approx\frac{n}{2\sqrt{\pi}L^{3/2}}, (5.6)

with nLn\leq L.

Example 2. This second example is defined for any θ>0{\theta}>0 by

f(k)=1ζ(1+θ)1k1+θ,k>0.f(k)=\frac{1}{\zeta(1+{\theta})}\frac{1}{k^{1+{\theta}}},\quad k>0. (5.7)

So

f~(z)=Li1+θ(z)ζ(1+θ),\tilde{f}(z)=\frac{\mathrm{Li}_{1+{\theta}}(z)}{\zeta(1+{\theta})}, (5.8)

where Lis(z)=k1zk/ks\mathrm{Li}_{s}(z)=\sum_{k\geq 1}z^{k}/k^{s} is the polylogarithm function. If θ>1{\theta}>1 the mean X=ζ(θ)/ζ(1+θ)\langle X\rangle=\zeta({\theta})/\zeta(1+{\theta}). This is the distribution used, e.g., in burda ; burda3 ; bar2 ; bar3 ; barma ; glzeta .

5.2 Phase diagram

Demonstrating the existence of a phase transition in the model defined by (4.2), with distribution f(k)f(k) given by (1.1), when ww crosses the value one, is a classical subject. This model is a particular instance of a linear system, as described in fisher , where the mechanism of the transition is explained in simple terms. This transition is also studied in burda3 ; bar2 ; bar3 ; barma for Example 2 (see (5.7)). Let us first analyse the large LL behaviour of Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L). Recalling (4.5) we have, for a contour encircling the origin,

Ztd(w,L)=dz2πiZ~td(w,z)zL+1=dz2πizL+111wf~(z).Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)=\oint\frac{{\rm d}z}{2\pi{\rm i}}\frac{\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)}{z^{L+1}}=\oint\frac{{\rm d}z}{2\pi{\rm i}\,z^{L+1}}\frac{1}{1-w\tilde{f}(z)}. (5.9)

Since f~(z)\tilde{f}(z) is monotonically increasing for z(0,1)z\in(0,1) the denominator of Z~td(w,z)\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z), 1wf~(z)1-w\tilde{f}(z), is monotonically decreasing between 1 and 1w1-w.

Disordered phase. If w>1w>1, the denominator vanishes for z=z0<1z=z_{0}<1 such that wf~(z0)=1w\tilde{f}(z_{0})=1, hence Z~td(w,z)\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z) has a pole at z0z_{0}, and therefore Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) is exponentially increasing,

Ztd(w,L)z0Lwz0f~(z0).Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)\approx\frac{z_{0}^{-L}}{wz_{0}\tilde{f}^{\prime}(z_{0})}. (5.10)

Critical regime. If w=1w=1, then z0=1z_{0}=1. The asymptotic estimates of Ztd(1,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L) are given in (7.5) and (7.23).

Condensed phase. If w<1w<1, the denominator 1wf~(z)1-w\tilde{f}(z) has no zero, but it is singular for z=z0=1z=z_{0}=1 (which is the singularity of f~(z)\tilde{f}(z)). Hence z0z_{0} sticks to 1. The asymptotic estimate of Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) is given in (8.1).

This is the switch mechanism of Fisher fisher : the condition determining z0z_{0} switches from z0z_{0} being the smallest root of the equation 1wf~(z)=01-w\tilde{f}(z)=0 to being the closest real singularity of f~(z)\tilde{f}(z), which is a cut at z=z0=1z=z_{0}=1. This non analytical switch signals the phase transition. The free energy density fisher

𝐟=limL1LlnZtd(w,L)=lnz0,\mathbf{f}=\lim_{L\to\infty}-\frac{1}{L}\ln Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)=\ln z_{0}, (5.11)

therefore vanishes when w1w\leq 1. The three cases above are successively reviewed in the next sections.

6 Disordered phase (w>1w>1) for tied-down renewal processes

The asymptotic expression at large LL of the distribution of the size of a generic interval is obtained by carrying (5.10) in (4.26), which leads to

π(k|L)wf(k)z0k=wf(k)ek/ξ,ξ=1|lnz0|,\pi(k|L)\approx wf(k)z_{0}^{k}=wf(k){\rm e}^{-k/\xi},\qquad\xi=\frac{1}{|\ln z_{0}|}, (6.1)

where ξ\xi is the correlation length, divergent at the transition. This expression is independent of LL and normalised, since summing on kk restores wf~(z0)=1w\tilde{f}(z_{0})=1. This exponentially decaying distribution has a finite mean,

X|Lwz0f~(z0),\langle X|L\rangle\approx wz_{0}\tilde{f}^{\prime}(z_{0}), (6.2)

an expression which can also be inferred from (4.32). Thus (5.10) can be recast as

Ztd(w,L)z0LX|L.Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)\approx\frac{z_{0}^{-L}}{\langle X|L\rangle}. (6.3)

The distribution of NLN_{L} is given by (4.9)

pn(L)=wnZn(L)Ztd(w,L)wnZn(L)wf~(z0)z0L+1.\mathrm{p}_{n}(L)=\frac{w^{n}Z_{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\approx w^{n}Z_{n}(L)\,w\tilde{f}^{\prime}(z_{0})z_{0}^{L+1}. (6.4)

This distribution obeys the central limit theorem, as illustrated on the example below. Using (4.15), we obtain the asymptotic expression of NL\langle N_{L}\rangle,

NLLwz0dz0dwLX|L,\langle N_{L}\rangle\approx-L\frac{w}{z_{0}}\frac{{\rm d}z_{0}}{{\rm d}w}\approx\frac{L}{\langle X|L\rangle}, (6.5)

which means that

1NL1NL.\frac{1}{\langle N_{L}\rangle}\approx\left\langle\frac{1}{N_{L}}\right\rangle. (6.6)

Let us denote the density of points (or intervals) for a finite system as

νL=NLL,\nu_{L}=\frac{\langle N_{L}\rangle}{L}, (6.7)

then, asymptotically, we have

ν=limLνL=limLNLL=limL1X|L.\nu=\lim_{L\to\infty}\nu_{L}=\lim_{L\to\infty}\frac{\langle N_{L}\rangle}{L}=\lim_{L\to\infty}\frac{1}{\langle X|L\rangle}. (6.8)

We illustrate these general statements on Example 1 (see (5.1)), for which z0z_{0} is explicit,

z0=2w1w2,z_{0}=\frac{2w-1}{w^{2}}, (6.9)

hence

ξ(w1)2,\xi\approx(w-1)^{-2}, (6.10)

and the following asymptotic expressions hold,

Ztd(w,L)\displaystyle Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) \displaystyle\approx 2(w1)w2L(2w1)L+1,\displaystyle\frac{2(w-1)w^{2L}}{(2w-1)^{L+1}},
X|L\displaystyle\langle X|L\rangle L\displaystyle\mathrel{\mathop{\to}\limits_{L\to\infty}} 2w12(w1),\displaystyle\frac{2w-1}{2(w-1)},
NL\displaystyle\langle N_{L}\rangle \displaystyle\approx LX|L+w(w1)(2w1),\displaystyle\frac{L}{\langle X|L\rangle}+\frac{w}{(w-1)(2w-1)},
VarNL\displaystyle\mathop{\rm Var}\nolimits N_{L} \displaystyle\approx L2w(2w1)2w(2w21)(2w1)2(w1)2,\displaystyle L\frac{2w}{(2w-1)^{2}}-\frac{w(2w^{2}-1)}{(2w-1)^{2}(w-1)^{2}},
pn(L)\displaystyle\mathrm{p}_{n}(L) \displaystyle\approx 12πVarNLexp((nNL)22VarNL),\displaystyle\frac{1}{\sqrt{2\pi\mathop{\rm Var}\nolimits{N_{L}}}}\exp\Big{(}-\frac{(n-\langle N_{L}\rangle)^{2}}{2\mathop{\rm Var}\nolimits{N_{L}}}\Big{)},
ν\displaystyle\nu =\displaystyle= 2(w1)2w1.\displaystyle\frac{2(w-1)}{2w-1}. (6.11)

Figure 6 depicts a comparison between the exact finite-size expression of the density νL\nu_{L} obtained by means of (4.14) for L=1000L=1000 as a function of ww, and the asymptotic expression (6.11). It vanishes at the transition w=1w=1, where the system becomes critical.

More generally, if θ<1{\theta}<1, close to the transition, we get

ν(w1)1/θ1,\nu\sim(w-1)^{1/{\theta}-1}, (6.12)

as can be easily inferred by means of the expansion (7.1), a result already present in burda3 , later recovered in bar2 .

Refer to caption
Figure 6: tdrp: density of intervals ν\nu against ww in the disordered phase, for Example 1 (see (5.1)). Exact refers to the finite-size expression (6.7) where NL\langle N_{L}\rangle is extracted from (4.14) with L=1000L=1000, asymptotic to (6.11).

If θ>1{\theta}>1, the density ν\nu tends to 1/X1/\langle X\rangle when w1w\to 1, as can be seen on (6.2) and (6.8), using the expansion (7.2) (see also (7.24)). The density vanishes in the condensed phase since NL\langle N_{L}\rangle is finite (see (8.5)), it is therefore discontinuous at the transition, as noted in burda3 ; bar2 . Likewise, it is easy to see that

ξ1{(w1)1/θθ<1w1θ>1.\xi^{-1}\sim\left\{\begin{array}[]{ll}(w-1)^{1/{\theta}}&{\theta}<1\vspace{12pt}\\ w-1&{\theta}>1.\end{array}\right. (6.13)

The correlation length diverges at the transition, while the order parameter ν\nu is either continuous (θ<1{\theta}<1) or discontinuous (θ>1{\theta}>1), as seen above. The transition is therefore of mixed order burda3 ; bar2 .

Finally we note that the intervals XiX_{i} behave essentially as iid random variables, with distribution (6.1), hence the statistics of the longest interval belongs to the Gumbel class gumbel ; gnedenko . This is detailed on Example 2 in bar3 .

7 Critical regime (w=1w=1) for tied-down renewal processes

In this regime, the behaviour of the quantities of interest strongly depends on whether the index θ{\theta} is smaller or larger than unity. The discussion below is organised accordingly. Part of the material of this section can be found in more details in wendel ; wendel1 and is also addressed in bar3 ; barma . Here we summarise these former studies and complement them by a detailed analysis of the distribution of the number of intervals NLN_{L} and of the distribution π(k|L)\pi(k|L) of the size of a generic interval. We also come back on the distribution of the longest interval.

If w=1w=1 the singularity is at z=1z=1, or, setting z=esz={\rm e}^{-s}, at s=0s=0. The generating function f~(z)\tilde{f}(z) becomes the Laplace transform f^(s)\hat{f}(s) which has the expansion

f^(s)s01|a|sθ,\displaystyle\hat{f}(s)\mathrel{\mathop{\approx}\limits_{s\to 0}}1-|a|s^{\theta},\qquad θ<1\displaystyle{\theta}<1 (7.1)
f^(s)s01sX++asθ,\displaystyle\hat{f}(s)\mathrel{\mathop{\approx}\limits_{s\to 0}}1-s\langle X\rangle+\cdots+as^{\theta},\qquad θ>1\displaystyle{\theta}>1 (7.2)

with

a=cΓ(θ)=θΓ(θ)k0θ,a=c\,\Gamma(-{\theta})={\theta}\Gamma(-{\theta})k_{0}^{\theta}, (7.3)

i.e., c=θk0θc={\theta}k_{0}^{\theta}, where k0k_{0} is a microscopic scale, defined as

g(k)=j>kf(j)k(k0k)θ.g(k)=\sum_{j>k}f(j)\mathrel{\mathop{\approx}\limits_{k\to\infty}}\Big{(}\frac{k_{0}}{k}\Big{)}^{\theta}. (7.4)

The parameter aa is negative if 0<θ<10<\theta<1, positive if 1<θ<21<\theta<2, and so on. For instance, Γ(1/2)=2π\Gamma(-1/2)=-2\sqrt{\pi}, Γ(3/2)=4π/3\Gamma(-3/2)=4\sqrt{\pi}/3, Γ(5/2)=8π/15\Gamma(-5/2)=-8\sqrt{\pi}/15, and so on.

7.1 Distribution f(k)f(k) with index θ<1{\theta}<1

Since Z~td(1,z)=1/(1f~(z))\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(1,z)=1/(1-\tilde{f}(z)), in Laplace space we have Z^td(1,s)1/asθ\hat{Z}^{\mathop{\scriptstyle\mathrm{td}}}(1,s)\approx 1/as^{{\theta}} which yields the expression of the partition function (see (4.6) in wendel1 )

Ztd(1,L)LθsinπθπcLθ1.Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{{\theta}\sin\pi{\theta}}{\pi c}L^{{\theta}-1}. (7.5)

For instance, setting θ=1/2{\theta}=1/2 and c=1/(2π)c=1/(2\sqrt{\pi}) restores (5.3).

7.1.1 The number of intervals

We have (see (4.10) in wendel1 ),

NLA(θ)cLθ,A(θ)=Γ(1+θ)Γ(1θ)Γ(2θ),\langle N_{L}\rangle\approx\frac{A({\theta})}{c}L^{\theta},\qquad A({\theta})=\frac{\Gamma(1+{\theta})}{\Gamma(1-{\theta})\Gamma(2{\theta})}, (7.6)

which can be easily deduced from (4.14). For the specific case of Example 1 (see (5.1)), we have the exact result (see (2.47) in wendel1 )

NL=1Ztd(1,L)1=22L(2LL)1πL,\langle N_{L}\rangle=\frac{1}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}-1=\frac{2^{2L}}{\left(2L\atop L\right)}-1\approx\sqrt{\pi L}, (7.7)

which is in agreement with (7.6), with θ=1/2,c=1/(2π){\theta}=1/2,c=1/(2\sqrt{\pi}). We know from (4.9) that the distribution of NLN_{L} is given by the ratio

pn(L)=Zn(L)Ztd(1,L).\mathrm{p}_{n}(L)=\frac{Z_{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}. (7.8)

For nn and LL large pn(L)\mathrm{p}_{n}(L) has a scaling form. On the one hand, according to the generalised central limit theorem, the scaling form of the numerator is given by

Zn(L)1n1/θθ,c(Ln1/θ),Z_{n}(L)\approx\frac{1}{n^{1/{\theta}}}\mathcal{L}_{{\theta},c}\left(\frac{L}{n^{1/{\theta}}}\right), (7.9)

where θ,c\mathcal{L}_{{\theta},c} is the density of the stable law of index θ{\theta}, tail parameter cc and asymmetry parameter β=1\beta=1 (see e.g., cg2019 ). Then using (7.5), we get, with u=L/n1/θu=L/n^{1/{\theta}},

pn(L)πcθsinπθ1LθLn1/θθ,c(Ln1/θ)πcθsinπθ1Lθuθ,c(u).\mathrm{p}_{n}(L)\approx\frac{\pi c}{{\theta}\sin\pi{\theta}}\frac{1}{L^{\theta}}\frac{L}{n^{1/{\theta}}}\mathcal{L}_{{\theta},c}\left(\frac{L}{n^{1/{\theta}}}\right)\approx\frac{\pi c}{{\theta}\sin\pi{\theta}}\frac{1}{L^{\theta}}u\mathcal{L}_{{\theta},c}(u). (7.10)

The Lévy distribution of index θ=1/2{\theta}=1/2 has the explicit expression

1/2,c(u)=ceπc2/uu3/2,\mathcal{L}_{1/2,c}(u)=\frac{c\,{\rm e}^{-\pi c^{2}/u}}{u^{3/2}}, (7.11)

hence (see (2.49) in wendel1 ), for Example 1 (see (5.1)),

pn(L)v2Lev2/4,v=1u=nL.\mathrm{p}_{n}(L)\approx\frac{v}{2\sqrt{L}}{\rm e}^{-v^{2}/4},\quad v=\frac{1}{\sqrt{u}}=\frac{n}{\sqrt{L}}. (7.12)

Moreover, for this example, for nn and LL finite, pn(L)\mathrm{p}_{n}(L) is explicit since both Zn(L)Z_{n}(L) given by (5.6) and Ztd(1,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L), given by (5.3), are known explicitly.

Refer to caption
Figure 7: tdrp: single interval distribution π(k|L)\pi(k|L), with L=100L=100, for Example 1 (see (5.1)), at criticality (w=1w=1). Exact refers to the middle expression in (7.13), asymptotic to the rightmost one. The yy-axis is on a logarithmic scale.

7.1.2 Single interval distribution

Two regimes are to be considered.

(i) In all regimes where =Lk\ell=L-k is large, using (7.5) we have

π(k|L)=f(k)Ztd(1,Lk)Ztd(1,L)f(k)(1kL)θ1.\pi(k|L)=f(k)\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L-k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx f(k)\left(1-\frac{k}{L}\right)^{{\theta}-1}. (7.13)

For instance, if 1kL1\ll k\ll L,

π(k|L)f(k)ck1+θ,\pi(k|L)\approx f(k)\approx\frac{c}{k^{1+{\theta}}}, (7.14)

while if k=λLk={\lambda}L, with λ(0,1){\lambda}\in(0,1),

π(k|L)cλ1+θ(1λ)1θ1L1+θ,\pi(k|L)\approx\frac{c}{{\lambda}^{1+{\theta}}(1-{\lambda})^{1-{\theta}}}\frac{1}{L^{1+{\theta}}}, (7.15)

which is minimum at λ=(1+θ)/2{\lambda}=(1+{\theta})/2.

(ii) On the other hand, if =Lk=O(1)\ell=L-k=O(1),

π(L|L)=f(L)Ztd(1,)Ztd(1,L)πc2Ztd(1,)θsinπθ1L2θ.\pi(L-\ell|L)=f(L-\ell)\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,\ell)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx\frac{\pi c^{2}Z^{\mathop{\scriptstyle\mathrm{td}}}(1,\ell)}{{\theta}\sin\pi{\theta}}\frac{1}{L^{2{\theta}}}. (7.16)

In particular, for k=Lk=L,

π(L|L)=f(L)Ztd(1,L)πc2θsinπθ1L2θ.\pi(L|L)=\frac{f(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx\frac{\pi c^{2}}{{\theta}\sin\pi{\theta}}\frac{1}{L^{2{\theta}}}. (7.17)

In this regime the ratio of π(k|L)\pi(k|L) to the estimate (7.13) reads

π(L|L)f(L)(1k/L)θ1=πc1θZtd(1,)θsinπθ,\frac{\pi(L-\ell|L)}{f(L-\ell)(1-k/L)^{{\theta}-1}}=\frac{\pi c\,\ell^{1-{\theta}}Z^{\mathop{\scriptstyle\mathrm{td}}}(1,\ell)}{{\theta}\sin\pi{\theta}}, (7.18)

which tends to one when \ell becomes large, if one refers to (7.5).

All these results can be illustrated on Example 1 (see (5.1)). For instance figure 7 gives a comparison between the exact expression of π(k|L)\pi(k|L) computed for L=100L=100 by means of the middle expression in (7.13) and its asymptotic form given by the rightmost expression in (7.13).

The generating function of the mean interval X|L\langle X|L\rangle given in (4.32) yields the estimate, in Laplace space,

L1zLX|L|numf^(s)1f^(s)θs,\sum_{L\geq 1}z^{L}\langle X|L\rangle_{|{\rm num}}\approx-\frac{\hat{f}^{\prime}(s)}{1-\hat{f}(s)}\approx\frac{{\theta}}{s}, (7.19)

hence (see (4.18) in wendel1 )

X|LθZtd(1,L)πcsinπθL1θ.\langle X|L\rangle\approx\frac{{\theta}}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx\frac{\pi c}{\sin\pi{\theta}}L^{1-{\theta}}. (7.20)

This result can be recovered by taking the average of the estimate (7.13). It predicts correctly that the product X|LNLL\langle X|L\rangle\langle N_{L}\rangle\sim L. For Example 1, (see (5.1)), the computation leads to the exact result (see (2.58) in wendel1 )

X|L=12Ztd(1,L)πL2.\langle X|L\rangle=\frac{1}{2Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx\frac{\sqrt{\pi L}}{2}. (7.21)

7.1.3 The longest interval

The study of the statistics of the longest interval for the critical case, including the scaling analysis of p(k|L)p(k|L) for θ<1{\theta}<1, is done in wendel ; wendel1 (see also bar3 for a study of Example 2).

For k>L/2k>L/2, in contrast with the case of random allocation models and zrp where p(k|L)=q(k|L)=nπ(k|L)p(k|L)=q(k|L)=n\pi(k|L) (see §2.1.2), the ‘enhancement factor’ nn is now replaced by the factor 1+NLk1+\langle N_{L-k}\rangle (see (4.49)), which is equal to one for k=Lk=L, where p(L|L)=π(L|L)p(L|L)=\pi(L|L) are equal, see (4.44). Using (7.6) for NL\langle N_{L}\rangle, and (7.13) for π(k|L)\pi(k|L) in (4.49) allows to recover the universal scaling expression valid for k>L/2k>L/2 (see equation (4.48) in wendel1 ), in the limit where 1kL1\ll k\sim L, with r=k/Lr=k/L kept fixed,

q(k|L)1LA(θ)r1+θ(1r)12θ,q(k|L)\approx\frac{1}{L}\frac{A({\theta})}{r^{1+{\theta}}(1-r)^{1-2{\theta}}}, (7.22)

where A(θ)A({\theta}) is given in (7.6).

Though there is no condensation at criticality, some features are precursors of this phenomenon. For instance, the mean longest interval Xmax\langle X_{\rm max}\rangle scales as LL while the typical interval X|L\langle X|L\rangle scales as L1θL^{1-{\theta}}. However, not only XmaxX(1)X_{\rm max}\equiv X^{(1)} scales as LL but also all the following maxima X(r)X^{(r)} (k=2,3,k=2,3,\dots) do so wendel ; wendel1 . Moreover XmaxX_{\rm max} continues to fluctuate when LL\to\infty while for genuine condensation as in section 2 above or in section 8 below, its distribution is peaked. Finally, the dominant contribution to the weight of π(k|L)\pi(k|L) comes from values of kk less than a small cutoff.

7.2 Distribution f(k)f(k) with index θ>1{\theta}>1

For θ>1{\theta}>1, we have, using (4.5) (see (4.73) in wendel1 ),

Ztd(1,L)1X+cθ(θ1)X2L1θ.Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)\approx\frac{1}{\langle X\rangle}+\frac{c}{{\theta}({\theta}-1)\langle X\rangle^{2}}L^{1-{\theta}}. (7.23)

The average value of NLN_{L} is obtained by means of (4.14)444Equation (7.24) corrects the inaccurate expression (4.74) given in wendel1 for this quantity.

NL{LX+c(θ1)(2θ)X2L2θ1<θ<2LX+VarXX2θ>2\langle N_{L}\rangle\approx\left\{\begin{array}[]{ll}\frac{L}{\langle X\rangle}+\frac{c}{({\theta}-1)(2-{\theta})\langle X\rangle^{2}}L^{2-{\theta}}&1<{\theta}<2\vspace{12pt}\\ \frac{L}{\langle X\rangle}+\frac{\mathop{\rm Var}\nolimits{X}}{\langle X\rangle^{2}}&{\theta}>2\end{array}\right. (7.24)

The subleading correction in the second line (i.e., for θ>2{\theta}>2) is given by the correction term of the first line which is now negative and decreasing. The distribution of NLN_{L} reads

pn(L)=Zn(L)Ztd(1,L)XZn(L).\mathrm{p}_{n}(L)=\frac{Z_{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx\langle X\rangle Z_{n}(L). (7.25)

The asymptotic estimate for X|L\langle X|L\rangle is obtained by analysing (4.32), yielding for θ>1{\theta}>1,

X|LXcθ1L1θ,\langle X|L\rangle\approx\langle X\rangle-\frac{c}{{\theta}-1}L^{1-{\theta}}, (7.26)

which is the same expression as for a free renewal process gl2001 . The single interval distribution has the form

π(k|L)=f(k)Ztd(1,Lk)Ztd(1,L)Lf(k),\pi(k|L)=f(k)\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L-k)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\mathrel{\mathop{\approx}\limits_{L\to\infty}}f(k), (7.27)

except for LkL-k finite, where in particular,

π(L|L)=f(L)Ztd(1,L)f(L)X.\pi(L|L)=\frac{f(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L)}\approx f(L)\langle X\rangle. (7.28)

The distribution of the longest interval is analysed in wendel1 (see also bar3 for the case of Example 2 defined in (5.7)). The result is

F(k|L)e[L/X](k0/k)θ,F(k|L)\approx{\rm e}^{-[L/\langle X\rangle](k_{0}/k)^{\theta}}, (7.29)

where k0k_{0} is related to the tail coefficient by c=θk0θc={\theta}k_{0}^{\theta}. Setting

Xmax=k0(LX)1/θYL,X_{\rm max}=k_{0}\left(\frac{L}{\langle X\rangle}\right)^{1/\theta}Y_{L}, (7.30)

we have, as LL\to\infty, YLYFY_{L}\to Y^{F}, with limiting distribution

Prob(YF<x)=e1/xθ,\mathop{\rm Prob}\nolimits(Y^{F}<x)={\rm e}^{-1/x^{\theta}}, (7.31)

which is the Fréchet law frechet ; gnedenko . Therefore

Xmaxk0(LX)1/θYFΓ(11/θ),\langle X_{{\rm max}}\rangle\approx k_{0}\left(\frac{L}{\langle X\rangle}\right)^{1/\theta}\underbrace{\langle Y^{F}\rangle}_{\Gamma(1-1/\theta)}, (7.32)

as was already the case for free renewal processes gms2015 .

8 Condensed phase (w<1w<1) for tied-down renewal processes

The aim of this section—central in the present work—is to investigate the statistics of the number of intervals and characterise the fluctuations of the condensate. We start by analysing the large LL behaviour of the quantities of interest which are functions of LL only (partition function, moments and distribution of NLN_{L}). We then investigate the regimes for the distributions of the size of a generic interval, π(k|L)\pi(k|L), and of the longest one, p(k|L)p(k|L). Related material can be found in gia1 ; bar3 ; barma .

8.1 Asymptotic estimates at large LL

Starting from (4.5) and linearising with respect to the singular part, we obtain, when LL\to\infty, for any value of θ{\theta},

Ztd(w,L)w(1w)2cL1+θw(1w)2f(L).Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)\approx\frac{w}{(1-w)^{2}}\frac{c}{L^{1+{\theta}}}\approx\frac{w}{(1-w)^{2}}f(L). (8.1)

Alternatively, it suffices to notice that, for nn fixed and LL large, for any subexponential distribution chistyakov ,

Zn(L)nf(L),Z_{n}(L)\approx nf(L), (8.2)

as for example in (5.6), which entails

Ztd(w,L)=n0wnZn(L)n0wnnf(L),Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)=\sum_{n\geq 0}w^{n}Z_{n}(L)\approx\sum_{n\geq 0}w^{n}nf(L), (8.3)

hence restores (8.1). Likewise, we find

(ZtdZtd)(w,L)2w(1w)3f(L).(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)\approx\frac{2w}{(1-w)^{3}}f(L). (8.4)

If one substitutes (8.1) in the expression of the mean NL\langle N_{L}\rangle (4.15), we obtain, the superuniversal result, independent of θ{\theta},

NLL1+w1w.\langle N_{L}\rangle\mathrel{\mathop{\to}\limits_{L\to\infty}}\frac{1+w}{1-w}. (8.5)

This result can be found alternatively using (4.17) together with (8.1) and (8.4), or else using (8.7) below.

Refer to caption
Figure 8: tdrp: superuniversal asymptotic distribution pn\mathrm{p}_{n} of the number NLN_{L} of intervals in the condensed phase for w=0.6w=0.6. Exact refers to pn(L)\mathrm{p}_{n}(L) extracted from (4.12) or (4.20) for L=200L=200 with Example 1 (see (5.1)), asymptotic refers to (8.7). For this value of ww, the asymptotic average NL=4\langle N_{L}\rangle=4.

More generally, this superuniversality also holds for the asymptotic distribution of NLN_{L}. Using (4.20), the latter reads

vNL=n0vnpn(L)=Ztd(vw,L)Ztd(w,L)y(1w)2(1vw)2,\langle v^{N_{L}}\rangle=\sum_{n\geq 0}v^{n}\mathrm{p}_{n}(L)=\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(vw,L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\approx\frac{y(1-w)^{2}}{(1-vw)^{2}}, (8.6)

hence extracting the coefficient of order nn in yy of this expression leads to the asymptotic distribution, independent of θ{\theta},

pn(L)Lpn=n(1w)2wn1.\mathrm{p}_{n}(L)\mathrel{\mathop{\to}\limits_{L\to\infty}}\mathrm{p}_{n}=n(1-w)^{2}w^{n-1}. (8.7)

This distribution is depicted in figure 8. The same result can be found by noting that

pn(L)=wnZn(L)Ztd(w,L)Zn(L)f(L)(1w)2wn1Ln(1w)2wn1,\mathrm{p}_{n}(L)=\frac{w^{n}Z_{n}(L)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)}\approx\frac{Z_{n}(L)}{f(L)}(1-w)^{2}w^{n-1}\mathrel{\mathop{\to}\limits_{L\to\infty}}n(1-w)^{2}w^{n-1}, (8.8)

using (8.2) again. The interpretation of (8.7) is simple: NL1N_{L}-1 is the sum of two independent geometric random variables (see (11.4)), which represent the fluid on either side of the remaining interval, which is the condensate.

The inverse moment 1/NL\langle 1/N_{L}\rangle can be obtained by using (8.7) above,

1NL1w.\Big{\langle}\frac{1}{N_{L}}\Big{\rangle}\to 1-w. (8.9)

As a consequence of (8.9) we have, for any value of θ{\theta},

X|L(1w)L,\langle X|L\rangle\approx(1-w)L, (8.10)

which can also be deduced from the asymptotic analysis of (4.32).

8.2 Regimes for the single interval distribution

For LL large, the asymptotic expression of the single interval distribution is obtained by substituting (8.1) in (4.24), leading to

π(k|L)L(1w)2f(k)Ztd(w,Lk)f(L).\pi(k|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}(1-w)^{2}f(k)\frac{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-k)}{f(L)}. (8.11)

Figure 9 depicts the distribution π(k|L)\pi(k|L) (together with the distribution of the longest interval p(k|L)p(k|L), see §8.3 below), for L=60L=60 and w=0.6w=0.6 computed from (4.26), with Example 1 (see (5.1)). As can be seen on this figure, there are three distinct regions for π(k|L)\pi(k|L), namely, from left to right, a downhill region, followed by a long dip region, then an uphill region which accounts for the fluctuations of the condensate.

Let us discuss the behaviour of π(k|L)\pi(k|L) given by (8.11) in each of these regions successively.

  1. 1.

    Downhill region. For kk finite we have, using again (8.1),

    π(k|L)wf(k).\pi(k|L)\approx wf(k). (8.12)

    Introducing a cutoff Λ\Lambda, such that 1ΛL1\ll\Lambda\ll L, the weight of this downhill region can be estimated as

    k=1Λπ(k|L)k=1Λwf(k)k=1wf(k)=w.\sum_{k=1}^{\Lambda}\pi(k|L)\approx\sum_{k=1}^{\Lambda}wf(k)\to\sum_{k=1}^{\infty}wf(k)=w. (8.13)
  2. 2.

    Dip region. In the dip region, where kk and LkL-k are simultaneously large, setting k=λLk={\lambda}L in (8.12) (0<λ<10<{\lambda}<1) and using (8.1) yields the estimate

    π(k|L)wf(k)f(Lk)f(L)w[λ(1λ)]1+θcL1+θ.\pi(k|L)\approx wf(k)\frac{f(L-k)}{f(L)}\approx\frac{w}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}\frac{c}{L^{1+{\theta}}}. (8.14)

    In this region the distribution is therefore U-shaped: the most probable configurations are those where almost all the particles are located on one of two sites. The dip centred around k=L/2k=L/2 becomes deeper and deeper with LL.

    The weight of the dip region can be estimated using (8.14), as

    k=ΛLΛπ(k|L)Lθwcϵ1ϵdλ[λ(1λ)]1+θ,\sum_{k=\Lambda}^{L-\Lambda}\pi(k|L)\approx L^{-{\theta}}w\,c\int_{{\epsilon}}^{1-{\epsilon}}\frac{{\rm d}{\lambda}}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}, (8.15)

    where, for the sake of simplicity, we chose Λ=ϵL\Lambda={\epsilon}L. The two downhill and uphill regions are therefore well separated by the dip region, as is conspicuous on figure 9. Note the similarity between (8.15) and (3.12), with the correspondence

    11n11NL=w.1-\frac{1}{n}\hookrightarrow 1-\left\langle\frac{1}{N_{L}}\right\rangle=w. (8.16)
  3. 3.

    Uphill region. The uphill region corresponds to =Lk\ell=L-k finite, where (8.11) simplifies into

    π(L|L)(1w)2Ztd(w,).\pi(L-\ell|L)\approx(1-w)^{2}Z^{\mathop{\scriptstyle\mathrm{td}}}(w,\ell). (8.17)

    The weight of this region can be estimated as

    =0Λπ(L|L)\displaystyle\sum_{\ell=0}^{\Lambda}\pi(L-\ell|L) \displaystyle\approx (1w)2=0ΛZtd(w,)\displaystyle(1-w)^{2}\sum_{\ell=0}^{\Lambda}Z^{\mathop{\scriptstyle\mathrm{td}}}(w,\ell) (8.18)
    \displaystyle\to (1w)2=0Ztd(w,)=1w\displaystyle(1-w)^{2}\sum_{\ell=0}^{\infty}Z^{\mathop{\scriptstyle\mathrm{td}}}(w,\ell)=1-w

    where the last step is obtained by setting z=1z=1 in the expression of the generating function (4.5). The right side of (8.18), 1w1-w, is precisely the limiting value of 1/NL\langle 1/N_{L}\rangle, see (8.9). This result is therefore the analogue of (3.14) with the correspondence given in (8.16).

In view of (8.13) and (8.18) we conclude that the weights of the downhill and uphill regions add to one, in line with the fact that the contribution of the dip region is subdominant, as shown above.

8.3 Regimes for the distribution of the longest interval

Refer to caption
Figure 9: tdrp: the main graph depicts the exact distributions of the size of a generic interval, π(k|L)\pi(k|L) (in green), and of the longest one, p(k|L)p(k|L) (in blue), in the condensed phase, for L=60L=60 and w=0.6w=0.6 computed from (4.26) and (4.43), with Example 1 (see (5.1)). The two curves join at wf(L)/Ztd(w,L)(1w)2wf(L)/Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L)\approx(1-w)^{2} on the yy-axis, for k=60k=60. The inset depicts π(k|L)(1+NLk)\pi(k|L)(1+\langle N_{L-k}\rangle) (in black) and p(k|L)p(k|L) (in blue). For this value of ww, the asymptotic average number NL4\langle N_{L}\rangle\to 4.

As can be seen on figure 9, there are two main regions for the distribution of the maximum, p(k|L)p(k|L). For kL/2k\leq L/2, the contribution of p(k|L)p(k|L) to the total weight is vanishingly small. The argument is the same as in §3.3. Hence we restrict the rest of the discussion to the region (L/2<kL)(L/2<k\leq L), where p(k|L)p(k|L) has the simpler expression q(k|L)q(k|L) given by (4.47). Using (8.1), its asymptotic estimate is

q(k|L)L(1w)2f(k)(ZtdZtd)(w,Lk)f(L),q(k|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}(1-w)^{2}f(k)\frac{(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L-k)}{f(L)}, (8.19)

or, equivalently, for =Lk(0,L/21)\ell=L-k\in(0,L/2-1),

q(L|L)L(1w)2f(L)(ZtdZtd)(w,)f(L).\displaystyle q(L-\ell|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}(1-w)^{2}f(L-\ell)\frac{(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell)}{f(L)}. (8.20)

Let us note that the ratio

r()=q(L|L)π(L|L)=(ZtdZtd)(w,)Ztd(w,)=1+Nr(\ell)=\frac{q(L-\ell|L)}{\pi(L-\ell|L)}=\frac{(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell)}{Z^{\mathop{\scriptstyle\mathrm{td}}}(w,\ell)}=1+\langle N_{\ell}\rangle (8.21)

is an increasing function of \ell, with first values

r(0)=1,r(1)=2,r(2)=3wf(1)2+2f(2)wf(1)2+f(2),r(0)=1,\quad r(1)=2,\quad r(2)=\frac{3wf(1)^{2}+2f(2)}{wf(1)^{2}+f(2)}, (8.22)

and so on, reaching the limit, for large \ell,

r()21w.r(\ell)\to\frac{2}{1-w}. (8.23)

The inset in figure 9 depicts the prediction q(k|L)=π(k|L)(1+NLk)q(k|L)=\pi(k|L)(1+\langle N_{L-k}\rangle) (see (4.49)) which coincides perfectly with p(k|L)p(k|L) in the second half k>L/2k>L/2. This graph is very similar to the inset in figure 5.

We now discuss the behaviour of q(L|L)q(L-\ell|L) in the two regions of interest.

  1. 1.

    For \ell finite, we have, see (8.20),

    q(L|L)(1w)2(ZtdZtd)(w,).q(L-\ell|L)\approx(1-w)^{2}(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell). (8.24)

    In particular for =0\ell=0,

    p(L|L)=q(L|L)=π(L|L)=Prob(NL=1)(1w)2.p(L|L)=q(L|L)=\pi(L|L)=\mathop{\rm Prob}\nolimits(N_{L}=1)\approx(1-w)^{2}. (8.25)

    If 1L1\ll\ell\ll L, (8.24) simplifies to

    q(L|L)2w1wf().q(L-\ell|L)\approx\frac{2w}{1-w}f(\ell). (8.26)
  2. 2.

    If \ell et LL-\ell are simultaneously large, with =λL\ell={\lambda}L, we have

    q(L|L)2w1wf()f(L)f(L)2w1w1[λ(1λ)]1+θcL1+θ,q(L-\ell|L)\approx\frac{2w}{1-w}f(\ell)\frac{f(L-\ell)}{f(L)}\approx\frac{2w}{1-w}\frac{1}{[{\lambda}(1-{\lambda})]^{1+{\theta}}}\frac{c}{L^{1+{\theta}}}, (8.27)

    which is proportional to (8.14), with ratio 2/(1w)2/(1-w). Note the similarity of (8.27) with (3.17) with the correspondence

    n1NL1=2w1w.n-1\hookrightarrow\langle N_{L}\rangle-1=\frac{2w}{1-w}. (8.28)

    The weight under the peak of the condensate tends to unity,

    =0Λq(L|L)\displaystyle\sum_{\ell=0}^{\Lambda}q(L-\ell|L) \displaystyle\approx (1w)2=0Λ(ZtdZtd)(w,)\displaystyle(1-w)^{2}\sum_{\ell=0}^{\Lambda}(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell) (8.29)
    \displaystyle\to (1w)2=0(ZtdZtd)(w,)=1,\displaystyle(1-w)^{2}\sum_{\ell=0}^{\infty}(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell)=1,

    by a computation similar to (8.18), using now the fact that the last sum is equal to 1/(1w)21/(1-w)^{2}, as can be seen by setting z=1z=1 in the expression of the generating function Z~td(w,z)2\tilde{Z}^{\mathop{\scriptstyle\mathrm{td}}}(w,z)^{2}.

We can further analyse the fluctuations of the condensate by considering the width of the peak,

LXmax==0L1q(L|L)=0L/2q(L|L).L-\langle X_{\rm max}\rangle=\sum_{\ell=0}^{L-1}\ell\,q(L-\ell|L)\approx\sum_{\ell=0}^{L/2}\ell\,q(L-\ell|L). (8.30)

The dominant contribution to this sum depends on whether θ{\theta} is smaller or larger than one.

\bullet If θ<1{\theta}<1, the dominant contribution comes from (8.27), i.e., for \ell comparable to LL,

LXmax\displaystyle L-\langle X_{\rm max}\rangle \displaystyle\approx 2wc1wL1θ01/2dλλ[(λ(1λ)]1+θ\displaystyle\frac{2wc}{1-w}L^{1-{\theta}}\int_{0}^{1/2}{\rm d}{\lambda}\frac{{\lambda}}{[({\lambda}(1-{\lambda})]^{1+{\theta}}} (8.31)
\displaystyle\approx 2wc1wB12(1θ,θ)L1θ,\displaystyle\frac{2wc\,}{1-w}\mathrm{B}_{\frac{1}{2}}\Big{(}1-{\theta},-{\theta}\Big{)}L^{1-{\theta}},

where B()\mathrm{B}(\cdot) is the incomplete beta function, which, for example, is equal to 22 for θ=1/2{\theta}=1/2.

\bullet If θ>1{\theta}>1, the main contribution comes from (8.24),

LXmax\displaystyle L-\langle X_{\rm max}\rangle \displaystyle\approx (1w)2=0Λ(ZtdZtd)(w,)\displaystyle(1-w)^{2}\sum_{\ell=0}^{\Lambda}\ell(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell) (8.32)
\displaystyle\to (1w)2=0(ZtdZtd)(w,)2w1wX,\displaystyle(1-w)^{2}\sum_{\ell=0}^{\infty}\ell(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,\ell)\to\frac{2w}{1-w}\langle X\rangle,

using (8.4). This last result (8.32) has a simple interpretation. It says that the correction LXmaxL-\langle X_{\rm max}\rangle is made of NL1=2w/(1w)\langle N_{L}\rangle-1=2w/(1-w) intervals, of mean length X\langle X\rangle. It is therefore the perfect parallel of the result (3.8). Likewise, (8.31) is the perfect parallel of (3.6). In the present case the expression in the right side of this equation is proportional to NL1\langle N_{L}\rangle-1 times the critical mean interval X|L\langle X|L\rangle, given in (7.20).

8.4 Discussion

In view of the above analysis the following picture emerges. The ‘contrast’ between the dip and the condensate region increases with LL, i.e., the dip centred around L/2L/2 becomes deeper and deeper as L1θL^{-1-{\theta}} (see (8.14)) relatively to the height of the peak, which is of order one. An estimate of the contribution of the condensate to the total weight can thus be operationally obtained by summing π(L|L)\pi(L-\ell|L) for \ell in the layer (0,Λ)(0,\Lambda). This sum is asymptotically equal to 1w1-w, according to (8.18), which turns out to be also the asymptotic estimate of 1/NL\langle 1/N_{L}\rangle. The interpretation of this result is clear. When X1X_{1} is larger than L/2L/2, this interval is necessarily the longest one, i.e., X1=XmaxX_{1}=X_{\rm max}. Furthermore, since this interval is chosen amongst NLN_{L} intervals, we expect that, in average,

=0Λπ(L|L)1NL1w=0Λp(L|L)1.\sum_{\ell=0}^{\Lambda}\pi(L-\ell|L)\approx\mathrel{\mathop{\underbrace{\left\langle\frac{1}{N_{L}}\right\rangle}}\limits_{1-w}}\ \mathrel{\mathop{\underbrace{\sum_{\ell=0}^{\Lambda}p(L-\ell|L)}}\limits_{\approx 1}}. (8.33)

But the sum on the right side, namely the weight of XmaxX_{\rm max} in the same layer is asymptotically equal to one. This simple heuristic reasoning therefore recovers (8.18).

In the condensed phase (w<1,Lw<1,L\to\infty) the number of intervals is finite and fluctuates around its mean, which is a superuniversal constant independent of θ{\theta}. This situation is akin to the case of random allocation models and zrp, when LL\to\infty and nn is kept fixed. Note that, for the latter, results were independent of the value of θ{\theta}, too. In both situations condensation is total, the condensed fraction is asymptotically equal to unity.

Table 1 summarises the results found in sections 6, 7 and 8, which demonstrate a large degree of universality.

Table 1: Dominant asymptotic behaviours at large LL for tied-down renewal processes with power-law distribution (1.1) for f(k)f(k), in the different phases.
disordered critical θ<1{\theta}<1 critical θ>1{\theta}>1 condensed
NL\langle N_{L}\rangle LX|L\frac{L}{\langle X|L\rangle} LθL^{\theta} LX\frac{L}{\langle X\rangle} 1+w1w\frac{1+w}{1-w}
X|L\langle X|L\rangle constant\mathrm{constant} L1θL^{1-{\theta}} X\langle X\rangle (1w)L(1-w)L
Xmax\langle X_{\rm max}\rangle lnL\ln L LL L1/θL^{1/{\theta}} LL
Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) eL/ξ{\rm e}^{L/\xi} Lθ1L^{{\theta}-1} 1X\frac{1}{\langle X\rangle} L1θL^{-1-{\theta}}

9 General statements on free renewal processes

We now turn to the case of free renewal processes. The random number NLN_{L} of intervals up to LL is defined through the condition (1.4), SNL<L<SNL+1S_{N_{L}}<L<S_{N_{L}+1}. The size of the current (unfinished) interval—named the backward recurrence time in renewal theory—is denoted by BL=LSNLB_{L}=L-S_{N_{L}}, see figure 2. As will appear shortly, free renewal processes are more complicated to analyse than tdrp, essentially because now there are two kinds of intervals to consider, the intervals XiX_{i} on one hand, and the last unfinished interval BLB_{L}, on the other hand.

In the present section, f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k) is any arbitrary distribution of the positive random variable XX. Later on, in sections 10 and 11, f(k)f(k) will obey the form (1.1).

9.1 Joint distribution

As for tdrp a weight ww is attached to each interval. The joint probability of the configuration {X1=k1,,XNL=kn,BL=b,NL=n}\{X_{1}=k_{1},\dots,X_{N_{L}}=k_{n},B_{L}=b,N_{L}=n\}, reads

p(k1,,kn,b,n|L)\displaystyle p(k_{1},\ldots,k_{n},b,n|L) =\displaystyle= Prob({Xi=ki},BL=b,NL=n)\displaystyle\mathop{\rm Prob}\nolimits(\{X_{i}=k_{i}\},B_{L}=b,N_{L}=n) (9.1)
=\displaystyle= 1Zf(w,L)wnf(k1)f(kn)g(b)δ(i=1nki+b,L),\displaystyle\frac{1}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}w^{n}f(k_{1})\ldots f(k_{n})\,g(b)\,\delta\big{(}\sum_{i=1}^{n}k_{i}+b,L\big{)},

where g(b)g(b) is the tail probability (or complementary distribution function) defined in (7.4),

g(b)=Prob(X>b)=k>bf(k)=f(b+1)+f(b+2)+g(b)=\mathop{\rm Prob}\nolimits(X>b)=\sum_{k>b}f(k)=f(b+1)+f(b+2)+\cdots (9.2)

As mentioned in section 4, since the summands XiX_{i} have the interpretation of the sizes of the intervals, we take f(0)=0f(0)=0. For n=0n=0,

p({},b,0|L)|num=g(b)δ(b,L),p(\{\},b,0|L)_{|{\rm num}}=g(b)\delta(b,L), (9.3)

corresponding to the event of no renewal occurring between 0 and LL, i.e., BL=LB_{L}=L, and where {}\{\} means empty. The generating function of g(b)g(b) is

g~(z)=b0zbg(b)=11zb1zbk=1bf(k)=1f~(z)1z,\tilde{g}(z)=\sum_{b\geq 0}z^{b}g(b)=\frac{1}{1-z}-\sum_{b\geq 1}z^{b}\sum_{k=1}^{b}f(k)=\frac{1-\tilde{f}(z)}{1-z}, (9.4)

with g~(0)=g(0)=1\tilde{g}(0)=g(0)=1.

The denominator of (9.1) is the free partition function, obtained by summing on nn, on the kik_{i} and on bb,

Zf(w,L)\displaystyle Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) =\displaystyle= n0b0{ki}p(k1,,kn,b,n|L)\displaystyle\sum_{n\geq 0}\sum_{b\geq 0}\sum_{\{k_{i}\}}\,p(k_{1},\ldots,k_{n},b,n|L) (9.5)
=\displaystyle= n0wnb0{ki}f(k1)f(kn)g(b)δ(i=1nki+b,L)\displaystyle\sum_{n\geq 0}w^{n}\sum_{b\geq 0}\sum_{\{k_{i}\}}f(k_{1})\cdots f(k_{n})g(b)\delta\big{(}\sum_{i=1}^{n}k_{i}+b,L\big{)}
=\displaystyle= b0g(b)[δ(b,L)+k1wf(k1)δ(k1+b,L)\displaystyle\sum_{b\geq 0}g(b)\Big{[}\delta(b,L)+\sum_{k_{1}}wf(k_{1})\delta(k_{1}+b,L)
+\displaystyle+ k1,k2w2f(k1)f(k2)δ(k1+k2+b,L)+]\displaystyle\sum_{k_{1},k_{2}}w^{2}f(k_{1})f(k_{2})\delta(k_{1}+k_{2}+b,L)+\dots\Big{]}
=\displaystyle= g(L)n=0+wgfn=1+w2gffn=2+=n0wn(g(f)n)(L).\displaystyle\,\mathrel{\mathop{\underbrace{g(L)}}\limits_{n=0}}+\mathrel{\mathop{\underbrace{w\,g\star f}}\limits_{n=1}}+\mathrel{\mathop{\underbrace{w^{2}\,g\star f\star f}}\limits_{n=2}}+\dots=\sum_{n\geq 0}w^{n}\big{(}g\star(f\star)^{n}\big{)}(L).

For instance,

Zf(w,0)=1,Zf(w,1)=g(1)+wf(1),\displaystyle Z^{\mathop{\scriptstyle\mathrm{f}}}(w,0)=1,\qquad Z^{\mathop{\scriptstyle\mathrm{f}}}(w,1)=g(1)+wf(1),\qquad (9.6)
Zf(w,2)=g(2)+wf(2)+wf(1)g(1)+w2f(1)2,\displaystyle Z^{\mathop{\scriptstyle\mathrm{f}}}(w,2)=g(2)+wf(2)+wf(1)g(1)+w^{2}f(1)^{2}, (9.7)

and so on. From (9.5) we have

Z~f(w,z)=L0zLZf(w,L)=n0(wf~(z))ng~(z)=g~(z)1wf~(z).\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)=\sum_{L\geq 0}z^{L}Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)=\sum_{n\geq 0}\big{(}w\tilde{f}(z)\big{)}^{n}\tilde{g}(z)=\frac{\tilde{g}(z)}{1-w\tilde{f}(z)}. (9.8)

Let us note that

Zf(w,L)=n0wnb0Prob(Sn=Lb)g(b)=(gZtd)(w,L),Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)=\sum_{n\geq 0}w^{n}\sum_{b\geq 0}\mathop{\rm Prob}\nolimits(S_{n}=L-b)g(b)=(g\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L), (9.9)

where Ztd(w,L)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L) is the partition function (4.3) for tdrp.

The case of usual (unweighted) free renewal processes is recovered by setting w=1w=1 in these expressions. This yields Zf(1,L)=1Z^{\mathop{\scriptstyle\mathrm{f}}}(1,L)=1, as can be seen from (9.8), and the joint probability distribution (9.1) simplifies accordingly.

9.2 Distribution of the number of intervals

As for tdrp we denote this distribution as555Whenever no ambiguity arises, we use the same notations for the observables of the tied-down and free renewal processes. Otherwise, when necessary, we add a superscript, as e.g., for ZtdZ^{\mathop{\scriptstyle\mathrm{td}}}, ZfZ^{\mathop{\scriptstyle\mathrm{f}}} or in (9.18).

pn(L)=Prob(NL=n).\mathrm{p}_{n}(L)=\mathop{\rm Prob}\nolimits(N_{L}=n). (9.10)

We read on the successive terms of Zf(w,L)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) that

p0(L)=g(L)Zf(w,L)=Prob(BL=L),p1(L)=w(gf)(L)Zf(w,L),\mathrm{p}_{0}(L)=\frac{g(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}=\mathop{\rm Prob}\nolimits(B_{L}=L),\quad\mathrm{p}_{1}(L)=\frac{w(g\star f)(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}, (9.11)

and so on. More generally,

pn(L)=wn(g(f)n)(L)Zf(w,L)=wn(gZn)(L)(gZtd)(w,L).\mathrm{p}_{n}(L)=\frac{w^{n}\big{(}g\star(f\star)^{n}\big{)}(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}=\frac{w^{n}(g\star Z_{n})(L)}{(g\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)}. (9.12)

Summing (9.1) on bb and on the kik_{i}, and taking the generating function with respect to LL yields

L0zLpn(L)|num=(wf~(z))ng~(z),\sum_{L\geq 0}z^{L}\mathrm{p}_{n}(L)_{|{\rm num}}=\Big{(}w\tilde{f}(z)\Big{)}^{n}\tilde{g}(z), (9.13)

to be compared to (9.8). Therefore

L0zLNL|num=wf~(z)g~(z)(1wf~(z))2=wddwZ~f(w,z),\sum_{L\geq 0}z^{L}\langle N_{L}\rangle_{|{\rm num}}=\frac{w\tilde{f}(z)\tilde{g}(z)}{(1-w\tilde{f}(z))^{2}}=w\frac{{\rm d}}{{\rm d}w}\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z), (9.14)

and

NL=wdlnZf(w,L)dw,\langle N_{L}\rangle=w\frac{{\rm d}\ln Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}{{\rm d}w}, (9.15)

as for tdrp, see (4.15). More generally,

L0zLvNL|num=n0vn(wf~(z))ng~(z)=Z~f(vw,z),\sum_{L\geq 0}z^{L}\left\langle v^{N_{L}}\right\rangle_{|{\rm num}}=\sum_{n\geq 0}v^{n}\Big{(}w\tilde{f}(z)\Big{)}^{n}\tilde{g}(z)=\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(vw,z), (9.16)

so we obtain, as for tdrp, see (4.20),

vNL=n0vnpn(L)=Zf(vw,L)Zf(w,L).\left\langle v^{N_{L}}\right\rangle=\sum_{n\geq 0}v^{n}\mathrm{p}_{n}(L)=\frac{Z^{\mathop{\scriptstyle\mathrm{f}}}(vw,L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}. (9.17)

Finally, comparing (9.14) to (4.14), we note the relationship between the free and tied-down cases,

NLf|num=k=0Lg(k)NLktd|num,\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L}\rangle_{|{\rm num}}=\sum_{k=0}^{L}g(k)\langle N^{\mathop{\scriptstyle\mathrm{td}}}_{L-k}\rangle_{|{\rm num}}, (9.18)

hence using (4.17), we have

NLf|num=(gZtdZtd)(w,L)Zf(w,L),\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L}\rangle_{|{\rm num}}=(g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)-Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L), (9.19)

or, equivalently,

(gZtdZtd)(w,L)=Zf(w,L)(1+NL).(g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)=Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)(1+\langle N_{L}\rangle). (9.20)

9.3 Distribution of SNLS_{N_{L}}

We recall that this quantity is the sum of the NLN_{L} intervals before LL, see (1.3) and figure 2. By definition,

Prob(SNL=j)=δ(SNL,j)=n0b0{ki}δ(SNL,j)p(k1,,kn,b,n|L).\mathop{\rm Prob}\nolimits(S_{N_{L}}=j)=\langle\delta(S_{N_{L}},j)\rangle=\sum_{n\geq 0}\sum_{b\geq 0}\sum_{\{k_{i}\}}\delta(S_{N_{L}},j)p(k_{1},\ldots,k_{n},b,n|L). (9.21)

Thus, using (9.1), we have

L0zLj=0LxjProb(SNL=j)|num=g~(z)1wf~(xz),\sum_{L\geq 0}z^{L}\sum_{j=0}^{L}x^{j}\mathop{\rm Prob}\nolimits(S_{N_{L}}=j)_{|{\rm num}}=\frac{\tilde{g}(z)}{1-w\tilde{f}(xz)}, (9.22)

which generalises the expression for this quantity when w=1w=1 gl2001 . By derivation with respect to xx then setting x=1x=1, leads to

L0zLSNL|num=wzf~(z)g~(z)(1wf~(z))2,\sum_{L\geq 0}z^{L}\langle S_{N_{L}}\rangle_{|{\rm num}}=\frac{wz\tilde{f}^{\prime}(z)\tilde{g}(z)}{(1-w\tilde{f}(z))^{2}}, (9.23)

whose summation with (9.31) below leads to the equality

L0zL(SNL|num+BL|num)=zdZ~f(w,z)dz,\sum_{L\geq 0}z^{L}\big{(}\langle S_{N_{L}}\rangle_{|{\rm num}}+\langle B_{L}\rangle_{|{\rm num}}\big{)}=z\frac{{\rm d}\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)}{{\rm d}z}, (9.24)

which expresses the sum rule

SNL+BL=L.\langle S_{N_{L}}\rangle+\langle B_{L}\rangle=L. (9.25)

The asymptotic behaviours of these quantities for w=1w=1 are simple gl2001 . If θ<1{\theta}<1, then SNLθL\langle S_{N_{L}}\rangle\approx{\theta}L, BL(1θ)L\langle B_{L}\rangle\approx(1-{\theta})L. If 1<θ<21<{\theta}<2, then SNLLcL2θ/[θ(2θ)X]\langle S_{N_{L}}\rangle\approx L-cL^{2-{\theta}}/[{\theta}(2-{\theta})\langle X\rangle], and BL\langle B_{L}\rangle follows by difference.

9.4 Distribution of BLB_{L}

The distribution of BLB_{L} is obtained by summing (9.1) on the kik_{i} and on nn

Prob(BL=b)=1Zf(w,L)g(b)n0wn{ki1}f(k1)f(kn)δ(i=1nki+b,L).\mathop{\rm Prob}\nolimits(B_{L}=b)=\frac{1}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}g(b)\sum_{n\geq 0}w^{n}\sum_{\{k_{i}\geq 1\}}f(k_{1})\cdots f(k_{n})\delta\Big{(}\sum_{i=1}^{n}k_{i}+b,L\Big{)}. (9.26)

This entails that

Prob(BL=b)|num=g(b)n0wnProb(Sn=Lb)=g(b)Ztd(w,Lb),\mathop{\rm Prob}\nolimits(B_{L}=b)_{|{\rm num}}=g(b)\sum_{n\geq 0}w^{n}\mathop{\rm Prob}\nolimits(S_{n}=L-b)=g(b)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-b), (9.27)

hence, using (9.9)

Prob(BL=b)=g(b)Ztd(w,Lb)(gZtd)(w,L).\mathop{\rm Prob}\nolimits(B_{L}=b)=\frac{g(b)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-b)}{(g\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L)}. (9.28)

The generating function with respect to LL of (9.27) reads

L0zLProb(BL=b)|num=zbg(b)1wf~(z),\sum_{L\geq 0}z^{L}\mathop{\rm Prob}\nolimits(B_{L}=b)_{|{\rm num}}=\frac{z^{b}g(b)}{1-w\tilde{f}(z)}, (9.29)

which summed upon bb gives Z~f(w,z)\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z) back. Taking now the generating function of (9.29) with respect to bb yields

L0zLb0ybProb(BL=b)|num=g~(yz)1wf~(z).\sum_{L\geq 0}z^{L}\sum_{b\geq 0}y^{b}\mathop{\rm Prob}\nolimits(B_{L}=b)_{|{\rm num}}=\frac{\tilde{g}(yz)}{1-w\tilde{f}(z)}. (9.30)

The mean BL\langle B_{L}\rangle ensues by taking the derivative of the right side of this equation with respect to yy and setting yy to one,

L0zLBL|num=zg~(z)1wf~(z).\sum_{L\geq 0}z^{L}\langle B_{L}\rangle_{|{\rm num}}=\frac{z\tilde{g}^{\prime}(z)}{1-w\tilde{f}(z)}. (9.31)

9.5 Single interval distribution

As for tdrp, the single interval distribution,

π(k|L)=δ(X1,k),\pi(k|L)=\langle\delta(X_{1},k)\rangle, (9.32)

is obtained by summing p(k1,,kn,b,n|L)p(k_{1},\dots,k_{n},b,n|L) on k1,,knk_{1},\dots,k_{n}, bb and n1n\geq 1

π(k|L)|num\displaystyle\pi(k|L)_{|{\rm num}} =\displaystyle= b0g(b)[k1δ(k1,k)wf(k1)δ(k1+b,L)\displaystyle\sum_{b\geq 0}g(b)\Big{[}\sum_{k_{1}}\delta(k_{1},k)wf(k_{1})\delta(k_{1}+b,L) (9.33)
+\displaystyle+ k1,k2δ(k1,k)w2f(k1)f(k2)δ(k1+k2+b,L)+]\displaystyle\sum_{k_{1},k_{2}}\delta(k_{1},k)w^{2}f(k_{1})f(k_{2})\delta(k_{1}+k_{2}+b,L)+\cdots\Big{]}
=\displaystyle= b0g(b)[wf(k)δ(k+b,L)+k2w2f(k)f(k2)δ(k+k2+b,L)+]\displaystyle\sum_{b\geq 0}g(b)\Big{[}wf(k)\delta(k+b,L)+\sum_{k_{2}}w^{2}f(k)f(k_{2})\delta(k+k_{2}+b,L)+\dots\Big{]}
=\displaystyle= wf(k)Zf(w,Lk).\displaystyle wf(k)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k).

So

π(k|L)=wf(k)Zf(w,Lk)Zf(w,L).\pi(k|L)=\frac{wf(k)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}. (9.34)

The generating function of the numerator is therefore

L0zLπ(k|L)|num=wzkf(k)Z~f(w,z)=wzkf(k)g~(z)1wf~(z),\displaystyle\sum_{L\geq 0}z^{L}\pi(k|L)_{|{\rm num}}=wz^{k}f(k)\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)=\frac{wz^{k}f(k)\tilde{g}(z)}{1-w\tilde{f}(z)}, (9.35)

to be compared to (4.32). Though (9.34) is formally identical to (4.26) the normalisations of these two distributions are different. Indeed,

L0zLk=1Lπ(k|L)|num=wf~(z)Z~f(w,z)=wf~(z)g~(z)1wf~(z)=Z~f(w,z)g~(z),\displaystyle\sum_{L\geq 0}z^{L}\sum_{k=1}^{L}\pi(k|L)_{|{\rm num}}=w\tilde{f}(z)\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)=\frac{w\tilde{f}(z)\tilde{g}(z)}{1-w\tilde{f}(z)}=\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)-\tilde{g}(z), (9.36)

which means that

1k=1Lπ(k|L)=g(L)Zf(w,L)=p0(L)=Prob(BL=L).1-\sum_{k=1}^{L}\pi(k|L)=\frac{g(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}=\mathrm{p}_{0}(L)=\mathop{\rm Prob}\nolimits(B_{L}=L). (9.37)

In other words

k=1LProb(X=k|L)+Prob(BL=L)=1.\sum_{k=1}^{L}\mathop{\rm Prob}\nolimits(X=k|L)+\mathop{\rm Prob}\nolimits(B_{L}=L)=1. (9.38)

The distribution π(k|L)\pi(k|L) is thus defective. The recursion relation for Zf(w,L)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) follows from (9.33) and (9.37)

Zf(w,L)=g(L)+k=1Lwf(k)Zf(w,Lk).Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)=g(L)+\sum_{k=1}^{L}wf(k)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k). (9.39)

At the end point, k=Lk=L, we have

π(L|L)=wf(L)Zf(w,L),\pi(L|L)=\frac{wf(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}, (9.40)

which corresponds to the event {X1=L,BL=0,NL=1}\{X_{1}=L,B_{L}=0,N_{L}=1\}, and which is formally the same as (4.25).

Both sides of (9.39) are equal to unity if w=1w=1. Note that the computation of π(k|L)\pi(k|L) for w=1w=1 is given in gl2001 , yielding π(k|L)=f(k)\pi(k|L)=f(k) with kLk\leq L, which shows that this distribution is already defective for w=1w=1, with k=1Lπ(k|L)=1g(L)\sum_{k=1}^{L}\pi(k|L)=1-g(L).

9.6 Mean interval X|L\langle X|L\rangle

We proceed as in §4.4. The mean interval is, by definition,

X|L=k1kπ(k|L).\langle X|L\rangle=\sum_{k\geq 1}k\pi(k|L). (9.41)

Multiplying (9.35) by kk and summing upon kk yields

L1zLX|L|num=wzf~(z)g~(z)1wf~(z),\sum_{L\geq 1}z^{L}\langle X|L\rangle_{|{\rm num}}=\frac{wz\tilde{f}^{\prime}(z)\tilde{g}(z)}{1-w\tilde{f}(z)}, (9.42)

to be compared to (4.32) for tdrp.

9.7 The longest interval

In the present case the longest interval is defined as

Xmax=max(X1,X2,,BL).X_{\rm max}={\rm max}(X_{1},X_{2},\dots,B_{L}). (9.43)

Its distribution function is

F(k|L)=Prob(Xmaxk|L)=n0b=0kk1=1kkn=1kp({ki},b,n|L)=F(k|L)|numZf(w,L),F(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}\leq k|L)=\sum_{n\geq 0}\sum_{b=0}^{k}\sum_{k_{1}=1}^{k}\dots\sum_{k_{n}=1}^{k}p(\{k_{i}\},b,n|L)=\frac{F(k|L)_{|{\rm num}}}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}, (9.44)

with initial value

F(k|0)|num=1.F(k|0)_{|{\rm num}}=1. (9.45)

As for tdrp, F(L|L)|num=Zf(w,L)F(L|L)_{|{\rm num}}=Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L), hence F(L|L)=1F(L|L)=1. The generating function of the numerator is

L0zLF(k|L)|num\displaystyle\sum_{L\geq 0}z^{L}F(k|L)_{|{\rm num}} =\displaystyle= g~(z,k)(1+n1i=1n(ki=1kwf(ki)zki))\displaystyle\tilde{g}(z,k)\Bigg{(}1+\sum_{n\geq 1}\prod_{i=1}^{n}\Big{(}\sum_{k_{i}=1}^{k}wf(k_{i})z^{k_{i}}\Big{)}\Bigg{)} (9.46)
=\displaystyle= g~(z,k)(1+n1(wf~(z,k))n)=g~(z,k)1wf~(z,k),\displaystyle\tilde{g}(z,k)\Bigg{(}1+\sum_{n\geq 1}\Big{(}w\tilde{f}(z,k)\Big{)}^{n}\Bigg{)}=\frac{\tilde{g}(z,k)}{1-w\tilde{f}(z,k)},

where

f~(z,k)=j=1kzjf(j),g~(z,k)=b=0kzbg(b),\tilde{f}(z,k)=\sum_{j=1}^{k}z^{j}f(j),\qquad\tilde{g}(z,k)=\sum_{b=0}^{k}z^{b}g(b), (9.47)

are related by

1f~(z,k)=zk+1g(k)+(1z)g~(z,k).1-\tilde{f}(z,k)=z^{k+1}g(k)+(1-z)\tilde{g}(z,k). (9.48)

Note that

Ff(k|L)|num\displaystyle F^{\mathop{\scriptstyle\mathrm{f}}}(k|L)_{|{\rm num}} =\displaystyle= b=0kg(b)Ftd(k|Lb)|num\displaystyle\sum_{b=0}^{k}g(b)F^{\mathop{\scriptstyle\mathrm{td}}}(k|L-b)_{|{\rm num}} (9.49)
=\displaystyle= b=0kg(b)n=0LwnFn(k|Lb)|num,\displaystyle\sum_{b=0}^{k}g(b)\sum_{n=0}^{L}w^{n}F_{n}(k|L-b)_{|{\rm num}},

where we used (4.38) in the last step.

The distribution of XmaxX_{\rm max} is given by the difference

p(k|L)=Prob(Xmax=k)=F(k|L)F(k1|L),p(k|L)=\mathop{\rm Prob}\nolimits(X_{{\rm max}}=k)=F(k|L)-F(k-1|L), (9.50)

where F(0|L)=δ(L,0)F(0|L)=\delta(L,0), with generating function

L0zLp(k|L)|num\displaystyle\sum_{L\geq 0}z^{L}p(k|L)_{|{\rm num}} =\displaystyle= g~(z,k)(11wf~(z,k)11wf~(z,k1))\displaystyle\tilde{g}(z,k)\Big{(}\frac{1}{1-w\tilde{f}(z,k)}-\frac{1}{1-w\tilde{f}(z,k-1)}\Big{)} (9.51)
=\displaystyle= wzkf(k)g~(z,k)[1wf~(z,k)][1wf~(z,k1)].\displaystyle\frac{wz^{k}f(k)\tilde{g}(z,k)}{[1-w\tilde{f}(z,k)][1-w\tilde{f}(z,k-1)]}.

At the end point, k=Lk=L, we have

p(L|L)=wf(L)+g(L)Zf(w,L)=π(L|L)+Prob(BL=L),p(L|L)=\frac{wf(L)+g(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}=\pi(L|L)+\mathop{\rm Prob}\nolimits(B_{L}=L), (9.52)

where the last two terms correspond respectively to the events {X1=L,BL=0,NL=1}\{X_{1}=L,B_{L}=0,N_{L}=1\}, cf (9.40) and {BL=L,NL=0}\{B_{L}=L,N_{L}=0\}, cf (9.11). The mean is given by the sum

Xmax=k=0L(1F(k|L))=Lk=1L1F(k|L),\langle X_{\rm max}\rangle=\sum_{k=0}^{L}\Big{(}1-F(k|L)\Big{)}=L-\sum_{k=1}^{L-1}F(k|L), (9.53)

which implies a relation between the generating functions

L0zLXmax|num=k0(Z~f(w,z)F~(k,z)|num)\displaystyle\sum_{L\geq 0}z^{L}{\langle X_{\rm max}\rangle}_{|{\rm num}}=\sum_{k\geq 0}\Big{(}\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)-\tilde{F}(k,z)_{|{\rm num}}\Big{)} (9.54)
=k0(g~(z)1wf~(z)g~(z,k)1wf~(z,k)),\displaystyle=\sum_{k\geq 0}\Big{(}\frac{\tilde{g}(z)}{1-w\tilde{f}(z)}-\frac{\tilde{g}(z,k)}{1-w\tilde{f}(z,k)}\Big{)}, (9.55)

where F~(k,z)|num\tilde{F}(k,z)_{|{\rm num}} is given by (9.46).

Denoting again the restriction of p(k|L)p(k|L) to k>L/2k>L/2 by q(k|L)q(k|L), we can obtain an expression of this quantity by a similar reasoning as done for (4.47) in §4.5. We thus obtain

q(k|L)\displaystyle q(k|L) =\displaystyle= Prob(BL=k)+wf(k)Zf(w,L)b=0Lkg(b)(ZtdZtd)(w,Lkb)\displaystyle\mathop{\rm Prob}\nolimits(B_{L}=k)+\frac{wf(k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}\sum_{b=0}^{L-k}g(b)(Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}})(w,L-k-b) (9.56)
=\displaystyle= Prob(BL=k)+wf(k)(gZtdZtd)(w,Lk)Zf(w,L),\displaystyle\mathop{\rm Prob}\nolimits(B_{L}=k)+\frac{wf(k)\big{(}g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}}\big{)}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)},

expressing the fact that the longest interval can be either BLB_{L} or a generic interval XiX_{i}. Equivalently, using (9.20), this reads

q(k|L)=Prob(BL=k)+wf(k)Zf(w,Lk)Zf(w,L)(1+NLk),\displaystyle q(k|L)=\mathop{\rm Prob}\nolimits(B_{L}=k)+\frac{wf(k)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}(1+\langle N_{L-k}\rangle), (9.57)

or else

q(k|L)=Prob(BL=k)+π(k|L)(1+NLk),\displaystyle q(k|L)=\mathop{\rm Prob}\nolimits(B_{L}=k)+\pi(k|L)(1+\langle N_{L-k}\rangle), (9.58)

to be compared to (4.48) and (4.49). For k=Lk=L, (9.52) is recovered.

Remarks

1. We can prove (9.56) otherwise. We start from the first line of (9.49) and take the discrete derivative with respect to kk of both sides, which yields

ΔkFf(k|L)|num=g(k)Ftd(k|Lk)|num+b=0kg(b)ΔkFtd(k|Lb)|num.\Delta_{k}F^{\mathop{\scriptstyle\mathrm{f}}}(k|L)_{|{\rm num}}=g(k)F^{\mathop{\scriptstyle\mathrm{td}}}(k|L-k)_{|{\rm num}}+\sum_{b=0}^{k}g(b)\Delta_{k}F^{\mathop{\scriptstyle\mathrm{td}}}(k|L-b)_{|{\rm num}}. (9.59)

If k>L/2k>L/2, using (4.54), we recognise the numerator of the first term of (9.56) in the first term of this equation. Likewise, if k>L/2k>L/2, it can be shown that the second term of the above equation is equal to the numerator of the second term of (9.56).

2. From (4.54) we infer that, if k>L/2k>L/2,

Ff(k|Lk)|num=Zf(w,Lk).F^{\mathop{\scriptstyle\mathrm{f}}}(k|L-k)_{|{\rm num}}=Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k). (9.60)

Lastly, consider the probability QL{Q_{L}} that the last unfinished interval is the longest one, that is

QL=Prob(BLmax(X1,,XNL))=n0b0k1=1bkn=1bp({ki},b,n|L).{Q_{L}}=\mathop{\rm Prob}\nolimits(B_{L}\geq{\rm max}(X_{1},\dots,X_{N_{L}}))=\sum_{n\geq 0}\sum_{b\geq 0}\sum_{k_{1}=1}^{b}\dots\sum_{k_{n}=1}^{b}p(\{k_{i}\},b,n|L). (9.61)

The generating function with respect to LL of its numerator reads

L0zLQL|num=b0zbg(b)1wf~(z,b),\sum_{L\geq 0}z^{L}{Q_{L}}_{|{\rm num}}=\sum_{b\geq 0}\frac{z^{b}g(b)}{1-w\tilde{f}(z,b)}, (9.62)

hence

Z~f(w,z)L0zLQL|num\displaystyle\tilde{Z}^{\mathop{\scriptstyle\mathrm{f}}}(w,z)-\sum_{L\geq 0}z^{L}{Q_{L}}_{|{\rm num}} =\displaystyle= g~(z)1wf~b0zbg(b)1wf~(z,b)\displaystyle\frac{\tilde{g}(z)}{1-w\tilde{f}}-\sum_{b\geq 0}\frac{z^{b}g(b)}{1-w\tilde{f}(z,b)} (9.63)
=\displaystyle= b0zbg(b)(11wf~11wf~(z,b)).\displaystyle\sum_{b\geq 0}z^{b}g(b)\Big{(}\frac{1}{1-w\tilde{f}}-\frac{1}{1-w\tilde{f}(z,b)}\Big{)}. (9.64)

10 Critical regime (w=1w=1) for free renewal processes

In this section and in the following one (section 11) we specialise the discussion to the case of a subexponential distribution f(k)=Prob(X=k)f(k)=\mathop{\rm Prob}\nolimits(X=k) with asymptotic power-law decay (1.1).

The critical regime is thoroughly described in gl2001 ; gms2015 and builds upon previous studies feller ; dynkin ; lamperti58 ; lamperti61 . The results are summarised in table 2, which also presents the main outcomes for the disordered regime (w>1w>1).

The initial analysis of the distribution of the longest interval is due to Lamperti lamperti61 . Let us just recover the universal asymptotic expression of q(k|L)q(k|L), for 1kL1\ll k\sim L, with r=k/Lr=k/L fixed, when θ<1{\theta}<1, given in lamperti61 ,

q(k|L)1Lsinπθπ1r1+θ(1r)1θ.q(k|L)\approx\frac{1}{L}\frac{\sin\pi{\theta}}{\pi}\frac{1}{r^{1+{\theta}}(1-r)^{1-{\theta}}}. (10.1)

This result can be simply inferred from (9.58). For the first term of (9.58), we obtain for LL large, using (7.5),

Prob(BL=k)=g(k)Ztd(1,Lk)Lsinπθπ1kθ(Lk)1θ,\mathop{\rm Prob}\nolimits(B_{L}=k)=g(k)Z^{\mathop{\scriptstyle\mathrm{td}}}(1,L-k)\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{\sin\pi{\theta}}{\pi}\frac{1}{k^{\theta}(L-k)^{1-{\theta}}}, (10.2)

i.e., an arcsine law in the variable r=k/Lr=k/L, which is a well-known result dynkin ; feller ; gl2001 . For the second term of (9.58), we need

NLkLsinπθπc(Lk)θ,\langle N_{L-k}\rangle\mathrel{\mathop{\approx}\limits_{L\to\infty}}\frac{\sin\pi{\theta}}{\pi c}(L-k)^{\theta}, (10.3)

which is obtained using (9.14). Adding the two terms of (9.58) yields (10.1).

11 Condensed phase (w<1w<1) for free renewal processes

We now focus on the case of most interest, namely the condensed phase, w<1w<1, for subexponential distributions (1.1). As in section 8, we investigate the statistics of the number of intervals, the single-interval distribution and we characterise the fluctuations of the condensate. We also address the statistics of the last interval BLB_{L}.

11.1 Asymptotic estimates at large LL

The asymptotic analysis of (9.8) yields, for large LL,

Zf(w,L)\displaystyle Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) \displaystyle\approx g(L)1w(1wcΓ(1θ)2(1w)Γ(12θ)θLθ),θ<1\displaystyle\frac{g(L)}{1-w}\Big{(}1-\frac{wc\Gamma(1-{\theta})^{2}}{(1-w)\Gamma(1-2{\theta}){\theta}L^{\theta}}\Big{)},\quad{\theta}<1 (11.1)
Zf(w,L)\displaystyle Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) \displaystyle\approx g(L)1w(1+2wθX(1w)L),θ>1.\displaystyle\frac{g(L)}{1-w}\Big{(}1+\frac{2w{\theta}\langle X\rangle}{(1-w)L}\Big{)},\quad{\theta}>1. (11.2)

where g()g(\cdot) is defined in (7.4). As a consequence, (9.17) yields

vNLL1w1vw,\left\langle v^{N_{L}}\right\rangle\mathrel{\mathop{\to}\limits_{L\to\infty}}\frac{1-w}{1-vw}, (11.3)

leading asymptotically to a geometric distribution for NLN_{L},

Prob(NL=n)pn=(1w)wn,\mathop{\rm Prob}\nolimits(N_{L}=n)\to\mathrm{p}_{n}=(1-w)w^{n}, (11.4)

independent of θ{\theta}, from which entails, in the same limit,

NLw1w.\langle N_{L}\rangle\to\frac{w}{1-w}. (11.5)

For LL large but finite, we find, using (9.15) and (11.1) or (11.2),

NLw1w(1c1wθΓ(1θ)2Γ(12θ)Lθ),θ<1.\displaystyle\langle N_{L}\rangle\approx\frac{w}{1-w}\Big{(}1-\frac{c}{1-w}\frac{{\theta}\,\Gamma(1-{\theta})^{2}}{\Gamma(1-2{\theta})L^{\theta}}\Big{)},\quad{\theta}<1.
NLw1w(1+2θX(1w)L),θ>1.\displaystyle\langle N_{L}\rangle\approx\frac{w}{1-w}\Big{(}1+\frac{2{\theta}\langle X\rangle}{(1-w)L}\Big{)},\quad{\theta}>1. (11.6)

The asymptotic estimate of the mean interval can be obtained from the analysis of (9.42),

X|L\displaystyle\langle X|L\rangle \displaystyle\approx wcΓ(1θ)2Γ(22θ)L1θ,θ<1,\displaystyle\frac{wc\Gamma(1-{\theta})^{2}}{\Gamma(2-2{\theta})}L^{1-{\theta}},\quad{\theta}<1, (11.7)
X|L\displaystyle\langle X|L\rangle \displaystyle\approx w(1+θ)X,θ>1.\displaystyle w(1+{\theta})\langle X\rangle,\quad{\theta}>1. (11.8)

These expressions can also be obtained by means of the marginal distribution π(k|L)\pi(k|L), see below. Likewise, the asymptotic estimate of the mean sum can be obtained from the analysis of (9.23),

SNL\displaystyle\langle S_{N_{L}}\rangle \displaystyle\approx wc1wΓ(1θ)2Γ(22θ)L1θ,θ<1,\displaystyle\frac{wc}{1-w}\frac{\Gamma(1-{\theta})^{2}}{\Gamma(2-2{\theta})}L^{1-{\theta}},\qquad{\theta}<1, (11.9)
SNL\displaystyle\langle S_{N_{L}}\rangle \displaystyle\approx w1w(1+θ)X,θ>1,\displaystyle\frac{w}{1-w}(1+{\theta})\langle X\rangle,\qquad{\theta}>1, (11.10)

hence

BL\displaystyle\langle B_{L}\rangle \displaystyle\approx Lwc1wΓ(1θ)2Γ(22θ)L1θ,θ<1,\displaystyle L-\frac{wc}{1-w}\frac{\Gamma(1-{\theta})^{2}}{\Gamma(2-2{\theta})}L^{1-{\theta}},\qquad{\theta}<1, (11.11)
BL\displaystyle\langle B_{L}\rangle \displaystyle\approx Lw1w(1+θ)X,θ>1.\displaystyle L-\frac{w}{1-w}(1+{\theta})\langle X\rangle,\qquad{\theta}>1. (11.12)

11.2 Regimes for the single interval distribution

We proceed as was done for tdrp (see section 8.2). Using (11.1) we have the estimate, at large LL,

π(k|L)Lw(1w)f(k)Zf(w,Lk)g(L).\pi(k|L)\mathrel{\mathop{\approx}\limits_{L\to\infty}}w(1-w)\frac{f(k)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k)}{g(L)}. (11.13)

Figure 10 depicts the distribution π(k|L)\pi(k|L) (together with the distribution Prob(BL=k)\mathop{\rm Prob}\nolimits(B_{L}=k), see §11.3 below), for L=60L=60 and w=0.8w=0.8 computed with Example 1 (see (5.1)). As can be seen on this figure, there are three distinct regions for π(k|L)\pi(k|L), that we consider in turn.

  1. 1.

    Downhill region. For kk finite, using (11.1) again, we have

    π(k|L)wf(k)g(Lk)g(L)wf(k).\pi(k|L)\approx\frac{wf(k)g(L-k)}{g(L)}\approx wf(k). (11.14)
  2. 2.

    Dip region. When kk and LkL-k are simultaneously large, setting k=λLk={\lambda}L in (11.13) (0<λ<10<{\lambda}<1) yields the estimate, at large LL,

    π(k|L)wf(k)g(Lk)g(L)1λ1+θ(1λ)θwcL1+θ1λ1+θ(1λ)θwf(L),\pi(k|L)\approx\frac{wf(k)g(L-k)}{g(L)}\approx\frac{1}{{\lambda}^{1+{\theta}}(1-{\lambda})^{\theta}}\frac{wc}{L^{1+{\theta}}}\approx\frac{1}{{\lambda}^{1+{\theta}}(1-{\lambda})^{\theta}}wf(L), (11.15)

    with a dip centred around kmin=L(1+θ)/(1+2θ)k_{\min}=L(1+{\theta})/(1+2{\theta}).

  3. 3.

    In the region corresponding to LkL-k finite, (11.13) simplifies into

    π(k|L)w(1w)θLZf(w,Lk).\pi(k|L)\approx\frac{w(1-w){\theta}}{L}Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L-k). (11.16)

    In particular, for k=Lk=L, π(L|L)w(1w)θ/L\pi(L|L)\approx w(1-w){\theta}/L, cf (9.40).

The weight of the downhill region is found to be asymptotically equal to ww, using the same reasoning as in §8.2. The complement is borne by g(L)/Zf(w,L)1wg(L)/Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)\approx 1-w, see (9.37) and (11.21). The two other regions therefore do not contribute to the total weight, asymptotically.

In order to complete the picture we now investigate the distribution of BLB_{L}, the last unfinished interval.

11.3 Regimes for the distribution of BLB_{L}

According to (9.27), and in view of (11.1), for LL large we have

Prob(BL=b)(1w)g(b)Ztd(w,Lb)g(L).\mathop{\rm Prob}\nolimits(B_{L}=b)\approx(1-w)\frac{g(b)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-b)}{g(L)}. (11.17)

Let us discuss the different regimes of this expression according to the magnitude of bb.

  1. 1.

    If bb is finite, the asymptotic estimate of Ztd(w,Lb)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-b) is given by (8.1), hence (11.17) becomes

    Prob(BL=b)w1wg(b)f(Lb)g(L).\mathop{\rm Prob}\nolimits(B_{L}=b)\approx\frac{w}{1-w}\frac{g(b)f(L-b)}{g(L)}. (11.18)
  2. 2.

    If bLb\sim L, the same estimate (11.18) still holds, then setting Lb=λLL-b={\lambda}L, we get

    Prob(BL=b)w1wg(b)f(Lb)g(L)wc1wL1θλ1+θ(1λ)θ,\mathop{\rm Prob}\nolimits(B_{L}=b)\approx\frac{w}{1-w}\frac{g(b)f(L-b)}{g(L)}\approx\frac{wc}{1-w}\frac{L^{-1-{\theta}}}{{\lambda}^{1+{\theta}}(1-{\lambda})^{{\theta}}}, (11.19)

    which has its minimum at kmin=Lθ/(1+2θ)k_{\min}=L\,{\theta}/(1+2{\theta}).

  3. 3.

    If LbL-b is finite, (11.17) becomes

    Prob(BL=b)(1w)Ztd(w,Lb),\mathop{\rm Prob}\nolimits(B_{L}=b)\approx(1-w)Z^{\mathop{\scriptstyle\mathrm{td}}}(w,L-b), (11.20)

    in particular, see (9.37),

    Prob(BL=L)=p0(L)=g(L)Zf(w,L)1w.\mathop{\rm Prob}\nolimits(B_{L}=L)=\mathrm{p}_{0}(L)=\frac{g(L)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}\to 1-w. (11.21)

Let us estimate, for later use, the probability that BLB_{L} is less than L/2L/2. The result depends on the value of θ{\theta}.

\bullet If θ<1{\theta}<1, using (11.19), we have

Prob(BLL/2)wc1wLθ1/21dλλ1+θ(1λ)θwc1wB12(1θ,θ)Lθ.\mathop{\rm Prob}\nolimits(B_{L}\leq L/2)\approx\frac{wc}{1-w}L^{-{\theta}}\int_{1/2}^{1}\frac{{\rm d}{\lambda}}{{\lambda}^{1+{\theta}}(1-{\lambda})^{{\theta}}}\approx\frac{wc}{1-w}\mathrm{B}_{\frac{1}{2}}\Big{(}1-{\theta},-{\theta}\Big{)}L^{-{\theta}}. (11.22)

\bullet If θ>1{\theta}>1, using (11.18), we have

Prob(BLL/2)wθ(1w)Lb=0L/2g(b)wθ(1w)Lb=0Lg(b)wθX(1w)L.\mathop{\rm Prob}\nolimits(B_{L}\leq L/2)\approx\frac{w\,{\theta}}{(1-w)L}\sum_{b=0}^{L/2}g(b)\approx\frac{w\,{\theta}}{(1-w)L}\sum_{b=0}^{L}g(b)\approx\frac{w\,{\theta}\langle X\rangle}{(1-w)L}. (11.23)
Refer to caption
Figure 10: Free renewal processes: distributions π(k|L)\pi(k|L) (in green) and Prob(BL=k)\mathop{\rm Prob}\nolimits(B_{L}=k) (in blue) for Example 1 (see (5.1)), with w=0.8w=0.8 and L=60L=60. For this value of ww, the asymptotic average number NL4\langle N_{L}\rangle\to 4.
Refer to caption
Figure 11: Free renewal processes: distributions Prob(BL=k)+π(k|L)(1+NLk)\mathop{\rm Prob}\nolimits(B_{L}=k)+\pi(k|L)(1+\langle N_{L-k}\rangle) (in green) and p(k|L)p(k|L) (in blue) for Example 1 (see (5.1)), with w=0.8w=0.8 and L=60L=60. The distribution q(k|L)q(k|L) is the restriction to k>L/2k>L/2 of Prob(BL=k)+π(k|L)(1+NLk)\mathop{\rm Prob}\nolimits(B_{L}=k)+\pi(k|L)(1+\langle N_{L-k}\rangle) according to (9.58). For this value of ww, the asymptotic average number NL4\langle N_{L}\rangle\to 4. This figure is to be compared to the insets of figures 5 and 9.

11.4 Regimes for the distribution of the longest interval

The bulk of the distribution of XmaxX_{\rm max} lies in the region k>L/2k>L/2, and is therefore given by q(k|L)q(k|L), see (9.56) or (9.58).

We start by giving an illustration. The distribution p(k|L)p(k|L) of XmaxX_{\rm max} on the whole interval (0,L)(0,L) is depicted in figure 11 for Example 1 (see (5.1)). The restriction of this distribution to the second half k>L/2k>L/2, that is q(k|L)q(k|L), computed from (9.58) is also depicted. The corresponding figure for θ>1{\theta}>1 is qualitatively alike.

Let us now compare the respective contributions of each of the two terms in (9.56) to the total weight in the region k>L/2k>L/2. The first term is investigated in §11.3 above. The asymptotic estimate of the second term is as follows. We start with the case θ<1{\theta}<1.

\bullet If θ<1{\theta}<1, we consider two regimes.
(i) The main contribution of the second term to the total weight comes from the regime LkLL-k\sim L. Using the asymptotic estimate for large LL,

(gZtdZtd)(w,L)g(L)(1w)2(12wc1wθΓ(θ)2Γ(12θ)Lθ),\big{(}g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}}\big{)}(w,L)\approx\frac{g(L)}{(1-w)^{2}}\Big{(}1-\frac{2wc}{1-w}\frac{{\theta}\Gamma(-{\theta})^{2}}{\Gamma(1-2{\theta})L^{\theta}}\Big{)}, (11.24)

and setting Lk=λLL-k={\lambda}L, the second term reads

wf(k)Zf(w,L)c(1w)2θ(Lk)θwc1wL1θλθ(1λ)1+θ,\frac{wf(k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}\frac{c}{(1-w)^{2}{\theta}(L-k)^{\theta}}\approx\frac{wc}{1-w}\frac{L^{-1-{\theta}}}{{\lambda}^{{\theta}}(1-{\lambda})^{1+{\theta}}}, (11.25)

which is similar to (11.19). Summing this expression upon λ{\lambda} from 0 to 1/21/2 yields (11.22), that is Prob(BLL/2)\mathop{\rm Prob}\nolimits(B_{L}\leq L/2). Therefore adding this contribution to the first one, namely Prob(BL>1/2)\mathop{\rm Prob}\nolimits(B_{L}>1/2), gives unity, up to small corrections, in agreement with the fact that the weight of p(k|L)p(k|L) in the left domain k<L/2k<L/2 is negligible.

(ii) If LkL-k is finite, then the second term reads

wf(k)(gZtdZtd)(w,Lk)Zf(w,L)(1w)wθL(gZtdZtd)(w,Lk),\frac{wf(k)\big{(}g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}}\big{)}(w,L-k)}{Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L)}\approx\frac{(1-w)w{\theta}}{L}\big{(}g\star Z^{\mathop{\scriptstyle\mathrm{td}}}\star Z^{\mathop{\scriptstyle\mathrm{td}}}\big{)}(w,L-k), (11.26)

which is subdominant compared to (11.20).

In order to get the correction of Xmax\langle X_{\rm max}\rangle to LL, we take the average of (9.56), by integrating each of the terms from 0 to 1/21/2 upon λ{\lambda}, using (11.19) and (11.25). Adding the contributions coming from the two terms, we finally obtain for the dominant correction,

LXmax\displaystyle L-\langle X_{\rm max}\rangle \displaystyle\approx wc1w(B12(1θ,1θ)+B12(2θ,θ))L1θ\displaystyle\frac{wc\,}{1-w}\left(\mathrm{B}_{\frac{1}{2}}(1-{\theta},1-{\theta})+\mathrm{B}_{\frac{1}{2}}(2-{\theta},-{\theta})\right)L^{1-{\theta}} (11.27)
=\displaystyle= wc1wB12(1θ,θ)L1θ,\displaystyle\frac{wc\,}{1-w}\mathrm{B}_{\frac{1}{2}}\Big{(}1-{\theta},-{\theta}\Big{)}L^{1-{\theta}},

which has the same structure as (8.31) or (3.6). The comment made below (8.32) also holds here. In the present case the expression in the right side of (11.27) is proportional to NL=w/(1w)\langle N_{L}\rangle=w/(1-w) times the critical mean interval X|LL1θ\langle X|L\rangle\sim L^{1-{\theta}}, as given in gl2001 .

\bullet Likewise, for θ>1{\theta}>1, the weight of the first term in (9.56) dominates upon the second one, and we find for the correction of the mean to LL,

LXmaxw1wX,L-\langle X_{\rm max}\rangle\approx\frac{w}{1-w}\langle X\rangle, (11.28)

showing that this correction is made of NL=w/(1w)\langle N_{L}\rangle=w/(1-w) intervals of size X\langle X\rangle. This expression is therefore the perfect parallel of (8.32) or (3.8).

11.5 Probability for the last interval to be the longest one

Lastly, we investigate the behaviour of QL{Q_{L}} defined in (9.61) as the probability that BLB_{L} is the longest interval. An estimate of QL{Q_{L}} for w<1w<1 can be obtained by means of the inequality

1QLProb(BLL/2).1-{Q_{L}}\lesssim\mathop{\rm Prob}\nolimits(B_{L}\leq L/2). (11.29)

In view of (11.22) and (11.27), we infer that asymptotically for LL large, if θ<1{\theta}<1,

QLXmaxL,{Q_{L}}\approx\frac{\langle X_{\rm max}\rangle}{L}, (11.30)

while, if θ>1{\theta}>1, in view of (11.23) and (11.28),

1QLθL(LXmax).1-Q_{L}\approx\frac{{\theta}}{L}(L-\langle X_{\rm max}\rangle). (11.31)

In other words, for w<1w<1, QL1Q_{L}\to 1. At criticality, w=1w=1, QLQ=0.626Q_{L}\to Q_{\infty}=0.626\dots, if θ<1{\theta}<1, while if θ>1{\theta}>1, QLL(11/θ)Q_{L}\sim L^{-(1-1/{\theta})} gms2015 . For w>1w>1, QL0Q_{L}\to 0.

Figure 12 depicts QLQ_{L} as a function of ww for Example 1 (see (5.1)) and for three different sizes, crossing at the universal critical value Q=0.626Q_{\infty}=0.626\dots for w=1w=1 gms2015 and the data collapse obtained by using the scaling variable x=(w1)L1/2x=(w-1)L^{1/2}.

Refer to caption
Figure 12: Free renewal processes: probability QL{Q_{L}} that the last interval, BLB_{L}, is the longest one, for Example 1 (see (5.1)), for three different values of LL. The curves cross at the universal critical value Q=0.626Q_{\infty}=0.626\dots for w=1w=1. Inset: after rescaling, QLQ_{L} against the scaling variable x=(w1)L1/2x=(w-1)L^{1/2}.

Table 2 summarises the results found in section 11 and recapitulates the results for the two other phases (disordered and critical). This table demonstrates a large degree of universality of the results, as was the case of table 1, with which it should be put in perspective.

Table 2: Dominant asymptotic behaviours at large LL for free renewal processes with power-law distribution (1.1) for f(k)f(k), in the different phases. The results in columns 2 and 3 (critical phase) are taken from gl2001 ; gms2015 . In the last column, X|LL1θ\langle X|L\rangle\sim L^{1-{\theta}} if θ<1{\theta}<1, or X|Lconstant\langle X|L\rangle\approx\mathrm{constant} if θ>1{\theta}>1.
disordered critical θ<1{\theta}<1 critical θ>1{\theta}>1 condensed
NL\langle N_{L}\rangle LX\frac{L}{\langle X\rangle} LθL^{\theta} LX\frac{L}{\langle X\rangle} w1w\frac{w}{1-w}
X|L\langle X|L\rangle constant\mathrm{constant} L1θL^{1-{\theta}} X\langle X\rangle L1θL^{1-{\theta}} or constant
BL\langle B_{L}\rangle constant LL L2θL^{2-{\theta}} LL
Xmax\langle X_{\rm max}\rangle lnL\ln L LL L1/θL^{1/{\theta}} LL
Zf(w,L)Z^{\mathop{\scriptstyle\mathrm{f}}}(w,L) eL/ξ{\rm e}^{L/\xi} 11 11 LθL^{-{\theta}}

12 Conclusion

Let us summarise the salient aspects of this study.

We first recalled the main features of the condensation transition taking place for random allocation models and zrp in the thermodynamic limit (L,nL,n\to\infty with fixed ratio ρ=L/n\rho=L/n), when the distribution of occupations is subexponential. These occupations are independent and identically distributed random variables conditioned by the value of their sum. The phase diagram is made of three phases: disordered, critical, and condensed. The critical line ρ=ρc(θ)\rho=\rho_{c}({\theta}), where θ>1{\theta}>1, separates the disordered phase at low density from the condensed phase at high density. Condensation manifests itself by the occurrence, in the thermodynamic limit, of a unique site with macroscopic occupation. In the language of particles and boxes (or sites), the condensate is by definition the site with the largest occupation. In the language of sums of random variables used all throughout the present work, the condensate XmaxX_{{\rm max}} is the unique summand with extensive value. In the thermodynamic limit, the fraction Xmax/LX_{\rm max}/L no longer fluctuates and takes the asymptotic value 1ρc/ρ1-\rho_{c}/\rho.

A second scenario for the same class of models consists in taking the LL\to\infty limit keeping the number of sites (or summands) fixed. In this limit there is again a single extensive summand XmaxX_{\rm max}, but now the fraction Xmax/LX_{\rm max}/L tends to unity, which means that condensation is total. The novelty is that this occurs irrespective of the existence of a first moment X\langle X\rangle, or in other words, irrespective of whether θ{\theta} is smaller or larger than one. If LL is large but finite, the distribution of XmaxX_{\rm max} is peaked, with a width LXmaxL-\langle X_{\rm max}\rangle scaling as L1θL^{1-{\theta}} if θ<1{\theta}<1, with a known amplitude, or asymptotically equal to (n1)X(n-1)\langle X\rangle, if θ>1{\theta}>1. Note that, in contrast to the previous case, one can no longer speak of a phase transition, nor even of a phase, since the system is made of a finite number nn of summands (or sites).

This scenario is a good preparation for the study of condensation in free and tied-down renewal processes, with power-law distribution of intervals (1.1), which is the main motivation of the present work. Instead of particle occupations and sites one speaks in terms of renewal events and intervals, whose sizes sum up to a fixed value LL. The novelty—and complication—is that the number of these renewal events, or equivalently of intervals, NLN_{L}, fluctuates. For instance these renewal points are the passages by the origin of a random walk, as depicted in figure 3. A weight ww is attached to each renewal event. In the language of random walks (or of polymer chains) ww represents the reward or penalty when the walk touches the origin fisher ; gia1 ; gia2 . A high value of ww favours configurations with a large number of intervals NLN_{L}, i.e., a disordered phase—or localised phase in the language of random walks. A low value of ww favours configurations with a small number of intervals NLN_{L}, i.e., a condensed phase—or delocalised phase in the language of random walks. It is therefore intuitively clear that the same scenario of total condensation as seen above should prevail, where now the driving force is no longer a change in the density, ρ\rho, but a change in the value of the weight ww attached to each interval (or summand). In this respect it is worth noting the similarity between equations (3.6), (8.31) and (11.27) on one hand, and the similarity between equations (3.8), (8.32) and (11.28) on the other hand.

It turns out that, in the condensed phase, when LL\to\infty, the distribution of the number of intervals, NLN_{L}, is superuniversal, i.e., model independent, since it only depends on ww and not even on the index θ{\theta} of the power-law decay (1.1). This distribution is geometric for free renewal processes, while it is the convolution of two such distributions for tdrp. More generally, an important distinction is to be made according to whether θ{\theta} is less or larger than unity. In the first case the distribution f(k)f(k) has no first moment, atypical events play a major role and the system becomes self-similar at criticality. In the second case the observables of interest depend on the first moment X\langle X\rangle, which is finite.

In closing, let us broaden the perspective. The phase transition occurring when ww passes through unity is second order for the density of intervals ν\nu (defined in (6.8)) if θ<1{\theta}<1, and first order if θ>1{\theta}>1. On the other hand the correlation length diverges at the transition, see (6.13). The transition is therefore mixed order as was pointed out for the particular case of tdrp with Example 2 (see (5.7)) in burda3 ; bar2 . Furthermore, the magnetisation, defined as the alternating sum m=(X1X2+X3)/Lm=(X_{1}-X_{2}+X_{3}-\cdots)/L, changes, when LL\to\infty, from the value 0 in the disordered phase to ±1\pm 1 in the condensed phase since condensation is total. More on this can be found in bar2 . If θ<1{\theta}<1 the distribution of the magnetisation at criticality is broad and self-similar, both for free lamperti58 ; gl2001 and tied-down renewal processes wendel2 . At criticality, for θ<1{\theta}<1, the non stationary two-time (or two-space) correlation function is also self-similar, again for both processes gl2001 ; wendel2 .666After submission of the present work, a study devoted to the statistics of XmaxX_{\rm max} in the range (L/2,L)(L/2,L) for tied-down or free renewal processes at criticality (w=1w=1) was presented in barkai . For the tdrp case the result (4.48) with w=1w=1 is obtained. For the free renewal case, barkai predicts, if w=1w=1, q(k|L)=g(k)(NLkfNLk1f)+f(k)(1+NLkf),q(k|L)=g(k)(\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L-k}\rangle-\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L-k-1}\rangle)+f(k)(1+\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L-k}\rangle), (12.1) which is (9.57), with w=1w=1, noting that Ztd(1,L)=NLfNL1fZ^{\mathop{\scriptstyle\mathrm{td}}}(1,L)=\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L}\rangle-\langle N^{\mathop{\scriptstyle\mathrm{f}}}_{L-1}\rangle, as is clear by taking the generating functions of both sides.

Acknowledgements.
It is a pleasure to thank G Giacomin, M Loulakis and J-M Luck for enlightening discussions. I am also indebted to S Grosskinsky and S Janson for useful correspondence.

Appendix

Appendix A On equation (3.18)

Let us explain the argument leading to (3.18) and the origin of the hierarchical structure mentioned in §3.3777I am indebted to M Loulakis for sharing his comments on this part with me..

  1. 1.

    In the uphill region where LXmaxL-X_{\rm max} is finite, XmaxX_{\rm max} is the unique big summand and (3.16) holds. This property stems from the fact that when the sum of nn subexponential random variables is conditioned to a large value LL, all the dependence is absorbed by the maximum and the ensemble of n1n-1 smaller variables becomes asymptotically independent. This property, initially put forward by early workers, has been progressively refined in subsequent studies ferrari ; armendariz2011 ; janson .

  2. 2.

    In the dip region, where Xmax>L/2X_{\rm max}>L/2, since LXmaxL-X_{\rm max} gets large, the sum i=1n1Xi\sum_{i=1}^{n-1}X_{i} becomes subjected to a large deviation event. This event will be realised by X(2)X^{(2)}, the second largest summand, typically equal to LXmaxL-X_{\rm max}. We thus obtain (3.17).

  3. 3.

    One can now iterate the reasoning. If Xmax=kL/2X_{\rm max}=k\leq L/2, the difference LkkL-k\geq k cannot accommodate a single big summand X(2)=jX^{(2)}=j since the latter should be less than XmaxX_{\rm max}. Now

    LXmaxX(2)Li=1n2Xi.L-X_{\rm max}-X^{(2)}\mathrel{\mathop{\approx}\limits_{L\to\infty}}\sum_{i=1}^{n-2}X_{i}. (A.1)

    where the sum in the right side is subjected to a large deviation, which will be realised by a third large summand X(3)X^{(3)}. Since LkjL-k-j should be less than jj, the constraint j(Lk)/2j\geq(L-k)/2 holds. Moreover X(2)XmaxX^{(2)}\leq X_{\rm max} imposes the condition (Lk)/2k(L-k)/2\leq k, i.e. kL/3k\geq L/3. We are thus lead to the asymptotic estimate

    Prob(Xmax=k|Sn=L)j=Lk2kn(n1)f(k)f(j)Zn2(Lkj)Zn(L),\mathop{\rm Prob}\nolimits(X_{\rm max}=k|S_{n}=L)\approx\sum_{j=\frac{L-k}{2}}^{k}n(n-1)\frac{f(k)f(j)Z_{n-2}(L-k-j)}{Z_{n}(L)}, (A.2)

    and therefore

    Prob(L/3XmaxL/2|Sn=L)k=L/3L/2j=Lk2kn(n1)f(k)f(j)Zn2(Lkj)Zn(L).\mathop{\rm Prob}\nolimits(L/3\leq X_{\rm max}\leq L/2|S_{n}=L)\mathrel{\mathop{\approx}\limits}\sum_{k=L/3}^{L/2}\ \sum_{j=\frac{L-k}{2}}^{k}n(n-1)\frac{f(k)f(j)Z_{n-2}(L-k-j)}{Z_{n}(L)}. (A.3)

For θ<1{\theta}<1, the analysis of this expression in the continuum limit leads to

Prob(L/3XmaxL/2|Sn=L)(n1)(n2)c2A(θ)L2θ,\mathop{\rm Prob}\nolimits(L/3\leq X_{\rm max}\leq L/2|S_{n}=L)\approx(n-1)(n-2)c^{2}\frac{A({\theta})}{L^{2{\theta}}}, (A.4)

where the amplitude A(θ)A({\theta}) is given by

A(θ)=1/31/2dx1x2xdy1[xy(1xy)]1+θ.A({\theta})=\int_{1/3}^{1/2}{\rm d}x\,\int_{\frac{1-x}{2}}^{x}{\rm d}y\,\frac{1}{[xy(1-x-y)]^{1+{\theta}}}. (A.5)

For instance, A(1/2)=2πA(1/2)=2\pi, A(1/3)=93Γ(2/3)3/(4π)A(1/3)=9\sqrt{3}\,\Gamma(2/3)^{3}/(4\pi).

For θ>1{\theta}>1, the analysis of (A.3) yields

Prob(L/3<Xmax<L/2|Sn=L)(n1)c 22+2θL1+θ(n2)X2,\mathop{\rm Prob}\nolimits(L/3<X_{\rm max}<L/2|S_{n}=L)\approx\frac{(n-1)c\,2^{2+2{\theta}}}{L^{1+{\theta}}}\frac{(n-2)\langle X\rangle}{2}, (A.6)

for continuous random variables, and

Prob(L/3XmaxL/2|Sn=L)(n1)c 22+2θL1+θ(n2)X+12,\mathop{\rm Prob}\nolimits(L/3\leq X_{\rm max}\leq L/2|S_{n}=L)\approx\frac{(n-1)c\,2^{2+2{\theta}}}{L^{1+{\theta}}}\frac{(n-2)\langle X\rangle+1}{2}, (A.7)

for discrete ones.

One can iterate the reasoning leading to (A.3) and derive the weights of the successive sectors (L/4,L/3)(L/4,L/3), (L/5,L/4(L/5,L/4), etc. For instance, for XmaxX_{\rm max} in the interval (L/4,L/3)(L/4,L/3), one finds

Prob(Xmax=k|Sn=L)j=Lk3ki=Lkj2jn(n1)(n2)f(k)f(j)f(i)Zn3(Lkji)Zn(L),\mathop{\rm Prob}\nolimits(X_{\rm max}=k|S_{n}=L)\approx\sum_{j=\frac{L-k}{3}}^{k}\sum_{i=\frac{L-k-j}{2}}^{j}n(n-1)(n-2)\frac{f(k)f(j)f(i)Z_{n-3}(L-k-j-i)}{Z_{n}(L)}, (A.8)

which yields

Prob(L/4XmaxL/3|Sn=L)Lγ,γ={3θif θ22(1+θ)if θ>2,\mathop{\rm Prob}\nolimits(L/4\leq X_{\rm max}\leq L/3|S_{n}=L)\sim L^{-\gamma},\qquad\gamma=\left\{\begin{array}[]{ll}3{\theta}&\textrm{if }{\theta}\leq 2\vspace{4pt}\\ 2(1+{\theta})&\textrm{if }{\theta}>2,\end{array}\right. (A.9)

For example, if θ=1/2{\theta}=1/2, one finds

Prob(L/4XmaxL/3|Sn=L)(n1)(n2)(n3)c32π/3L3/2.\mathop{\rm Prob}\nolimits(L/4\leq X_{\rm max}\leq L/3|S_{n}=L)\approx(n-1)(n-2)(n-3)c^{3}\frac{\sqrt{2}\,\pi/3}{L^{3/2}}. (A.10)

Appendix B Weight of the maximum in the left region for a Lévy 12\frac{1}{2} stable law

We want to determine the weight of the maximum in the left region considered in §3.3,

Pn=Prob(XmaxL/2|Sn=L)=k=0L/2pn(k|L)=1k=L/2+1Lpn(k|L)=1k=L/2+1Lnπn(k|L),P_{n}=\mathop{\rm Prob}\nolimits(X_{\rm max}\leq L/2|S_{n}=L)=\sum_{k=0}^{L/2}p_{n}(k|L)=1-\sum_{k=L/2+1}^{L}p_{n}(k|L)=1-\sum_{k=L/2+1}^{L}n\pi_{n}(k|L), (B.1)

on the particular example of a a Lévy 12\frac{1}{2} stable law. We use a continuum formalism where distributions are densities and variables are real numbers for the particular case where f(k)f(k) is the Lévy 12\frac{1}{2} stable density (7.11),

f(k)=12,c(k)=ck3/2eπc2/k.f(k)=\mathcal{L}_{\frac{1}{2},c}(k)=\frac{c}{k^{3/2}}{\rm e}^{-\pi c^{2}/k}. (B.2)

Likewise, considering LL as a real number and Zn(L)Z_{n}(L) as a density,

Zn(L)=12,nc(k)=ncL3/2en2πc2/L.Z_{n}(L)=\mathcal{L}_{\frac{1}{2},nc}(k)=\frac{nc}{L^{3/2}}{\rm e}^{-n^{2}\pi c^{2}/L}. (B.3)

Thus

πn(k|L)=f(k)Zn1(Lk)Zn(L)\pi_{n}(k|L)=\frac{f(k)Z_{n-1}(L-k)}{Z_{n}(L)} (B.4)

is explicit. Setting k=L/tk=L/t, we obtain

Pn=1(n1)cL12dtteπc2(nt)2L(t1)(t1)3/2.P_{n}=1-\frac{(n-1)c}{\sqrt{L}}\int_{1}^{2}{\rm d}t\,\frac{t\,{\rm e}^{-\frac{\pi c^{2}(n-t)^{2}}{L(t-1)}}}{(t-1)^{3/2}}. (B.5)

Setting c/L=b/(2π)c/\sqrt{L}=b/(2\sqrt{\pi}), we finally get

Pn=12(2nerfc((n2)b2)+(n2)e(n1)b2erfc(nb2)).P_{n}=\frac{1}{2}\Bigg{(}2-n\mathop{\rm erfc}\Big{(}\frac{(n-2)b}{2}\Big{)}+(n-2){\rm e}^{(n-1)b^{2}}\mathop{\rm erfc}\Big{(}\frac{nb}{2}\Big{)}\Bigg{)}. (B.6)

For LL large, expanding in powers of bb, we obtain

Pn\displaystyle P_{n} =\displaystyle= (n1)(n2)(b22nb33π+(n1)b44+)\displaystyle(n-1)(n-2)\Big{(}\frac{b^{2}}{2}-n\frac{b^{3}}{3\sqrt{\pi}}+(n-1)\frac{b^{4}}{4}+\cdots\Big{)} (B.7)
=\displaystyle= (n1)(n2)(2πc2L8πn3c3L3/2+(n1)4π2c4L2+).\displaystyle(n-1)(n-2)\Big{(}2\pi\frac{c^{2}}{L}-\frac{8\pi n}{3}\frac{c^{3}}{L^{3/2}}+(n-1)4\pi^{2}\frac{c^{4}}{L^{2}}+\cdots\Big{)}.

For Example 1, c=1/(2π)c=1/(2\sqrt{\pi}) (see (5.1)), the first term of the expansion,

Pn(n1)(n2)2L,P_{n}\approx\frac{(n-1)(n-2)}{2L}, (B.8)

matches the predictions made in (A.4) and (A.5).

References

  • (1) Bialas P, Burda Z and Johnston D 1997 Nucl. Phys. B 493 505 Condensation in the Backgammon model
  • (2) Bialas P, Bogacz L, Burda Z and Johnston D 2000 Nucl. Phys. B 575 599 Finite size scaling of the balls in boxes model
  • (3) Janson S 2012 Prob. Surveys 9 103 Simply generated trees, conditioned Galton-Watson trees, random allocations and condensation
  • (4) Spitzer F 1970 Advances in Math. 5 246 Interaction of Markov Processes
  • (5) Andjel ED 1982 Ann. Prob. 10 525 Invariant measures for the zero-range process
  • (6) Drouffe JM, Godrèche C and Camia F 1998 J. Phys. A 31 L19 A simple stochastic model for the dynamics of condensation
  • (7) Evans MR 2000 Braz. J. Phys. 30 42 Phase transitions in one-dimensional nonequilibrium systems
  • (8) Jeon I, March P and Pittel B 2000 Ann. Probab. 28 1162 Size of the largest cluster under zero-range invariant measures
  • (9) Godrèche C and Luck JM 2002 J. Phys.: Condens. Matter 14 1601 Nonequilibrium dynamics of urn models
  • (10) Godrèche C 2003 J. Phys. A 36 6313 Dynamics of condensation in zero-range processes
  • (11) Grosskinsky S, Schütz GM and Spohn H 2003 J. Stat. Phys. 113 389 Condensation in the zero range process: stationary and dynamical properties
  • (12) Evans MR and Hanney T 2005 J. Phys. A 38 R195 Nonequilibrium Statistical Mechanics of the Zero-Range Process and Related Models
  • (13) Godrèche C and Luck J M 2005 J. Phys. A 38 7215 Dynamics of the condensate in zero-range processes
  • (14) Godrèche C 2007 Lect. Notes Phys. Springer 716 261 From urn models to zero-range processes: statics and dynamics
  • (15) Evans MR, Majumdar SN and Zia RKP 2006 J. Stat. Phys. 123 357 Canonical Analysis of Condensation in Factorised Steady States
  • (16) Ferrari PA, Landim C and Sisko VV 2007 J. Stat. Phys. 128 1153 Condensation for a Fixed Number of Independent Random Variables
  • (17) Evans MR and Majumdar SN 2008 J. Stat. Mech. P05004 Condensation and Extreme Value Statistics
  • (18) Armendariz I and Loulakis M 2009 Probab. Theory Relat. Fields 145 175 Thermodynamic limit for the invariant measures in supercritical zero range processes
  • (19) Armendariz I and Loulakis M 2011 Stochastic Process. Appl. 121 1138 Conditional distribution of heavy tailed random variables on large deviations of their sum
  • (20) Armendariz I, Grosskinsky S and Loulakis M 2013 Stochastic Process. Appl. 123 3466 Zero-range condensation at criticality
  • (21) Armendariz I, Grosskinsky S and Loulakis M 2017 Probab. Theory Relat. Fields 169 105 Metastability in a condensing zero-range process in the thermodynamic limit
  • (22) Godrèche C 2019 J. Stat. Mech. 063207 Condensation for random variables conditioned by the value of their sum
  • (23) Feller W 1968 1971 An Introduction to Probability Theory and its Applications volumes 1&2 (New York: Wiley)
  • (24) Doob JL 1948 Trans. Amer. Math. Soc. 63 422 Renewal theory from the point of view of the theory of probability
  • (25) Smith WL 1958 J. Roy. Statist. Soc. Ser. B 20 243 Renewal Theory and its Ramifications
  • (26) Cox DR 1962 Renewal theory (London: Methuen)
  • (27) Wendel JG 1964 Math. Scand. 14 21 Zero-Free Intervals of Semi-Stable Markov Processes
  • (28) Godrèche C 2017 J. Phys. A 50 195003 Longest interval between zeros of the tied-down random walk, the Brownian bridge and related renewal processes
  • (29) Godrèche C 2017 J. Stat. Mech. P073205 Two-time correlation and occupation time for the Brownian bridge and tied-down renewal processes
  • (30) Labarbe JM and Marckert JF 2007 Elect. Journ. Prob. 12 229 Asymptotics of Bernoulli random walks, bridges, excursions and meanders with a given number of peaks
  • (31) Vervaat W 1979 Ann. Probab. 7 143 A relation between Brownian bridge and Brownian excursion
  • (32) Fisher M 1984 J. Stat. Phys. 34 667 Walks, Walls, Wetting and Melting
  • (33) Poland D and Scheraga HA 1966 J. Chem. Phys. 45 1456 Phase Transitions in One Dimension and the Helix-Coil Transition in Polyamino Acids
  • (34) Poland D and Scheraga HA 1966 J. Chem. Phys. 45 1464 Occurrence of a Phase Transition in Nucleic Acid Models
  • (35) Bialas P, Burda Z and Johnston D 1999 Nucl. Phys. B 542 413 Phase diagram of the mean field model of simplicial gravity
  • (36) Bar A and Mukamel D 2014 J. Stat. Mech. P11001 Mixed order transition and condensation in an exactly soluble one dimensional spin model
  • (37) Bar A, Majumdar SN, Schehr G and Mukamel D 2016 Phys. Rev. 93 052130 Exact extreme-value statistics at mixed-order transitions
  • (38) Barma M, Majumdar SN and Mukamel D 2019 J. Phys. A 52 254001 Fluctuation-dominated phase ordering at a mixed order transition
  • (39) Giacomin G 2007 Random polymer models (Imperial College Press: World Scientific)
  • (40) Giacomin G 2011 Disorder and critical phenomena through basic probability models École d’été de probabilités de Saint-Flour Lect. Notes Math. Springer
  • (41) Godrèche C and Luck JM 2001 J. Stat. Phys. 104 489 Statistics of the occupation time of renewal processes
  • (42) Godrèche C Majumdar SN and Schehr G 2015 J. Stat. Mech. P03014 Statistics of the longest interval in renewal processes
  • (43) Bardou F, Bouchaud J P, Aspect A and Cohen-Tannoudji C 2002 Lévy Statistics and Laser Cooling (Cambridge: Cambridge University Press)
  • (44) Barkai E 2003 Phys. Rev. Lett. 90 104101 Aging in Subdiffusion Generated by a Deterministic Dynamical System
  • (45) J. Beltrán J and C. Landim C 2012 Probab. Theory Relat. Fields 152 781 Metastability of reversible condensed zero range processes on a finite set
  • (46) Gnedenko B V and Kolmogorov A 1954 Limit Distributions for Sums of Independent Random Variables (Cambridge, Mass: Addison-Wesley)
  • (47) Godrèche C and Luck JM 2001 Eur. Phys. J. B 23, 473 Nonequilibrium dynamics of the zeta urn model
  • (48) Gumbel E J 1958 Statistics of Extremes (New York: Columbia University Press)
  • (49) Gnedenko B V 1943 Ann. Math. 44 423 Sur la distribution limite du terme maximum d’une série aléatoire
  • (50) Fréchet M 1927 Ann. Soc. Polon. Math. 6 93 Sur la loi de probabilité de l’écart maximum
  • (51) Chistyakov VP 1964 Theor. Probab. Appl. 9 640 A theorem on sums of independent positive random variables and its applications to branching random processes
  • (52) Dynkin EB 1955 Izv. Akad. Nauk. SSSR Ser. Math. 19 247; 1961 Selected Translations Math. Stat. Prob. 1 171 Limit theorems for sums of independent random quantities
  • (53) Lamperti J 1958 Trans. Amer. Math. Soc. 88 380 An occupation time theorem for a class of stochastic processes
  • (54) Lamperti J 1961 Am. Math. Soc. 12 724 A contribution to renewal theory
  • (55) Höll M, Wang W and Barkai E arXiv:2006.06253 Extreme value theory for constrained physical systems