Approximation of smooth numbers for harmonic samples: a Stein method approach
Abstract.
We present a de Bruijn-type approximation for quantifying the content of -smooth numbers, derived from samples obtained through a probability measure over the set with point mass function at proportional to . 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 method1991 Mathematics Subject Classification:
60F05, 11K651. Introduction
Let denote the set of non-negative prime numbers and the integer interval . Define the function through
We say that a given natural number is -smooth, with , if . Namely, if the prime factors of are at most . The primary goal of this manuscript is to introduce a new probabilistic measurement for the content of the set of -smooth numbers bounded by , and find sharp approximations for it. Our principal contribution is an explicit estimation for , 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 -smooth numbers
Let denote the collection of -smooth numbers bounded by and its cardinality
(1.1) |
The set 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 , and serves as a sufficient criterion for validating the Riemann hypothesis.
Many pieces of work have been devoted to the study of the function . 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
where is a uniform random variable defined over a probability space . This perspective can be utilized as a pivot for proposing alternative ways for describing , based on modifications of the law of (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 . 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 .
Our modification on the law of is largely inspired by the recent paper [4], which introduced the idea of using the harmonic distribution in as a tool for understanding divisibility properties of uniform samples. More precisely, we let be a collection of random variables defined in , with probability mass at proportional to . We then consider the function
(1.2) |
as measurement of content for . The advantage of dealing with the law of instead of resides in the fact that the prime multiplicities of 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 , to approximating in sup-norm, the cumulative distribution of a random sum of the form
where the are independent geometric variables supported over , with success probability . 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 and stems from [4, Lemma 3.1], which asserts that the variable can be sharply approximated by a product of the form , where , conditional on the value of , is uniformly distributed in the set . The precise statement establishes that for all ,
(1.3) |
where denotes the distance in total variation between the probability distributions associated to the inputs. In particular, if it holds that for all strictly positive , the set , defined through (2.3), is such that probability can be sharply approximated, uniformly over , 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 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 is a fundamental first step for this task. Indeed, in virtue of (1.3), sharp estimations of can be obtained from their counterpart for . In order to verify this, we write
Splitting the event into its components induced by the partition
and using the fact that for all , it holds that
where denotes the prime-counting function
we deduce the inequality
By a conditioning argument, the above identity yields
From the above discussion, we conclude that a sharp approximation for the distribution function of (or equivalently, for the function ) can be used to transfer approximations of
(1.4) |
to approximations for the distribution of , and ultimately, to . 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
(1.5) |
holds for (see also [4, Section 3]), suggesting that an estimation of (1.4) could be obtained through an analysis of
(1.6) |
The above term can be approximated, contingent on sharp estimates for
for
Approximations on this object are closely related to the description of the limiting behavior of the measure , obtained as the distribution of , conditional to . We conjecture that the measure 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
Now we turn our attention to the function , which is the main focus of our investigation. Similarly to , the description of the asymptotic behavior of turns out to be closely related to the so-called Dickman function , defined as the unique solution to the initial value problem of the delay type
(1.7) |
However, the structure of dependence on and through for the approximation of , is different from its counterpart for . In particular, our main result, presented in Theorem 3.1, combined with Theorem 2.2, imply that the quotient
converges to one as the quotient tends to infinity. In particular, is not asymptotically equivalent to as tends to infinity. A second remarkable feature of is the fact that it can be sharply estimated uniformly over , unlike , whose approximation has shown to be tractable only for regimes of and for which , with being an adequate function satisfying .
The approach we follow for studying is entirely probabilistic, and up to some important but tractable technical computations, consists of two fundamental steps: (i) First, we utilize a representation of in terms of independent geometric random variables conditioned on an adequate event. This reduces the analysis of , to the estimation of the distribution of the sum of weighted geometric random variables satisfying that converges pointwise as 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 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 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
In this section, we revise some of the known results regarding the study of approximations for . The material of this section is largely taken from [14].
2.1. The function and its relation to the Dickman function
In the seminal paper [7], Dickman presented a groundbreaking result regarding the behavior of the function and its relation to the solution to the delay equation (1.7). Dickman showed that for every the following convergence holds
This relation was further refined by Ramaswami in [18], who proved that for any fixed , there exists , such that
(2.1) |
These two results where subsequently explored thoroughly by many authors, seeking for extensions that allow for uniformity over . To pin down precisely the aforementioned notion of uniformity, we observe that if we set the smoothness threshold appearing in (2.1), to be equal to , we have that , where is given by
so that the Dickman approximation (2.1), reads , provided that we take . Ramaswani’s approximation, on the other hand, becomes
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 , define the set
Then, there exists a constant only depending on , such that
(2.2) |
where
Otherwise said, it holds that
The mathematical community has put many efforts in finding sharp upper bounds asymptotically equivalent to (2.2), uniformly over a larger set than . 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
holds uniformly over considerably larger set than , thus reducing the problem of estimating the exact value of , 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 , there exits a constant , such that the estimation (2.2) holds uniformly over , with
Remark 2.1.
The function appearing in the threshold is asymptotically smaller than any function of the form , for every .
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 , we define the set
(2.3) |
Then, provided that exists a mapping satisfying
(2.4) |
then the Riemann hypothesis holds.
Remark 2.2.
The condition (2.4) can be rephrased in the ”big ” notation as
Although we do not treat directly with the function , the above results will serve as benchmarks for setting up appropriate parallelisms for .
3. Main results
This section describes our findings regarding the analysis of . Since all of the prime divisors of lie in , then for all , so the problem of estimating can be localized into the regime .
In the sequel, will denote the Harmonic partial sum in and the product of the prime numbers belonging to , namely,
(3.1) |
The next result is the main contribution of our paper
Theorem 3.1.
There exists a constant , such that for ,
(3.2) |
where denotes the integral operator
Otherwise said, the function satisfies
uniformly over .
A version of Theorem 3.1 was first addressed by Bruijn and Van Lint in [6] for the case where . Related work followed in the contributions of Song [19] and Tenenbaum [21], where was replaced by a more general weight . 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 in terms of a sequence of independent geometric random variables , with satisfying
(3.3) |
for . As explained in detail in Section 5, a stochastic representation of the multiplicities of in terms of the , will allow us to reduce the proof of Theorem 4.1, to showing that the variables
(3.4) |
satisfy
(3.5) |
where and is a universal constant independent of and . 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 (iii) Some elementary number theory estimations by Mertens and (iv) Basic formulas regarding bias transforms.
4.1. Dickman distributional approximations
In the sequel, will denote the Euler-Mascheroni constant. Useful information on can be obtained from a probabilistic perspective by regarding the mapping as the density of a random variable. The verification of the condition can be obtained from an adequate analysis of the Laplace transform of , as explained in [14, Lemma 2.6.] (see also [15]). We say that a random variable is distributed according to the Dickman law if its cumulative distribution function is given by , where denotes the integral of over . 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 , it holds that
(4.1) |
This identity can be easily obtained from the fact that if is distributed according to and is a uniform random variable independent of , then
(4.2) |
The reader is referred to [17] for a proof of (4.2). The study of distributional approximations of random variables by means of sharp estimations on , for 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 in (4.1) by , one obtains the following characterization, which can be found in [1, Lemma 3.2]
Lemma 4.1.
The variable follows a Dickman law if and only if for every Lipchitz function , the following identity holds
Let denote the collection of test functions
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 for by an approximation of the type
for belonging to a family of test functions determined by . This idea can be formalized by considering the Stein equation
(4.3) |
where . Evaluation at equal to the variable of interest, taking expectation and then supremum with respect to , 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 be a random variable following the Dickman law. There exists a solution to the equation (4.3), admitting a decomposition of the type , where
(4.4) |
and satisfies , for functions non-negative and non-increasing, with
Moreover, for every random variable and constants , and ,
for some constant , where 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 and , 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
(4.5) |
The second Merten’s formula reads
(4.6) |
where is a universal constant, while the third guarantees the existence of a constant , such that
(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 be a sequence of independent random variables with distribution given as in (3.3). For , let denote the largest satisfying . Therefore is uniquely determined by ’s through the prime factorisation . Recall the definitions of and , given by (3.1). The following result, first presented in [4], describes the join distribution of the ’s.
Proposition 4.1.
The law of can be represented in terms of the ’s as
where
(4.8) |
Moreover, if , then
(4.9) |
The next result gives some useful identities for exploiting the distributional properties of the ’s
Lemma 4.3.
Let be a geometric random variable, with distribution
Then, for every compactly supported function ,
where is an independent copy of
Proof.
It suffices to consider the case for some . For such instance, we have that
The result follows from here. ∎
5. Proof of Theorem 3.1
In the sequel, will denote a sequence of random variables defined over , with harmonically distributed over , namely, for and zero otherwise, where 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.
This way, the problem reduces to the analysis of two pieces: the estimation of the term , 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 , such that
for some universal constant .
Contingent on the validity of these lemmas, we can prove Theorem 3.1.
Proof of Theorem 3.1.
Firstly, by Lemma 5.1,
This relation, together with Merten’s lemma and the approximation
leads to the inequality
By Merten’s approximation (4.7), we then conclude that
(5.1) |
Thus, by Lemma 5.2,
(5.2) |
for some universal constant . By the mean value theorem and the boundedness of , we deduce the existence of a constant such that
An application of Merten’s approximation (4.5), then yields
Combining this with inequality (5.2) then gives
Theorem 3.1 follows from here. ∎
6. Proof of Lemma 5.1
7. Proof of Lemma 5.2
Let a Dickman distributed random variable. Let be uniform in independent of . Let be the solution to the Stein equation (4.3). For convenience in the notation, we will simply write instead of . Define
with
and observe that
By Lemma 4.3,
where is an independent copy of . By conditioning the expectation in the right hand side on the value of , we obtain
where
Let be a random variable independent of the ’s, with distribution given by
(7.1) |
for and zero otherwise. With this notation in hand, we can write
By Lemma 4.2, , where is given by (4.4) and are non-negative and non-increasing, with . We can thus write
where
(7.2) |
Let be the distribution function of and its left inverse. Since is distributed according to , it follows that is a coupling for and , which allows us to write instead of in (7). Observe that the function remains constant equal to in the intervals , where
(7.3) |
for , defined as the smallest prime strictly bigger than . Observe that and also depend on and , but we have avoided this in the notation for convenience. By the Mertens’ approximation (4.5), there exists a constant such that
(7.4) |
Localizing and over the intervals of constancy of , we can write the terms as
In order to handle the term , we proceed as follows. Let denote the term
so that can be written as
Since is supported in and for , by distinguishing the instances for which is bigger than or smaller than or , we can write , where
By (7.4), there exists such that
Consequently, in the event , we have that
Using the fact that satisfies and for , we can easily show that
We will assume without loss of generality that 1. From here we deduce that
Splitting the sum in the right-hand side into the regimes and , we deduce that
Mertens’ approximation (4.5), yields the existence of a possibly different constant , such that
For handling the term , we use the mean value theorem, as well as the fact that for to write
Bounding by from below, adding over and using Mertens’ second formula (4.6), we deduce the existence of a constant such that
For handling the term , we use the non-increasing property of to write
Applying summation by parts, and , we thus get
The second sum is telescopic, with boundary terms bounded by a constant multiple of , which gives the existence of a constant such that
The term can be handled in a similar fashion.
In order to handle the term we proceed as follows. Using the decomposition , we obtain the bound
The conditional expectation in the right can be estimated by conditioning over and applying Lemma 4.2. Indeed, using Lemma 4.2 with , , and , we obtain
Observe that
which in turn implies the relation
Therefore, for a possibly different constant , the following inequality holds
guaranteeing the existence of a constant , such that
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 and free of prime factors . 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 and free of prime factors . 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 and free of prime factors . 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 and free of prime divisors greater than . 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.