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

Approximation of smooth numbers for harmonic samples: a Stein method approach

Arturo Jaramillo, Xiaochuan Yang Arturo Jaramillo: Department of Probability and statistics, Centro de Investigación en matemáticas (CIMAT) [email protected] Xiaochuan Yang: Department of Mathematics, Brunel University London [email protected]
Abstract.

We present a de Bruijn-type approximation for quantifying the content of mm-smooth numbers, derived from samples obtained through a probability measure over the set {1,,n},\{1,\dots,n\}, with point mass function at kk proportional to 1/k1/k. Our analysis is based on a stochastic representation of the measure of interest, utilizing weighted independent geometric random variables. This representation is analyzed through the lens of Stein’s method for the Dickman distribution. A pivotal element of our arguments relies on precise estimations concerning the regularity properties of the solution to the Dickman-Stein equation for Heaviside functions, recently developed by Bhattacharjee and Schulte in [2]. Remarkably, our arguments remain mostly in the realm of probability theory, with Mertens’ first and third theorems standing as the only number theory estimations required.

Key words and phrases:
Probabilistic number theory, Dickman approximation, Stein method
1991 Mathematics Subject Classification:
60F05, 11K65

1. Introduction

Let 𝒫\mathcal{P} denote the set of non-negative prime numbers and [n][n] the integer interval [n]:={1,,n}[n]:=\{1,\dots,n\}. Define the function ψ:\psi:\mathbb{N}\rightarrow\mathbb{N} through

ψ(k):=max{p𝒫;p divides k}.\psi(k):=\max\{p\in\mathcal{P}\ ;\ p\text{ divides }k\}.

We say that a given natural number kk is mm-smooth, with mm\in\mathbb{N}, if ψ(k)m\psi(k)\leq m. Namely, if the prime factors of kk are at most mm. The primary goal of this manuscript is to introduce a new probabilistic measurement ΨH[m,n]\Psi_{H}[m,n] for the content of the set of mm-smooth numbers bounded by nn, and find sharp approximations for it. Our principal contribution is an explicit estimation for ΨH[m,n]\Psi_{H}[m,n], expressed in terms of the Dickman function, in the spirit of the work developed by de Bruijn in [5].

1.1. Overview on the asymptotics for the size of mm-smooth numbers

Let 𝒵[n,m]\mathcal{Z}[n,m] denote the collection of mm-smooth numbers bounded by nn and Ψ(n,m)\Psi(n,m) its cardinality

Ψ(n,m):=|𝒵[n,m]|.\displaystyle\Psi(n,m):=|\mathcal{Z}[n,m]|. (1.1)

The set 𝒵[n,m]\mathcal{Z}[n,m] is of fundamental importance in the field of number theory, finding its most notable application in the formulation of the Hildebrand condition for the resolution of the Riemann hypothesis. This condition, first presented in [12] and succinctly outlined in Theorem 2.3 below, involves a precise estimation for Ψ(n,m)\Psi(n,m), and serves as a sufficient criterion for validating the Riemann hypothesis.

Many pieces of work have been devoted to the study of the function Ψ(n,m)\Psi(n,m). Some of the most significant ones for purposes of our paper, have been documented in the compendium [14] (see also [10]), to which the reader is referred for a thorough examination of the topic. From a probabilistic number theory point of view, we can write

Ψ(n,m)\displaystyle\Psi(n,m) =n[Jn𝒵[n,m]]=n[ψ(Jn)m],\displaystyle=n\mathbb{P}[J_{n}\in\mathcal{Z}[n,m]]=n\mathbb{P}[\psi(J_{n})\leq m],

where JnJ_{n} is a uniform random variable defined over a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}). This perspective can be utilized as a pivot for proposing alternative ways for describing 𝒵[m,n]\mathcal{Z}[m,n], based on modifications of the law of JnJ_{n} (see (1.2) below). These adjustments should be designed in such a way that the resulting problem becomes more tractable while retaining the capacity to recover information for Ψ(n,m)\Psi(n,m). Although we will mainly focus on providing a detailed resolution to the “simplified problem”, we will briefly describe a methodology for transferring useful estimations to the study of Ψ(m,n)\Psi(m,n).

Our modification on the law of JnJ_{n} is largely inspired by the recent paper [4], which introduced the idea of using the harmonic distribution in [n][n] as a tool for understanding divisibility properties of uniform samples. More precisely, we let HnH_{n} be a collection of random variables defined in (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), with probability mass at kk proportional to k1𝟙[n](k)k^{-1}\mathbbm{1}_{[n]}(k). We then consider the function

ΨH[n,m]\displaystyle\Psi_{H}[n,m] :=n[Hn𝒵[n,m]]=n[ψ(Hn)m],\displaystyle:=n\mathbb{P}[H_{n}\in\mathcal{Z}[n,m]]=n\mathbb{P}[\psi(H_{n})\leq m], (1.2)

as measurement of content for 𝒵[n,m]\mathcal{Z}[n,m]. The advantage of dealing with the law of HnH_{n} instead of JnJ_{n} resides in the fact that the prime multiplicities of HnH_{n} are easily described in terms of independent geometric random variables, as highlighted in Proposition 4.1 below. This will allow us to simplify the task of estimating [ψ(Hn)m]\mathbb{P}[\psi(H_{n})\leq m], to approximating in sup-norm, the cumulative distribution of a random sum of the form

1λmp𝒫[n]log(p)ξp,\displaystyle\frac{1}{\lambda_{m}}\sum_{p\in\mathcal{P}\cap[n]}\log(p)\xi_{p},

where the ξp\xi_{p} are independent geometric variables supported over 0:={0}\mathbb{N}_{0}:=\mathbb{N}\cup\{0\}, with success probability 11/p1-1/p. This reduction transforms the nature of the problem to a purely probabilistic one, where the recent advances in the theory of Stein’s method for Dickman distributed random variables can be implemented (see Section 4 for further information regarding the Dickman distribution).

The aforementioned connection between the functions ΨH\Psi_{H} and Ψ\Psi stems from [4, Lemma 3.1], which asserts that the variable JnJ_{n} can be sharply approximated by a product of the form QnHnQ_{n}H_{n}, where QnQ_{n}, conditional on the value of HnH_{n}, is uniformly distributed in the set (𝒫[n/Hn]){1}(\mathcal{P}\cap[n/H_{n}])\cup\{1\}. The precise statement establishes that for all n21n\geq 21,

dTV(Jn,HnQn)\displaystyle d_{TV}(J_{n},H_{n}Q_{n}) 61loglog(n)log(n),\displaystyle\leq 61\frac{\log\log(n)}{\log(n)}, (1.3)

where dTVd_{TV} denotes the distance in total variation between the probability distributions associated to the inputs. In particular, if it holds that for all strictly positive ε\varepsilon, the set 𝒮^ε\hat{\mathcal{S}}_{\varepsilon}, defined through (2.3), is such that probability [ψ(QnHn)m]\mathbb{P}[\psi(Q_{n}H_{n})\leq m] can be sharply approximated, uniformly over (n,m)𝒮^ε(n,m)\in\hat{\mathcal{S}}_{\varepsilon}, then we would obtain the validity of the Hildebrand condition [12] and consequently, the resolution of the Riemann hypothesis. This problem escapes the reach of our manuscript and is postponed for future research.

1.2. Some questions for future research

Although the problem of obtaining sharp estimates for Ψ(n,m)\Psi(n,m) will not be addressed in this paper, we would like to bring the reader’s attention to some ideas that might be useful for the treatment of this particular problem. First we notice that a good approximation for [ψ(Hn)m]\mathbb{P}[\psi(H_{n})\leq m] is a fundamental first step for this task. Indeed, in virtue of (1.3), sharp estimations of Ψ(n,m)\Psi(n,m) can be obtained from their counterpart for ΨH(n,m)\Psi_{H}(n,m). In order to verify this, we write

[ψ(HnQn)m]\displaystyle\mathbb{P}[\psi(H_{n}Q_{n})\leq m] =𝔼[𝟙{ψ(Hn)m}𝟙{Qnm}]=𝔼[𝟙{ψ(Hn)m}𝔼[𝟙{Qnm}|Hn]].\displaystyle=\mathbb{E}[\mathbbm{1}_{\{\psi(H_{n})\leq m\}}\mathbbm{1}_{\{Q_{n}\leq m\}}]=\mathbb{E}[\mathbbm{1}_{\{\psi(H_{n})\leq m\}}\mathbb{E}[\mathbbm{1}_{\{Q_{n}\leq m\}}\ |\ H_{n}]].

Splitting the event {ψ(Hn)m}\{\psi(H_{n})\leq m\} into its components induced by the partition

{{n/Hnm},{n/Hn>m}},\{\{n/H_{n}\leq m\},\{n/H_{n}>m\}\},

and using the fact that for all hh\in\mathbb{N}, it holds that

[Qnm|Hn=h]\displaystyle\mathbb{P}[Q_{n}\leq m\ |\ H_{n}=h] ={1 if h>n/m(π(m)+1)/(π(n/h)+1) if hn/m,\displaystyle=\left\{\begin{array}[]{ccc}1&\text{ if }&h>n/m\\ (\pi(m)+1)/(\pi(n/h)+1)&\text{ if }&h\leq n/m,\end{array}\right.

where π:+\pi:\mathbb{R}_{+}\rightarrow\mathbb{R} denotes the prime-counting function

π(x)\displaystyle\pi(x) :=|𝒫[1,x]|,\displaystyle:=|\mathcal{P}\cap[1,x]|,

we deduce the inequality

[ψ(HnQn)m]\displaystyle\mathbb{P}[\psi(H_{n}Q_{n})\leq m] =[ψ(Hn)m,n/m<Hn]+𝔼[𝟙{ψ(Hn)m}𝟙{Hnn/m}π(m)+1π(n/Hn)+1],\displaystyle=\mathbb{P}[\psi(H_{n})\leq m,n/m<H_{n}]+\mathbb{E}\left[\mathbbm{1}_{\{\psi(H_{n})\leq m\}}\mathbbm{1}_{\{H_{n}\leq n/m\}}\frac{\pi(m)+1}{\pi(n/H_{n})+1}\right],

By a conditioning argument, the above identity yields

[ψ(HnQn)m]\displaystyle\mathbb{P}[\psi(H_{n}Q_{n})\leq m] =[ψ(Hn)m][n/m<Hn|ψ(Hn)m]\displaystyle=\mathbb{P}[\psi(H_{n})\leq m]\mathbb{P}[n/m<H_{n}\ |\ \psi(H_{n})\leq m]
+[ψ(Hn)m]𝔼[π(m)+1π(n/Hn)+1𝟙{Hnn/m}|ψ(Hn)m].\displaystyle+\mathbb{P}[\psi(H_{n})\leq m]\mathbb{E}\left[\frac{\pi(m)+1}{\pi(n/H_{n})+1}\mathbbm{1}_{\{H_{n}\leq n/m\}}\ |\ \psi(H_{n})\leq m\right].

From the above discussion, we conclude that a sharp approximation for the distribution function of ψ(Hn)\psi(H_{n}) (or equivalently, for the function ΨH\Psi_{H}) can be used to transfer approximations of

𝔼[(1+π(m)+1π(n/Hn)+1)𝟙{Hnn/m}|ψ(Hn)m]\displaystyle\mathbb{E}\left[\left(1+\frac{\pi(m)+1}{\pi(n/H_{n})+1}\right)\mathbbm{1}_{\{H_{n}\leq n/m\}}\ |\ \psi(H_{n})\leq m\right] (1.4)

to approximations for the distribution of ψ(HnQn)\psi(H_{n}Q_{n}), and ultimately, to Ψ(n,m)\Psi(n,m). Arguably, the most natural approach for carrying out an analysis for (1.4) would be to utilize the available estimations regarding the prime counting function. In particular, by [23], the inequality

|π(n)nlog(n)n(logn)2|\displaystyle|\pi(n)-\frac{n}{\log(n)}-\frac{n}{(\log n)^{2}}| 184n(logn)3,\displaystyle\leq\frac{184n}{(\log n)^{3}}, (1.5)

holds for n229n\geq 229 (see also [4, Section 3]), suggesting that an estimation of (1.4) could be obtained through an analysis of

𝔼[(1+m/log(m)+m/log(m)2+1n/(Hnlog(n/Hn))+n/(Hnlog(n/Hn)2)+1)𝟙{Hnn/m}|ψ(Hn)m].\displaystyle\mathbb{E}\left[\left(1+\frac{m/\log(m)+m/\log(m)^{2}+1}{n/(H_{n}\log(n/H_{n}))+n/(H_{n}\log(n/H_{n})^{2})+1}\right)\mathbbm{1}_{\{H_{n}\leq n/m\}}\ |\ \psi(H_{n})\leq m\right]. (1.6)

The above term can be approximated, contingent on sharp estimates for

𝔼[gm(n/Hn)|ψ(Hn)m],\displaystyle\mathbb{E}\left[g_{m}(n/H_{n})\ |\ \psi(H_{n})\leq m\right],

for

gm(x)\displaystyle g_{m}(x) :=(1+m/log(m)+m/log(m)2+1x(log(x)+log(x)2)+1)𝟙{mx}.\displaystyle:=\left(1+\frac{m/\log(m)+m/\log(m)^{2}+1}{x(\log(x)+\log(x)^{2})+1}\right)\mathbbm{1}_{\{m\leq x\}}.

Approximations on this object are closely related to the description of the limiting behavior of the measure μn,m\mu_{n,m}, obtained as the distribution of n/Hnn/H_{n}, conditional to ψ(Hn)m\psi(H_{n})\leq m. We conjecture that the measure μn,m\mu_{n,m} concentrates around its mean and, up to a suitable normalization constant, admits an asymptotic limiting distribution. We intend to address this topic in a forthcoming paper.

1.3. An overview on the description of ΨH\Psi_{H}

Now we turn our attention to the function ΨH\Psi_{H}, which is the main focus of our investigation. Similarly to Ψ\Psi, the description of the asymptotic behavior of ΨH\Psi_{H} turns out to be closely related to the so-called Dickman function ρ\rho, defined as the unique solution to the initial value problem of the delay type

uρ(u)=ρ(u1),ρ(u)=1 for u[0,1].\displaystyle u\rho^{\prime}(u)=-\rho(u-1),\quad\quad\quad\quad\quad\rho(u)=1\text{ for }u\in[0,1]. (1.7)

However, the structure of dependence on mm and nn through ρ\rho for the approximation of ΨH(m,n)\Psi_{H}(m,n), is different from its counterpart for Ψ(m,n)\Psi(m,n). In particular, our main result, presented in Theorem 3.1, combined with Theorem 2.2, imply that the quotient

(Ψ(m,n)/ΨH(m,n))/(log(n)/log(m))(\Psi(m,n)/\Psi_{H}(m,n))/(\log(n)/\log(m))

converges to one as the quotient log(n)/log(m)\log(n)/\log(m) tends to infinity. In particular, Ψ(m,n)\Psi(m,n) is not asymptotically equivalent to ΨH(m,n)\Psi_{H}(m,n) as log(n)/log(m)\log(n)/\log(m) tends to infinity. A second remarkable feature of ΨH\Psi_{H} is the fact that it can be sharply estimated uniformly over m,nm,n\in\mathbb{N}, unlike Ψ(m,n)\Psi(m,n), whose approximation has shown to be tractable only for regimes of mm and nn for which log(n)/log(m)f(n)\log(n)/\log(m)\leq f(n), with f(x)f(x) being an adequate function satisfying f(x)xf(x)\leq\sqrt{x}.

The approach we follow for studying ΨH\Psi_{H} is entirely probabilistic, and up to some important but tractable technical computations, consists of two fundamental steps: (i) First, we utilize a representation of HnH_{n} in terms of independent geometric random variables conditioned on an adequate event. This reduces the analysis of [Hn𝒵[m,n]]\mathbb{P}[H_{n}\in\mathcal{Z}[m,n]], to the estimation of the distribution FnF_{n} of the sum of weighted geometric random variables satisfying that FnF_{n} converges pointwise as nn tends to infinity to the Dickman function (ii) As a second step, we measure the quality of the aforementioned approximation in the sup-norm, which is done by interpreting ρ\rho as a constant multiple of the cumulative distribution function of a random variable following the so-called Dickman law and then applying the Stein-method theory for the Dickman distribution introduced in [1]. For readability, the proof of our main result is presented conditional on the validity of a series of key steps, whose proof is postponed to later sections.

The rest of the paper is organized as follows: in Section 2, we review some of the work related to the estimations of Ψ\Psi and their relevance in the field of number theory. In Section 3, we present our main results and briefly explain the key ingredients involved in its resolution. In Section 4, we present some preliminary tools to be utilized throughout the paper and set the appropriate notation. In Section 5 we present the proof of the main theorem and finally, in sections 6 and 7, we prove some technical lemmas utilized in the proof of our main result.

2. Some known approximation results for Ψ(n,m)\Psi(n,m)

In this section, we revise some of the known results regarding the study of approximations for Ψ(n,m)\Psi(n,m). The material of this section is largely taken from [14].

2.1. The function Ψ\Psi and its relation to the Dickman function

In the seminal paper [7], Dickman presented a groundbreaking result regarding the behavior of the function Ψ\Psi and its relation to the solution to the delay equation (1.7). Dickman showed that for every x+,x\in\mathbb{R}_{+}, the following convergence holds

limn1nΨ(n,n1/x)\displaystyle\lim_{n\rightarrow\infty}\frac{1}{n}\Psi(n,n^{1/x}) =ρ(x).\displaystyle=\rho(x).

This relation was further refined by Ramaswami in [18], who proved that for any fixed x1x\geq 1, there exists Cx>0C_{x}>0, such that

|1nΨ(n,n1/x)ρ(x)|\displaystyle\left|\frac{1}{n}\Psi\left(n,n^{1/x}\right)-\rho(x)\right| Cxlog(n).\displaystyle\leq\frac{C_{x}}{\log(n)}. (2.1)

These two results where subsequently explored thoroughly by many authors, seeking for extensions that allow for uniformity over xx. To pin down precisely the aforementioned notion of uniformity, we observe that if we set the smoothness threshold n1/xn^{1/x} appearing in (2.1), to be equal to mm, we have that x=Υ(n,m)x=\Upsilon(n,m), where Υ:\Upsilon:\mathbb{N}\rightarrow\mathbb{R} is given by

Υ(n,m)\displaystyle\Upsilon(n,m) :=log(n)log(m),\displaystyle:=\frac{\log(n)}{\log(m)},

so that the Dickman approximation (2.1), reads Ψ(n,m)/nρΥ(n,m)\Psi\left(n,m\right)/n\approx\rho\circ\Upsilon(n,m), provided that we take m=n1/xm=n^{1/x}. Ramaswani’s approximation, on the other hand, becomes

|1nΨ(n,m)ρΥ(n,m)|\displaystyle\left|\frac{1}{n}\Psi(n,m)-\rho\circ\Upsilon(n,m)\right| CΥ(n,m)/log(n).\displaystyle\leq C_{\Upsilon(n,m)}/\log(n).

One of the most significant generalizations to the above results was done by de Bruijn in [5], who proved Theorem 2.1 below

Theorem 2.1 (de Bruijn, 1951).

For a fixed value of ε>0\varepsilon>0, define the set

𝒮ε\displaystyle\mathcal{S}_{\varepsilon} ={(n,m)2|n3 and Υ(n,m)log(x)3/5ε}\displaystyle=\{(n,m)\in\mathbb{N}^{2}\ |\ n\geq 3\text{ and }\Upsilon(n,m)\leq\log(x)^{3/5-\varepsilon}\}

Then, there exists a constant CεC_{\varepsilon} only depending on ε\varepsilon, such that

sup(m,n)𝒮ε1Mm,n|Ψ(n,m)nρΥ(n,m)|\displaystyle\sup_{(m,n)\in\mathcal{S}_{\varepsilon}}\frac{1}{M_{m,n}}\left|\Psi(n,m)-n\rho\circ\Upsilon(n,m)\right| Cε,\displaystyle\leq C_{\varepsilon}, (2.2)

where

Mm,n\displaystyle M_{m,n} :=nρΥ(n,m)log(Υ(n,m)+1)log(m).\displaystyle:=\frac{n\rho\circ\Upsilon(n,m)\log(\Upsilon(n,m)+1)}{\log(m)}.

Otherwise said, it holds that

Ψ(n,m)\displaystyle\Psi(n,m) =nρΥ(n,m)(1+O(log(Υ(n,m)+1)log(m)))\displaystyle=n\rho\circ\Upsilon(n,m)\left(1+O\left(\frac{\log(\Upsilon(n,m)+1)}{\log(m)}\right)\right)

The mathematical community has put many efforts in finding sharp upper bounds asymptotically equivalent to (2.2), uniformly over a larger set than 𝒮ε\mathcal{S}_{\varepsilon}. Next, we mention some of the most relevant ones for the purposes of our paper. We begin with the work by Hensley in [11], where it was shown that the relaxation

Ψ(n,m)\displaystyle\Psi(n,m) nρΥ(n,m)(1+O(log(Υ(n,m)+1)log(m)))\displaystyle\geq n\rho\circ\Upsilon(n,m)\left(1+O\left(\frac{\log(\Upsilon(n,m)+1)}{\log(m)}\right)\right)

holds uniformly over considerably larger set than 𝒮ε\mathcal{S}_{\varepsilon}, thus reducing the problem of estimating the exact value of Ψ(n,m)\Psi(n,m), to simply finding upper bounds for it. A crucial improvement to the de Bruijn estimation was done by Hildebrand in [13], where the following result was proved

Theorem 2.2 (Hildebrand, 1986).

For every ε>0\varepsilon>0, there exits a constant Cε>0C_{\varepsilon}>0, such that the estimation (2.2) holds uniformly over (n,m)𝒮~ε(n,m)\in\tilde{\mathcal{S}}_{\varepsilon}, with

𝒮~ε\displaystyle\tilde{\mathcal{S}}_{\varepsilon} ={(n,m)2|m1 and Υ(n,m)elog(m)3/5ε}.\displaystyle=\{(n,m)\in\mathbb{N}^{2}\ |\ m\neq 1\text{ and }\Upsilon(n,m)\leq e^{\log(m)^{3/5-\varepsilon}}\}.
Remark 2.1.

The function elog(m)3/5εe^{\log(m)^{3/5-\varepsilon}} appearing in the threshold is asymptotically smaller than any function of the form yαy^{\alpha}, for every α>0\alpha>0.

Shortly after, Hildebrand presents in [12], an equivalence between the resolution of this problem and the Riemann hypothesis. The precise statement reads as follows

Theorem 2.3 (Hildebrand, 1984).

For a fixed value of ε>0\varepsilon>0, we define the set

𝒮^ε\displaystyle\hat{\mathcal{S}}_{\varepsilon} :={(n,m)2|m1 and Υ(n,m)m1/2ε}.\displaystyle:=\{(n,m)\in\mathbb{N}^{2}\ |\ m\neq 1\text{ and }\Upsilon(n,m)\leq m^{1/2-\varepsilon}\}. (2.3)

Then, provided that exists a mapping εCε\varepsilon\mapsto C_{\varepsilon} satisfying

sup(m,n)𝒮^εlog(m)log(Υ(n,m)+1)|log(Ψ(n,m)nρΥ(n,m))|\displaystyle\sup_{(m,n)\in\hat{\mathcal{S}}_{\varepsilon}}\frac{\log(m)}{\log(\Upsilon(n,m)+1)}\left|\log\left(\frac{\Psi(n,m)}{n\rho\circ\Upsilon(n,m)}\right)\right| Cε,\displaystyle\leq C_{\varepsilon}, (2.4)

then the Riemann hypothesis holds.

Remark 2.2.

The condition (2.4) can be rephrased in the ”big OO” notation as

Ψ(n,m)\displaystyle\Psi(n,m) =nρΥ(n,m)exp{O(log(Υ(n,m)+1)log(m))}.\displaystyle=n\rho\circ\Upsilon(n,m)\exp\left\{O\left(\frac{\log(\Upsilon(n,m)+1)}{\log(m)}\right)\right\}.

Although we do not treat directly with the function Ψ\Psi, the above results will serve as benchmarks for setting up appropriate parallelisms for ΨH\Psi_{H}.

3. Main results

This section describes our findings regarding the analysis of ΨH\Psi_{H}. Since all of the prime divisors of HnH_{n} lie in [n][n], then ψH(n,m)=1\psi_{H}(n,m)=1 for all mnm\geq n, so the problem of estimating ψH(n,m)\psi_{H}(n,m) can be localized into the regime mnm\leq n.

In the sequel, LmL_{m} will denote the Harmonic partial sum in [m][m] and ImI_{m} the product of the prime numbers belonging to [m][m], namely,

Lm:=j=1m1j Im:=pm(11/p).\displaystyle L_{m}:=\sum_{j=1}^{m}\frac{1}{j}\ \ \ \ \ \ \ \ \ \ \ \ \  I_{m}:=\prod_{p\leq m}(1-1/p). (3.1)

The next result is the main contribution of our paper

Theorem 3.1.

There exists a constant C>0C>0, such that for 16m16\leq m,

supmnlog(n)|ΨH(n,m)nΥ(n,m)[ρ]Υ(n,m)|\displaystyle\sup_{m\leq n}\log(n)\left|\Psi_{H}(n,m)-\frac{n}{\Upsilon(n,m)}\mathcal{I}[\rho]\circ\Upsilon(n,m)\right| C,\displaystyle\leq C, (3.2)

where \mathcal{I} denotes the integral operator

[ρ](x)\displaystyle\mathcal{I}[\rho](x) :=0xρ(s)𝑑s.\displaystyle:=\int_{0}^{x}\rho(s)ds.

Otherwise said, the function ΨH\Psi_{H} satisfies

ΨH(n,m)\displaystyle\Psi_{H}(n,m) =nΥ(n,m)[ρ]Υ(n,m)+O(1/log(n)),\displaystyle=\frac{n}{\Upsilon(n,m)}\mathcal{I}[\rho]\circ\Upsilon(n,m)+O(1/\log(n)),

uniformly over 16mn16\leq m\leq n.

A version of Theorem 3.1 was first addressed by Bruijn and Van Lint in [6] for the case where log(n)log(m)log(n)\sqrt{\log(n)}\leq\log(m)\leq\log(n). Related work followed in the contributions of Song [19] and Tenenbaum [21], where 1/n1/n was replaced by a more general weight h(n)/nh(n)/n. While these results share some similarities with our approach, there are two fundamental differences. First, our method is entirely probabilistic, whereas all the aforementioned works rely on techniques from complex analysis. Second, the range of values for which our result is applicable is significantly broader, contrasting with the conditions in [6], [19], and [21].

Our approach for proving (3.1) is based on a representation of the law of HnH_{n} in terms of a sequence of independent geometric random variables {ξp}p𝒫\{\xi_{p}\}_{p\in\mathcal{P}}, with ξp\xi_{p} satisfying

[ξp=k]\displaystyle\mathbb{P}[\xi_{p}=k] =(11/p)pk,\displaystyle=(1-1/p)p^{-k}, (3.3)

for k0k\geq 0. As explained in detail in Section 5, a stochastic representation of the multiplicities of HnH_{n} in terms of the ξp\xi_{p}, will allow us to reduce the proof of Theorem 4.1, to showing that the variables

Zm\displaystyle Z_{m} :=p𝒫[m]log(p)ξp,\displaystyle:=\sum_{p\in\mathcal{P}\cap[m]}\log(p)\xi_{p}, (3.4)

satisfy

|[𝔼[Zm]1Zmz]eγρ(z)|\displaystyle|\mathbb{P}[\mathbb{E}[Z_{m}]^{-1}Z_{m}\leq z]-e^{-\gamma}\rho(z)| Clog(m)(1+z1),\displaystyle\leq\frac{C}{\log(m)}(1+z^{-1}), (3.5)

where z0z\geq 0 and CC is a universal constant independent of mm and zz. This approximation is the main achievement of our proof, and its resolution will make use of the Stein method for the Dickman distribution, first introduced by Bhattacharjee and Goldstein in [1] for handling approximations in the Wasserstein distance for Dickman distributed random variables, and further refined by Bhattacharjee and Schulte in [2] for dealing with its counterpart for the Kolmogorov metric. Equation (3.5) resambles the findings from [2, Theorem 1.1] and is of interest on its own.

4. Preliminaries

In this section, we present some important preliminaries required for the proof of Theorem 3.1. There are four fundamental parts that will be utilized: (i) A methodology for assessing Kolmogorov distances for Dickman approximation, previously studied in [2, 1] (ii) A stochastic representation of the multiplicities of HnH_{n} (iii) Some elementary number theory estimations by Mertens and (iv) Basic formulas regarding bias transforms.

4.1. Dickman distributional approximations

In the sequel, γ\gamma will denote the Euler-Mascheroni constant. Useful information on ρ\rho can be obtained from a probabilistic perspective by regarding the mapping xeγρ(x)x\mapsto e^{-\gamma}\rho(x) as the density of a random variable. The verification of the condition +eγρ(x)=1\int_{\mathbb{R}_{+}}e^{-\gamma}\rho(x)=1 can be obtained from an adequate analysis of the Laplace transform of ρ\rho, as explained in [14, Lemma 2.6.] (see also [15]). We say that a random variable DD is distributed according to the Dickman law if its cumulative distribution function is given by eγ[ρ]e^{-\gamma}\mathcal{I}[\rho], where [ρ]\mathcal{I}[\rho] denotes the integral of ρ\rho over [0,x][0,x]. The main use we will give to this interpretation is the well-known characterization of the Dickman distribution in terms of its Bias transform (see [1]), which states that for every bounded and smooth function f:+f:\mathbb{R}_{+}\rightarrow\mathbb{R}, it holds that

𝔼[Df(D)]\displaystyle\mathbb{E}[Df(D)] =01𝔼[f(D+t)]𝑑t.\displaystyle=\int_{0}^{1}\mathbb{E}[f(D+t)]dt. (4.1)

This identity can be easily obtained from the fact that if DD is distributed according to eγρe^{-\gamma}\rho and UU is a uniform random variable independent of DD, then

D\displaystyle D =LawU(D+1).\displaystyle\stackrel{{\scriptstyle Law}}{{=}}U(D+1). (4.2)

The reader is referred to [17] for a proof of (4.2). The study of distributional approximations of random variables XX by means of sharp estimations on 𝔼[Xf(X)]\mathbb{E}[Xf(X)], for ff ranging over an adequate collection of test functions initiated with the work by Stein in [20] in the study of Gaussian approximations for sums of weakly dependent random variables and has been developed for many different distributions, including exponential, beta, gamma, Poisson, geometric and Dickman, among many others (see [16, 3, 9, 8]). A very useful adaptation of this methodology to the realm of weighted sums of random variables was presented in Theorem 1.4 and Theorem 1.5 in [2], whose proof serves as starting point for our application in hand. Replacing the role of ff in (4.1) by ff^{\prime}, one obtains the following characterization, which can be found in [1, Lemma 3.2]

Lemma 4.1.

The variable WW follows a Dickman law if and only if for every Lipchitz function ff, the following identity holds

𝔼[Wf(W)f(W+1)+f(W)]=0.\displaystyle\mathbb{E}[Wf^{\prime}(W)-f(W+1)+f(W)]=0.

Let 𝒞\mathcal{C} denote the collection of test functions

𝒞\displaystyle\mathcal{C} :={𝟙[0,z];z+}.\displaystyle:=\{\mathbbm{1}_{[0,z]}\ ;\ z\in\mathbb{R}_{+}\}.

Taking into consideration the above lemma, we implement Stein’s heuristics in the context for the Dickman law, which roughly speaking, aims to conclude an approximation of the type 𝔼[h(W)]𝔼[h(D)]\mathbb{E}[h(W)]\approx\mathbb{E}[h(D)] for h𝒞h\in\mathcal{C} by an approximation of the type

𝔼[Wf(W)f(W+1)+f(W)]0,\mathbb{E}[Wf^{\prime}(W)-f(W+1)+f(W)]\approx 0,

for ff belonging to a family of test functions determined by 𝒞\mathcal{C}. This idea can be formalized by considering the Stein equation

xfz(x)+fz(x)fz(x+1)=hz(x)𝔼[hz(D)],\displaystyle xf_{z}^{\prime}(x)+f_{z}(x)-f_{z}(x+1)=h_{z}(x)-\mathbb{E}[h_{z}(D)], (4.3)

where hz(x):=𝟙[0,z](x)h_{z}(x):=\mathbbm{1}_{[0,z]}(x). Evaluation at xx equal to the variable of interest, taking expectation and then supremum with respect to zz, then yields a tractable expression for the distance in Kolmogorov of the variable of interest and the Dickman law. The solution to (4.3), along with a description of its regularity properties is presented in Theorem 1.9 and Lemma 3.3, from [2]. These results are summarized in Lemma 4.2 below.

Lemma 4.2.

Let DD be a random variable following the Dickman law. There exists a solution fzf_{z} to the equation (4.3), admitting a decomposition of the type fz=f1,z+f2,zf_{z}=f_{1,z}+f_{2,z}, where

f1,z(x)\displaystyle f_{1,z}(x) :=1(z/x)[Dz],\displaystyle:=1\wedge(z/x)-\mathbb{P}[D\leq z], (4.4)

and f2,zf_{2,z} satisfies f2,z=uz,+uz,f_{2,z}^{\prime}=u_{z,+}-u_{z,-}, for functions uz,+,uz,u_{z,+},u_{z,-} non-negative and non-increasing, with

uz,+,uz,1.\|u_{z,+}\|_{\infty},\|u_{z,-}\|_{\infty}\leq 1.

Moreover, for every random variable W0W\geq 0 and constants t>0t>0, α>0\alpha>0 and u0u\geq 0,

𝔼[αt(W+u)1+t𝟙{W+u>α}]\displaystyle\mathbb{E}\left[\frac{\alpha^{t}}{(W+u)^{1+t}}\mathbbm{1}_{\{W+u>\alpha\}}\right] ζ(1+t)(C+2dK(W,D)αu),\displaystyle\leq\zeta(1+t)\left(C+2\frac{d_{K}(W,D)}{\alpha\vee u}\right),

for some constant C0C\geq 0, where ζ\zeta denotes the Riemann zeta function.

4.2. Mertens’ formulas

Next we present some elementary number theory estimations related to partial products and partial sums over the primes for the functions (11/p)(1-1/p) and log(p)/p\log(p)/p, respectively. Such results are attributed to Franz Mertens and the estimations are referred to as “first, second and third Mertens’ formulas”. The first one establishes that

|p𝒫nlog(p)plog(n)|\displaystyle|\sum_{p\in\mathcal{P}_{n}}\frac{\log(p)}{p}-\log(n)| 2log(n).\displaystyle\leq\frac{2}{\log(n)}. (4.5)

The second Merten’s formula reads

|p𝒫n1ploglog(n)c1|\displaystyle\left|\sum_{p\in\mathcal{P}_{n}}\frac{1}{p}-\log\log(n)-c_{1}\right| 5/log(n),\displaystyle\leq 5/\log(n), (4.6)

where c1c_{1} is a universal constant, while the third guarantees the existence of a constant C>0C>0, such that

|log(m)Imeγ|\displaystyle|\log(m)I_{m}-e^{-\gamma}| C1log(m).\displaystyle\leq C\frac{1}{\log(m)}. (4.7)

The reader is referred to [22], pages 15-17, for a proof of these estimations.

4.3. The geometric distribution and the multiplicities of Harmonic samples

Let {ξp}p𝒫\{\xi_{p}\}_{p\in\mathcal{P}} be a sequence of independent random variables with distribution given as in (3.3). For k,p𝒫k\in\mathbb{N},p\in\mathcal{P}, let αp(k)\alpha_{p}(k) denote the largest mm satisfying pm|kp^{m}|k. Therefore kk is uniquely determined by αp\alpha_{p}’s through the prime factorisation k=p𝒫pαp(k)k=\prod_{p\in\mathcal{P}}p^{\alpha_{p}(k)}. Recall the definitions of LnL_{n} and InI_{n}, given by (3.1). The following result, first presented in [4], describes the join distribution of the αp(Hn)\alpha_{p}(H_{n})’s.

Proposition 4.1.

The law of HnH_{n} can be represented in terms of the ξp\xi_{p}’s as

((αp(Hn);p𝒫n))\displaystyle\mathcal{L}((\alpha_{p}(H_{n})\ ;\ p\in\mathcal{P}_{n})) =((ξp;p𝒫[n])|An),\displaystyle=\mathcal{L}((\xi_{p}\ ;\ p\in\mathcal{P}\cap[n])\ |\ A_{n}),

where

An\displaystyle A_{n} :={p𝒫[n]pξpn}.\displaystyle:=\{\prod_{p\in\mathcal{P}\cap[n]}p^{\xi_{p}}\leq n\}. (4.8)

Moreover, if m21m\geq 21, then

[An]=LnIn1/2.\displaystyle\mathbb{P}[A_{n}]=L_{n}I_{n}\geq 1/2. (4.9)

The next result gives some useful identities for exploiting the distributional properties of the ξp\xi_{p}’s

Lemma 4.3.

Let ξ\xi be a geometric random variable, with distribution

[ξ=k]\displaystyle\mathbb{P}[\xi=k] =(1θ)θk,k0,θ(0,1).\displaystyle=(1-\theta)\theta^{k},\quad k\geq 0,~{}\theta\in(0,1).

Then, for every compactly supported function f:f:\mathbb{R}\rightarrow\mathbb{R},

𝔼[ξf(ξ)]\displaystyle\mathbb{E}[\xi f(\xi)] =θ1θ𝔼[f(ξ+ξ~+1)].\displaystyle=\frac{\theta}{1-\theta}\mathbb{E}[f(\xi+\tilde{\xi}+1)].

where ξ~\tilde{\xi} is an independent copy of ξ.\xi.

Proof.

It suffices to consider the case f(x)=exp{iλx}f(x)=\exp\{\textbf{i}\lambda x\} for some λ\lambda\in\mathbb{R}. For such instance, we have that

𝔼[ξf(ξ)]\displaystyle\mathbb{E}[\xi f(\xi)] =1iddλ𝔼[exp{iλξ}]=1iddλ1θ1θeiλ\displaystyle=\frac{1}{\textbf{i}}\frac{d}{d\lambda}\mathbb{E}[\exp\{\textbf{i}\lambda\xi\}]=\frac{1}{\textbf{i}}\frac{d}{d\lambda}\frac{1-\theta}{1-\theta e^{\textbf{i}\lambda}}
=θ(1θ)eiλ(1θeiλ)2=θ1θ𝔼[exp{iλ(ξ+ξ~+1)}].\displaystyle=\frac{\theta(1-\theta)e^{\textbf{i}\lambda}}{(1-\theta e^{\textbf{i}\lambda})^{2}}=\frac{\theta}{1-\theta}\mathbb{E}[\exp\{\textbf{i}\lambda(\xi+\tilde{\xi}+1)\}].

The result follows from here. ∎

5. Proof of Theorem 3.1

In the sequel, {Hn}n1\{H_{n}\}_{n\geq 1} will denote a sequence of random variables defined over (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), with HnH_{n} harmonically distributed over [n][n], namely, [Hn=k]:=1Lnk\mathbb{P}[H_{n}=k]:=\frac{1}{L_{n}k} for k[n]k\in[n] and zero otherwise, where LnL_{n} is given by (3.1).

The proof of the bound consists on proving the following steps

Step I
As a first step, we will show the following result

Lemma 5.1.

For every m1m\geq 1, it holds that

[ψ(Hn)m]\displaystyle\mathbb{P}[\psi(H_{n})\leq m] =1LnIm[Smlog(n)λm],\displaystyle=\frac{1}{L_{n}I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right],

where Sm:=Zm/λmS_{m}:=Z_{m}/\lambda_{m}, with ZmZ_{m} defined by (3.4) and λm\lambda_{m} defined through

λm:=q𝒫mlog(q)q.\lambda_{m}:=\sum_{q\in\mathcal{P}_{m}}\frac{\log(q)}{q}.

This way, the problem reduces to the analysis of two pieces: the estimation of the term LnImL_{n}I_{m}, which can be handled by means of Merten’s formula, and the analysis of the probability in the right, which now consists on the treatment of a sum of independent random variables.

Step II
In the second step we will show

Lemma 5.2.

There exists a universal constant C>0C>0, such that

|[Smz]eγ[ρ](z)|\displaystyle|\mathbb{P}[S_{m}\leq z]-e^{-\gamma}\mathcal{I}[\rho](z)| C(1+1/z2)log(m),\displaystyle\leq\frac{C(1+1/z^{2})}{\log(m)},

for some universal constant C>0C>0.

Contingent on the validity of these lemmas, we can prove Theorem 3.1.

Proof of Theorem 3.1.

Firstly, by Lemma 5.1,

|[ψ(Hn)m]1log(m)Υ(n,m)Im[Smlog(n)λm]|=|1LnIm[Smlog(n)λm]1log(n)Im[Smlog(n)λm]|=|1log(n)Ln|1log(n)Im[Smlog(n)λm].\left|\mathbb{P}[\psi(H_{n})\leq m]-\frac{1}{\log(m)\Upsilon(n,m)I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right]\right|\\ \begin{aligned} &=\left|\frac{1}{L_{n}I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right]-\frac{1}{\log(n)I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right]\right|\\ &=\left|1-\frac{\log(n)}{L_{n}}\right|\frac{1}{\log(n)I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right].\end{aligned}

This relation, together with Merten’s lemma and the approximation

log(n)Lnlog(n)+1,\log(n)\leq L_{n}\leq\log(n)+1,

leads to the inequality

|[ψ(Hn)m]1log(m)Υ(n,m)Im[Smlog(n)λm]|\displaystyle\left|\mathbb{P}[\psi(H_{n})\leq m]-\frac{1}{\log(m)\Upsilon(n,m)I_{m}}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right]\right| Clog(n)Υ(n,m).\displaystyle\leq\frac{C}{\log(n)\Upsilon(n,m)}.

By Merten’s approximation (4.7), we then conclude that

|[ψ(Hn)m]eγΥ(n,m)[Smlog(n)λm]|\displaystyle\left|\mathbb{P}[\psi(H_{n})\leq m]-\frac{e^{\gamma}}{\Upsilon(n,m)}\mathbb{P}\left[S_{m}\leq\frac{\log(n)}{\lambda_{m}}\right]\right| CΥ(n,m)(1log(m)+1log(n))K/log(n),\displaystyle\leq\frac{C}{\Upsilon(n,m)}\left(\frac{1}{\log(m)}+\frac{1}{\log(n)}\right)\leq K/\log(n), (5.1)

Thus, by Lemma 5.2,

|[ψ(Hn)m]eγΥ(n,m)[ρ](Υ(n,m)/λmlog(m))|K(1Υ(n,m)log(m)+1log(m))C/log(n),\left|\mathbb{P}[\psi(H_{n})\leq m]-\frac{e^{-\gamma}}{\Upsilon(n,m)}\mathcal{I}[\rho]\left(\Upsilon(n,m)/\frac{\lambda_{m}}{\log(m)}\right)\right|\\ \leq K\left(\frac{1}{\Upsilon(n,m)\log(m)}+\frac{1}{\log(m)}\right)\leq C/\log(n), (5.2)

for some universal constant C>0C>0. By the mean value theorem and the boundedness of ρ\rho, we deduce the existence of a constant C>0C>0 such that

|[ρ](Υ(n,m)/λmlog(m))[ρ](Υ(n,m))|\displaystyle\left|\mathcal{I}[\rho]\left(\Upsilon(n,m)/\frac{\lambda_{m}}{\log(m)}\right)-\mathcal{I}[\rho]\left(\Upsilon(n,m)\right)\right| CΥ(n,m)|1λmlog(m)|.\displaystyle\leq C\Upsilon(n,m)\left|1-\frac{\lambda_{m}}{\log(m)}\right|.

An application of Merten’s approximation (4.5), then yields

|[ρ](Υ(n,m)λmlog(m))[ρ](Υ(n,m))|\displaystyle\left|\mathcal{I}[\rho]\left(\Upsilon(n,m)\frac{\lambda_{m}}{\log(m)}\right)-\mathcal{I}[\rho]\left(\Upsilon(n,m)\right)\right| C1log(m).\displaystyle\leq C\frac{1}{\log(m)}.

Combining this with inequality (5.2) then gives

|[ψ(Hn)m]eγΥ(n,m)[ρ](Υ(n,m))|\displaystyle\left|\mathbb{P}[\psi(H_{n})\leq m]-\frac{e^{-\gamma}}{\Upsilon(n,m)}\mathcal{I}[\rho]\left(\Upsilon(n,m)\right)\right| C1Υ(n,m)log(m)C1log(n).\displaystyle\leq C\frac{1}{\Upsilon(n,m)\log(m)}\leq C\frac{1}{\log(n)}.

Theorem 3.1 follows from here. ∎

6. Proof of Lemma 5.1

Observe that by Proposition 4.1

[ψ(Hn)m]\displaystyle\mathbb{P}[\psi(H_{n})\leq m] =[ψ(p𝒫npαp(Hn))m]\displaystyle=\mathbb{P}[\psi(\prod_{p\in\mathcal{P}_{n}}p^{\alpha_{p}(H_{n})})\leq m]
=[ψ(p𝒫npξp)m|An]\displaystyle=\mathbb{P}[\psi(\prod_{p\in\mathcal{P}_{n}}p^{\xi_{p}})\leq m\ |\ A_{n}]
=1[An][ξq=0 for all q𝒫(m,n] and p𝒫npξpn],\displaystyle=\frac{1}{\mathbb{P}[A_{n}]}\mathbb{P}[\xi_{q}=0\ \text{ for all }\ q\in\mathcal{P}\cap(m,n]\ \text{ and }\ \prod_{p\in\mathcal{P}_{n}}p^{\xi_{p}}\leq n],

Observe that in the event {ξq=0 for all q𝒫(m,n]}\{\xi_{q}=0\ \text{ for all }\ q\in\mathcal{P}\cap(m,n]\}, it holds that the product p𝒫npξp\prod_{p\in\mathcal{P}_{n}}p^{\xi_{p}} coincides with p𝒫mpξp\prod_{p\in\mathcal{P}_{m}}p^{\xi_{p}}. Consequently, by the independence of the ξp\xi_{p}’s as well as identity (4.9), we have that

[ψ(Hn)m]\displaystyle\mathbb{P}[\psi(H_{n})\leq m] =1LnIn(q𝒫(m,n](11/q))[p𝒫mpξpn]\displaystyle=\frac{1}{L_{n}I_{n}}\big{(}\prod_{q\in\mathcal{P}\cap(m,n]}(1-1/q)\big{)}\mathbb{P}[\prod_{p\in\mathcal{P}_{m}}p^{\xi_{p}}\leq n]
=1LnIm[p𝒫mlog(p)ξplog(n)].\displaystyle=\frac{1}{L_{n}I_{m}}\mathbb{P}[\sum_{p\in\mathcal{P}_{m}}\log(p)\xi_{p}\leq\log(n)].

The result easily follows from here.

7. Proof of Lemma 5.2

Let DD a Dickman distributed random variable. Let UU be uniform in (0,1)(0,1) independent of {ξp}p𝒫\{\xi_{p}\}_{p\in\mathcal{P}}. Let fzf_{z} be the solution to the Stein equation (4.3). For convenience in the notation, we will simply write ff instead of fzf_{z}. Define

Smq\displaystyle S_{m}^{q} :=Smlog(q)λmξq,q𝒫m,\displaystyle:=S_{m}-\frac{\log(q)}{\lambda_{m}}\xi_{q},\quad q\in\mathcal{P}_{m},

with

λm:=q𝒫mlog(q)q,\lambda_{m}:=\sum_{q\in\mathcal{P}_{m}}\frac{\log(q)}{q},

and observe that

[Smz][Dz]\displaystyle\mathbb{P}[S_{m}\leq z]-\mathbb{P}[D\leq z] =𝔼[Smf(Sm)f(Sm+U)]\displaystyle=\mathbb{E}[S_{m}f^{\prime}(S_{m})-f^{\prime}(S_{m}+U)]
=1λmq𝒫[m]log(q)𝔼[ξqf(Smq+log(q)λmξq)]𝔼[f(Sm+U)].\displaystyle=\frac{1}{\lambda_{m}}\sum_{q\in\mathcal{P}\cap[m]}\log(q)\mathbb{E}[\xi_{q}f^{\prime}(S_{m}^{q}+\frac{\log(q)}{\lambda_{m}}\xi_{q})]-\mathbb{E}[f^{\prime}(S_{m}+U)].

By Lemma 4.3,

[Smz][Dz]=1λmq𝒫[m]log(q)q(11/q)1𝔼[f(Sm+log(q)λmξ~q+log(q)λm)]𝔼[f(Sm+U)],\mathbb{P}[S_{m}\leq z]-\mathbb{P}[D\leq z]\\ =\frac{1}{\lambda_{m}}\sum_{q\in\mathcal{P}\cap[m]}\frac{\log(q)}{q}(1-1/q)^{-1}\mathbb{E}[f^{\prime}(S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}})]-\mathbb{E}[f^{\prime}(S_{m}+U)],

where {ξ~p}p𝒫\{\tilde{\xi}_{p}\}_{p\in\mathcal{P}} is an independent copy of {ξp}p𝒫\{\xi_{p}\}_{p\in\mathcal{P}}. By conditioning the expectation in the right hand side on the value of ξ~q\tilde{\xi}_{q}, we obtain

[Smz][Dz]\displaystyle\mathbb{P}[S_{m}\leq z]-\mathbb{P}[D\leq z] =T1+T2,\displaystyle=T_{1}+T_{2},

where

T1\displaystyle T_{1} :=1λmq𝒫[m]log(q)q𝔼[f(Sm+log(q)λm)]𝔼[f(Sm+U)]\displaystyle:=\frac{1}{\lambda_{m}}\sum_{q\in\mathcal{P}\cap[m]}\frac{\log(q)}{q}\mathbb{E}[f^{\prime}(S_{m}+\frac{\log(q)}{\lambda_{m}})]-\mathbb{E}[f^{\prime}(S_{m}+U)]
T2\displaystyle T_{2} :=1λmq𝒫[m]log(q)q2(11/q)1𝔼[f(Sm+log(q)λmξ~q+log(q)λm)|ξ~q1].\displaystyle:=\frac{1}{\lambda_{m}}\sum_{q\in\mathcal{P}\cap[m]}\frac{\log(q)}{q^{2}}(1-1/q)^{-1}\mathbb{E}[f^{\prime}(S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}})\ |\ \tilde{\xi}_{q}\geq 1].

Let VmV_{m} be a random variable independent of the ξp\xi_{p}’s, with distribution given by

[Vm=log(q)/λm]\displaystyle\mathbb{P}[V_{m}=\log(q)/\lambda_{m}] =log(q)qλm,\displaystyle=\frac{\log(q)}{q\lambda_{m}}, (7.1)

for q𝒫[m]q\in\mathcal{P}\cap[m] and zero otherwise. With this notation in hand, we can write

T1\displaystyle T_{1} =𝔼[f(Sm+Vm)]𝔼[f(Sm+U)].\displaystyle=\mathbb{E}[f^{\prime}(S_{m}+V_{m})]-\mathbb{E}[f^{\prime}(S_{m}+U)].

By Lemma 4.2, f:=f1+u+uf^{\prime}:=f_{1}^{\prime}+u_{+}-u_{-}, where f1f_{1} is given by (4.4) and u+,uu_{+},u_{-} are non-negative and non-increasing, with u+,u1\|u_{+}\|_{\infty},\|u_{-}\|_{\infty}\leq 1. We can thus write

|T1|\displaystyle|T_{1}| |T1,1|+|T1,2|+|T1,3|,\displaystyle\leq|T_{1,1}|+|T_{1,2}|+|T_{1,3}|,

where

T1,1\displaystyle T_{1,1} :=𝔼[f1(Sm+Vm)]𝔼[f1(Sm+U)]\displaystyle:=\mathbb{E}[f_{1}^{\prime}(S_{m}+V_{m})]-\mathbb{E}[f_{1}^{\prime}(S_{m}+U)]
T1,2\displaystyle T_{1,2} :=𝔼[u+(Sm+Vm)]𝔼[u+(Sm+U)]\displaystyle:=\mathbb{E}[u_{+}(S_{m}+V_{m})]-\mathbb{E}[u_{+}(S_{m}+U)]
T1,3\displaystyle T_{1,3} :=𝔼[u(Sm+Vm)]𝔼[u(Sm+U)].\displaystyle:=\mathbb{E}[u_{-}(S_{m}+V_{m})]-\mathbb{E}[u_{-}(S_{m}+U)]. (7.2)

Let FmF_{m} be the distribution function of VmV_{m} and Fm1F_{m}^{-1} its left inverse. Since Fm1(U)F_{m}^{-1}(U) is distributed according to VmV_{m}, it follows that (Fm1(U),U)(F_{m}^{-1}(U),U) is a coupling for VmV_{m} and UU, which allows us to write Fm1(U)F_{m}^{-1}(U) instead of VmV_{m} in (7). Observe that the function Fm1F_{m}^{-1} remains constant equal to log(p)/λm\log(p)/\lambda_{m} in the intervals Ip=[ap,bp)I_{p}=[a_{p},b_{p}), where

ap:=q𝒫[p]log(q)qλm and bp:=p𝒫[p]log(q)qλm+log(p+)p+λm,\displaystyle a_{p}:=\sum_{q\in\mathcal{P}\cap[p]}\frac{\log(q)}{q\lambda_{m}}\ \ \ \ \ \text{ and }\ \ \ \ \ b_{p}:=\sum_{p\in\mathcal{P}\cap[p]}\frac{\log(q)}{q\lambda_{m}}+\frac{\log(p_{+})}{p_{+}\lambda_{m}}, (7.3)

for x+x_{+}, defined as the smallest prime strictly bigger than xx. Observe that apa_{p} and bpb_{p} also depend on tt and mm, but we have avoided this in the notation for convenience. By the Mertens’ approximation (4.5), there exists a constant C>0C>0 such that

|aplog(p)λm|,|bplog(p)λm|\displaystyle\left|a_{p}-\frac{\log(p)}{\lambda_{m}}\right|,\left|b_{p}-\frac{\log(p)}{\lambda_{m}}\right| Clog(p)log(m).\displaystyle\leq\frac{C}{\log(p)\log(m)}. (7.4)

Localizing VmV_{m} and UU over the intervals of constancy of Fm1F^{-1}_{m}, we can write the terms T1,iT_{1,i} as

T1,1\displaystyle T_{1,1} =p𝒫[m]𝔼[(f1(Sm+log(p)/λm)f1(Sm+U))𝟙Ip(U)]\displaystyle=\sum_{p\in\mathcal{P}\cap[m]}\mathbb{E}[(f_{1}^{\prime}(S_{m}+\log(p)/\lambda_{m})-f_{1}^{\prime}(S_{m}+U))\mathbbm{1}_{I_{p}}(U)]
T1,2\displaystyle T_{1,2} =p𝒫[m]𝔼[(u+(Sm+log(p)/λm)u+(Sm+U))𝟙Ip(U)]\displaystyle=\sum_{p\in\mathcal{P}\cap[m]}\mathbb{E}[(u_{+}(S_{m}+\log(p)/\lambda_{m})-u_{+}(S_{m}+U))\mathbbm{1}_{I_{p}}(U)]
T1,3\displaystyle T_{1,3} =p𝒫[m]𝔼[(u(Sm+log(p)/λm)u(Sm+U))𝟙Ip(U)].\displaystyle=\sum_{p\in\mathcal{P}\cap[m]}\mathbb{E}[(u_{-}(S_{m}+\log(p)/\lambda_{m})-u_{-}(S_{m}+U))\mathbbm{1}_{I_{p}}(U)].

In order to handle the term T1,1T_{1,1}, we proceed as follows. Let ψp,m\psi_{p,m} denote the term

ψp,m\displaystyle\psi_{p,m} :=𝔼[(f1(Sm+log(p)/λm)f1(Sm+U))𝟙Ip(U)],\displaystyle:=\mathbb{E}[(f_{1}^{\prime}(S_{m}+\log(p)/\lambda_{m})-f_{1}^{\prime}(S_{m}+U))\mathbbm{1}_{I_{p}}(U)],

so that T1,1T_{1,1} can be written as

T1,1\displaystyle T_{1,1} =p𝒫[m]ψp,m.\displaystyle=\sum_{p\in\mathcal{P}\cap[m]}\psi_{p,m}.

Since f1f_{1}^{\prime} is supported in [z,)[z,\infty) and f1(x)=zx1f_{1}(x)=zx^{-1} for xzx\geq z, by distinguishing the instances for which Sm+log(p)/λmS_{m}+\log(p)/\lambda_{m} is bigger than or smaller than zz or UU, we can write ψp,m=θp,m+ϑp,m1+ϑp,m2\psi_{p,m}=\theta_{p,m}+\vartheta_{p,m}^{1}+\vartheta_{p,m}^{2}, where

θp,m\displaystyle\theta_{p,m} :=𝔼[(f1(Sm+log(p)/λm)f1(Sm+U))𝟙Ip(U)𝟙{zSm+log(p)/λm,Sm+U})]\displaystyle:=\mathbb{E}[(f_{1}^{\prime}(S_{m}+\log(p)/\lambda_{m})-f_{1}^{\prime}(S_{m}+U))\mathbbm{1}_{I_{p}}(U)\mathbbm{1}_{\{z\leq S_{m}+\log(p)/\lambda_{m},S_{m}+U\}})]
ϑp,m1\displaystyle\vartheta_{p,m}^{1} :=𝔼[f1(Sm+U)𝟙Ip(U)𝟙{Sm+log(p)/λm<zSm+U}]\displaystyle:=-\mathbb{E}[f_{1}^{\prime}(S_{m}+U)\mathbbm{1}_{I_{p}}(U)\mathbbm{1}_{\{S_{m}+\log(p)/\lambda_{m}<z\leq S_{m}+U\}}]
ϑp,m2\displaystyle\vartheta_{p,m}^{2} :=𝔼[f1(Sm+log(p)/λm)𝟙Ip(U)𝟙{Sm+U<zSm+log(p)/λm}].\displaystyle:=\mathbb{E}[f_{1}^{\prime}(S_{m}+\log(p)/\lambda_{m})\mathbbm{1}_{I_{p}}(U)\mathbbm{1}_{\{S_{m}+U<z\leq S_{m}+\log(p)/\lambda_{m}\}}].

By (7.4), there exists C>0C>0 such that

logpλmCλmlog(p)apbplogpλm+Cλmlog(p).\displaystyle\frac{\log p}{\lambda_{m}}-\frac{C}{\lambda_{m}\log(p)}\leq a_{p}\leq b_{p}\leq\frac{\log p}{\lambda_{m}}+\frac{C}{\lambda_{m}\log(p)}.

Consequently, in the event {UIp}\{U\in I_{p}\}, we have that

{Sm+log(p)/λm<zSm+U}{Sm+U<zSm+log(p)/λm}{zSmC/(λmlog(p))log(p)/λmzSm+C/(λmlog(p))}.\{S_{m}+\log(p)/\lambda_{m}<z\leq S_{m}+U\}\cup\{S_{m}+U<z\leq S_{m}+\log(p)/\lambda_{m}\}\\ \subset\{z-S_{m}-C/(\lambda_{m}\log(p))\leq\log(p)/\lambda_{m}\leq z-S_{m}+C/(\lambda_{m}\log(p))\}.

Using the fact that f1(x)f_{1}^{\prime}(x) satisfies f1(x)=zx2f_{1}^{\prime}(x)=-zx^{-2} and |f1(x)|z1|f_{1}^{\prime}(x)|\leq z^{-1} for xzx\geq z, we can easily show that

|ϑp,m2|,|ϑp,m1|\displaystyle|\vartheta_{p,m}^{2}|,|\vartheta_{p,m}^{1}| z1[zSmC/λmlogpλmzSm+C/λm][UIp].\displaystyle\leq z^{-1}\mathbb{P}[z-S_{m}-C/\lambda_{m}\leq\frac{\log p}{\lambda_{m}}\leq z-S_{m}+C/\lambda_{m}]\mathbb{P}[U\in I_{p}].

We will assume without loss of generality that CC\geq1. From here we deduce that

T1,1\displaystyle T_{1,1} p𝒫[m]|θp,m|+z1𝔼[p𝒫[eC+(zSm)λm,eC+(zSm)λm]log(p)plog(m)].\displaystyle\leq\sum_{p\in\mathcal{P}\cap[m]}|\theta_{p,m}|+z^{-1}\mathbb{E}\left[\sum_{p\in\mathcal{P}\cap[e^{-C+(z-S_{m})\lambda_{m}},e^{C+(z-S_{m})\lambda_{m}}]}\frac{\log(p)}{p\log(m)}\right].

Splitting the sum in the right-hand side into the regimes pCp\leq C and p>Cp>C, we deduce that

T1,1\displaystyle T_{1,1} p𝒫[m]|θp,m|+z1supy+p𝒫[CeC+y,CeC+y]log(p)plog(m)+z1p𝒫[1,C]log(p)plog(m).\displaystyle\leq\sum_{p\in\mathcal{P}\cap[m]}|\theta_{p,m}|+z^{-1}\sup_{y\in\mathbb{R}_{+}}\sum_{p\in\mathcal{P}\cap[C\vee e^{-C+y},C\vee e^{C+y}]}\frac{\log(p)}{p\log(m)}+z^{-1}\sum_{p\in\mathcal{P}\cap[1,C]}\frac{\log(p)}{p\log(m)}.

Mertens’ approximation (4.5), yields the existence of a possibly different constant K>0K>0, such that

T1,1\displaystyle T_{1,1} K(1+z1)/log(m)+p𝒫[m]|θp,m|.\displaystyle\leq K(1+z^{-1})/\log(m)+\sum_{p\in\mathcal{P}\cap[m]}|\theta_{p,m}|.

For handling the term θp,m\theta_{p,m}, we use the mean value theorem, as well as the fact that f1(x)=zx2f_{1}^{\prime}(x)=-zx^{-2} for xzx\geq z to write

θp,m\displaystyle\theta_{p,m} C1log(p)log(m)𝔼[z(Sm+log(p)λm)3𝟙{zSm+log(p)/λm}][UIp]\displaystyle\leq C\frac{1}{\log(p)\log(m)}\mathbb{E}[z(S_{m}+\frac{\log(p)}{\lambda_{m}})^{-3}\mathbbm{1}_{\{z\leq S_{m}+\log(p)/\lambda_{m}\}}]\mathbb{P}[U\in I_{p}]
z2Cp+(log(m))2.\displaystyle\leq z^{-2}\frac{C}{p_{+}(\log(m))^{2}}.

Bounding p+p_{+} by pp from below, adding over p𝒫p\in\mathcal{P} and using Mertens’ second formula (4.6), we deduce the existence of a constant C>0C>0 such that

T1,1\displaystyle T_{1,1} C(1+1/z2)1log(m).\displaystyle\leq C(1+1/z^{2})\frac{1}{\log(m)}.

For handling the term T1,2T_{1,2}, we use the non-increasing property of u+u_{+} to write

|T1,2|p𝒫[m]log(p)pλm𝔼[(u+(Sm+log(p)/λm)u+(Sm+log(p)/λm+log(p+)/(p+λm)))].\displaystyle|T_{1,2}|\leq\sum_{p\in\mathcal{P}\cap[m]}\frac{\log(p)}{p\lambda_{m}}\mathbb{E}[(u_{+}(S_{m}+\log(p)/\lambda_{m})-u_{+}(S_{m}+\log(p)/\lambda_{m}+\log(p_{+})/(p_{+}\lambda_{m})))].

Applying summation by parts, u+0u_{+}\geq 0 and u+1\|u_{+}\|_{\infty}\leq 1, we thus get

T1,2log(pπ(m))pπ(m)λm+1λmp𝒫m(log(p)plog(p+)p+).\displaystyle T_{1,2}\leq\frac{\log(p_{\pi(m)})}{p_{\pi(m)}\lambda_{m}}+\frac{1}{\lambda_{m}}\sum_{p\in\mathcal{P}_{m}}(\frac{\log(p)}{p}-\frac{\log(p_{+})}{p_{+}}).

The second sum is telescopic, with boundary terms bounded by a constant multiple of λm1\lambda_{m}^{-1}, which gives the existence of a constant C>0C>0 such that

T1,2\displaystyle T_{1,2} Cλm.\displaystyle\leq\frac{C}{\lambda_{m}}.

The term T1,3T_{1,3} can be handled in a similar fashion.

In order to handle the term T2T_{2} we proceed as follows. Using the decomposition f=f1+u+uf^{\prime}=f_{1}^{\prime}+u_{+}-u_{-}, we obtain the bound

T2\displaystyle T_{2} 1λmq𝒫[m]log(q)q2(11/q)1𝔼[f1(Sm+log(q)λmξ~q+log(q)λm)|ξ~q1]+π23λm.\displaystyle\leq\frac{1}{\lambda_{m}}\sum_{q\in\mathcal{P}\cap[m]}\frac{\log(q)}{q^{2}}(1-1/q)^{-1}\mathbb{E}[f_{1}^{\prime}(S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}})\ |\ \tilde{\xi}_{q}\geq 1]+\frac{\pi^{2}}{3\lambda_{m}}.

The conditional expectation in the right can be estimated by conditioning over ξ~q\tilde{\xi}_{q} and applying Lemma 4.2. Indeed, using Lemma 4.2 with W=SmW=S_{m}, u=log(q)/λmξ~q+log(q)/λmu=\log(q)/\lambda_{m}\tilde{\xi}_{q}+\log(q)/\lambda_{m}, α=z\alpha=z and t=1t=1, we obtain

𝔼[f1(Sm+log(q)λmξ~q+log(q)λm)|ξ~q]=z𝔼[(Sm+log(q)λmξ~q+log(q)λm)2𝟙{Sm+log(q)λmξ~q+log(q)λm>z}|ξ~q]ζ(2)(C+2dK(Sm,D)((ξ~q+1)1λm/logq)).\mathbb{E}\left[f_{1}^{\prime}\left(S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}}\right)\ |\ \tilde{\xi}_{q}\right]\\ \begin{aligned} &=z\mathbb{E}\left[\left(S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}}\right)^{-2}\mathbbm{1}_{\{S_{m}+\frac{\log(q)}{\lambda_{m}}\tilde{\xi}_{q}+\frac{\log(q)}{\lambda_{m}}>z\}}\ |\ \tilde{\xi}_{q}\right]\\ &\leq\zeta(2)(C+2d_{K}(S_{m},D)((\tilde{\xi}_{q}+1)^{-1}\lambda_{m}/\log q)).\end{aligned}

Observe that

𝔼[(ξ~q+1)1]\displaystyle\mathbb{E}[(\tilde{\xi}_{q}+1)^{-1}] =k=1q(11/q)qk/k)=q(11/q)log(11/q)\displaystyle=\sum_{k=1}^{\infty}q(1-1/q)q^{-k}/k)=-q(1-1/q)\log(1-1/q)

which in turn implies the relation

𝔼[f1(Sm+log(q)λm(ξ~q+1)|ξ~q1]\displaystyle\mathbb{E}[f_{1}^{\prime}(S_{m}+\frac{\log(q)}{\lambda_{m}}(\tilde{\xi}_{q}+1)|\ \tilde{\xi}_{q}\geq 1] C2ζ(2)q(11/q)λmlog(11/q)dK(Sm,D)log(q).\displaystyle\leq C-2\zeta(2)q(1-1/q)\lambda_{m}\log(1-1/q)\frac{d_{K}(S_{m},D)}{\log(q)}.

Therefore, for a possibly different constant C>0C>0, the following inequality holds

T2\displaystyle T_{2} C/log(m)+π23dK(Sm,D)q𝒫(11/q)log(11/q)q2\displaystyle\leq C/\log(m)+\frac{\pi^{2}}{3}d_{K}(S_{m},D)\sum_{q\in\mathcal{P}}\frac{(1-1/q)\log(1-1/q)}{q^{2}}
C/log(m)+π215dK(Sm,D)\displaystyle\leq C/\log(m)+\frac{\pi^{2}}{15}d_{K}(S_{m},D)

guaranteeing the existence of a constant δ(0,1)\delta\in(0,1), such that

T2\displaystyle T_{2} dK(Sm,D)(1δ)+Clog(m).\displaystyle\leq d_{K}(S_{m},D)(1-\delta)+\frac{C}{\log(m)}.

By the above analysis,

dK(Sm,D)\displaystyle d_{K}(S_{m},D) C(1+z1)/log(m)+(1δ)dK(Sm,D).\displaystyle\leq C(1+z^{-1})/\log(m)+(1-\delta)d_{K}(S_{m},D).

Relation (4.5) follows from here.

Acknowledgements
We thank Louis Chen, Chinmoy Bhattacharjee, Ofir Gorodetsky and Larry Goldstein for helpful guidance in the elaboration of this paper. Arturo Jaramillo Gil was supported by the grant CBF2023-2024-2088.

References

  • [1] Chinmoy Bhattacharjee and Larry Goldstein. Dickman approximation in simulation, summations and perpetuities. Bernoulli, 25(4A):2758–2792, 2019.
  • [2] Chinmoy Bhattacharjee and Matthias Schulte. Dickman approximation of weighted random sums in the kolmogorov distance. Arxiv, 2022.
  • [3] Louis H. Y. Chen. Poisson approximation for dependent trials. Ann. Probability, 3(3):534–545, 1975.
  • [4] Louis H. Y. Chen, Arturo Jaramillo, and Xiaochuan Yang. A probabilistic approach to the erdös-kac theorem for additive functions. Arxiv, 2021.
  • [5] N. G. de Bruijn. On the number of positive integers x\leq x and free of prime factors >y>y. Nederl. Akad. Wetensch. Proc. Ser. A, 54:50–60, 1951.
  • [6] N. G. de Bruijn and J. H. van Lint. Incomplete sums of multiplicative functions. I. II. Indag. Math., 26:339–347, 348–359, 1964. Nederl. Akad. Wetensch. Proc. Ser. A 67.
  • [7] Karl Dickman. On the frequency of numbers containing prime factors of a certain relative magnitude. Arkiv för Matematik, Astronomi och Fysik., 22A(10):1–14, 1971.
  • [8] Christian Döbler. Stein’s method of exchangeable pairs for the beta distribution and generalizations. Electron. J. Probab., 20:no. 109, 34, 2015.
  • [9] Christian Döbler and Giovanni Peccati. The gamma Stein equation and noncentral de Jong theorems. Bernoulli, 24(4B):3384–3421, 2018.
  • [10] Andrew Granville. Smooth numbers: computational number theory and beyond. In Algorithmic number theory: lattices, number fields, curves and cryptography, volume 44 of Math. Sci. Res. Inst. Publ., pages 267–323. Cambridge Univ. Press, Cambridge, 2008.
  • [11] Douglas Hensley. The number of positive integers x\leq x and free of prime factors >y>y. J. Number Theory, 21(3):286–298, 1985.
  • [12] Adolf Hildebrand. Integers free of large prime factors and the Riemann hypothesis. Mathematika, 31(2):258–271, 1984.
  • [13] Adolf Hildebrand. On the number of positive integers x\leq x and free of prime factors >y>y. J. Number Theory, 22(3):289–307, 1986.
  • [14] Adolf Hildebrand and Gérald Tenenbaum. Integers without large prime factors. J. Théor. Nombres Bordeaux, 5(2):411–484, 1993.
  • [15] Hugh L. Montgomery and Robert C. Vaughan. Multiplicative number theory. I. Classical theory, volume 97 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2007.
  • [16] Erol A. Peköz and Adrian Röllin. New rates for exponential approximation and the theorems of Rényi and Yaglom. Ann. Probab., 39(2):587–608, 2011.
  • [17] Ross G. Pinsky. A natural probabilistic model on the integers and its relation to Dickman-type distributions and Buchstab’s function. In Probability and analysis in interacting physical systems, volume 283 of Springer Proc. Math. Stat., pages 267–294. Springer, Cham, 2019.
  • [18] V. Ramaswami. On the number of positive integers less than xx and free of prime divisors greater than xcx^{c}. Bull. Amer. Math. Soc., 55:1122–1127, 1949.
  • [19] Joung Min Song. Sums of nonnegative multiplicative functions over integers without large prime factors. II. Acta Arith., 102(2):105–129, 2002.
  • [20] Charles Stein. A bound for the error in the normal approximation to the distribution of a sum of dependent random variables. In Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability (Univ. California, Berkeley, Calif., 1970/1971), Vol. II: Probability theory, pages 583–602. Univ. California Press, Berkeley, CA, 1972.
  • [21] Gérald Tenenbaum. Note on a paper by J. M. Song: “Sums of nonnegative multiplicative functions over integers without large prime factors. I” [Acta Arith. 97 (2001), no. 4, 329–351; MR1823551 (2002f:11130)]. Acta Arith., 97(4):353–360, 2001.
  • [22] Gérald Tenenbaum. Introduction to analytic and probabilistic number theory, volume 163 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, third edition, 2015.
  • [23] Tim Trudgian. Updating the error term in the prime number theorem. Ramanujan J., 39(2):225–234, 2016.