You are currently browsing the monthly archive for July 2024.

I’ve just uploaded to the arXiv my paper “Dense sets of natural numbers with unusually large least common multiples“. This short paper answers (in the negative) a somewhat obscure question of Erdös and Graham:

Problem 1 Is it true that if {A} is a set of natural numbers for which

\displaystyle  \frac{1}{\log\log x} \sum_{n \in A: n \leq x} \frac{1}{n} \ \ \ \ \ (1)

goes to infinity as {x \rightarrow \infty}, then the quantity

\displaystyle  \frac{1}{(\sum_{n \in A: n \leq x} \frac{1}{n})^2} \sum_{n,m \in A: n < m \leq x} \frac{1}{\mathrm{lcm}(n,m)} \ \ \ \ \ (2)

also goes to infinity as {x \rightarrow \infty}?

At first glance, this problem may seem rather arbitrary, but it can be motivated as follows. The hypothesis that (1) goes to infinity is a largeness condition on {A}; in view of Mertens’ theorem, it can be viewed as an assertion that {A} is denser than the set of primes. On the other hand, the conclusion that (2) grows is an assertion that {\frac{1}{\mathrm{lcm}(n,m)}} becomes significantly larger than {\frac{1}{nm}} on the average for large {n,m \in A}; that is to say, that many pairs of numbers in {A} share a common factor. Intuitively, the problem is then asking whether sets that are significantly denser than the primes must start having lots of common factors on average.

For sake of comparison, it is easy to see that if (1) goes to infinity, then at least one pair {(n,m)} of distinct elements in {A} must have a non-trivial common factor. For if this were not the case, then the elements of {A} are pairwise coprime, so each prime {p} has at most one multiple in {A}, and so can contribute at most {1/p} to the sum in (1), and hence by Mertens’ theorem, and the fact that every natural number greater than one is divisible by at least one prime {p}, the quantity (1) stays bounded, a contradiction.

It turns out, though, that the answer to the above problem is negative; one can find sets {A} that are denser than the primes, but for which (2) stays bounded, so that the least common multiples in the set are unusually large. It was a bit surprising to me that this question had not been resolved long ago (in fact, I was not able to find any prior literature on the problem beyond the original reference of Erdös and Graham); in contrast, another problem of Erdös and Graham concerning sets with unusually small least common multiples was extensively studied (and essentially solved) about twenty years ago, while the study of sets with unusually large greatest common divisor for many pairs in the set has recently become somewhat popular, due to their role in the proof of the Duffin-Schaeffer conjecture by Koukoulopoulos and Maynard.

To search for counterexamples, it is natural to look for numbers with relatively few prime factors, in order to reduce their common factors and increase their least common multiple. A particularly simple example, whose verification is on the level of an exercise in a graduate analytic number theory course, is the set of semiprimes (products of two primes), for which one can readily verify that (1) grows like {\log\log x} but (2) stays bounded. With a bit more effort, I was able to optimize the construction and uncover the true threshold for boundedness of (2), which was a little unexpected:

Theorem 2
  • (i) For any {C>0}, there exists a set of natural numbers {A} with

    \displaystyle  \sum_{n \in A: n \leq x} \frac{1}{n} = \exp( (C+o(1)) (\log\log x)^{1/2} \log\log\log x )

    for all large {x}, for which (2) stays bounded.
  • (ii) Conversely, if (2) stays bounded, then

    \displaystyle  \sum_{n \in A: n \leq x} \frac{1}{n} \ll \exp( O( (\log\log x)^{1/2} \log\log\log x ) )

    for all large {x}.

The proofs are not particularly long or deep, but I thought I would record here some of the process towards finding them. My first step was to try to simplify the condition that (2) stays bounded. In order to use probabilistic intuition, I first expressed this condition in probabilistic terms as

\displaystyle  \mathbb{E} \frac{\mathbf{n} \mathbf{m}}{\mathrm{lcm}(\mathbf{n}, \mathbf{m})} \ll 1

for large {x}, where {\mathbf{n}, \mathbf{m}} are independent random variables drawn from {\{ n \in A: n \leq x \}} with probability density function

\displaystyle  \mathbb{P} (\mathbf{n} = n) = \frac{1}{\sum_{m \in A: m \leq x} \frac{1}{m}} \frac{1}{n}.

The presence of the least common multiple in the denominator is annoying, but one can easily flip the expression to the greatest common divisor:

\displaystyle  \mathbb{E} \mathrm{gcd}(\mathbf{n}, \mathbf{m}) \ll 1.

If the expression {\mathrm{gcd}(\mathbf{n}, \mathbf{m})} was a product of a function of {\mathbf{n}} and a function of {\mathbf{m}}, then by independence this expectation would decouple into simpler averages involving just one random variable instead of two. Of course, the greatest common divisor is not of this form, but there is a standard trick in analytic number theory to decouple the greatest common divisor, namely to use the classic Gauss identity {n = \sum_{d|n} \varphi(d)}, with {\varphi} the Euler totient function, to write

\displaystyle  \mathrm{gcd}(\mathbf{n}, \mathbf{m}) = \sum_{d | \mathbf{n}, \mathbf{m}} \varphi(d).

Inserting this formula and interchanging the sum and expectation, we can now express the condition as bounding a sum of squares:

\displaystyle  \sum_d \varphi(d) \mathbb{P}(d|\mathbf{n})^2 \ll 1.

Thus, the condition (2) is really an assertion to the effect that typical elements of {A} do not have many divisors. From experience in sieve theory, the probabilities {\mathbb{P}(d|\mathbf{n})} tend to behave multiplicatively in {d}, so the expression here heuristically behaves like an Euler product that looks something like

\displaystyle  \prod_p (1 + \varphi(p) \mathbb{P}(p|\mathbf{n})^2)

and so the condition (2) is morally something like

\displaystyle  \sum_p p \mathbb{P}(p|\mathbf{n})^2 \ll 1. \ \ \ \ \ (3)

Comparing this with the Mertens’ theorems, this leads to the heuristic prediction that {\mathbb{P}(p|\mathbf{n})} (for a typical prie {p} much smaller than {x}) should decay somewhat like {\frac{1}{p (\log\log p)^{1/2}}} (ignoring for now factors of {\log\log\log p}). This can be compared to the example of the set of primes or semiprimes on one hand, where the probability is like {\frac{1}{p \log\log p}}, and the set of all natural numbers on the other hand, where the probability is like {\frac{1}{p}}. So the critical behavior should come from sets that are in some sense “halfway” between the primes and the natural numbers.

It is then natural to try a random construction, in which one sieves out the natural numbers by permitting each natural number {n} to survive with a probability resembling {\prod_{p|n} \frac{1}{(\log\log p)^{1/2}}}, in order to get the predicted behavior for {\mathbb{P}(p|\mathbf{n})}. Performing some standard calculations, this construction could ensure (2) bounded with a density a little bit less than the one stated in the main theorem; after optimizing the parameters, I could only get something like

\displaystyle  \sum_{n \in A: n \leq x} \frac{1}{n} = \exp( (\log\log x)^{1/2} (\log\log\log x)^{-1/2-o(1)} ).

I was stuck on optimising the construction further, so I turned my attention to a positive result in the spirit of (ii) of the main theorem. On playing around with (3), I observed that one could use Cauchy-Schwarz and Mertens’ theorem to obtain the bound

\displaystyle  \sum_{p \leq x} \mathbb{P}(p|\mathbf{n}) \ll (\log\log x)^{1/2}

which was in line with the previous heuristic that {\mathbb{P}(p|\mathbf{n})} should behave like {\frac{1}{p (\log\log p)^{1/2}}}. The left-hand side had a simple interpretation: by linearity of expectation, it was the expected number {\mathbb{E} \omega(\mathbf{n})} of prime factors of {\mathbf{n}}. So the boundedness of (2) implied that a typical element of {A} only had about {(\log\log x)^{1/2}} prime factors, in contrast to the {\log\log x} predicted by the Hardy-Ramanujan law. Standard methods from the anatomy of integers can then be used to see how dense a set with that many prime factors could be, and this soon led to a short proof of part (ii) of the main theorem (I eventually found for instance that Jensen’s inequality could be used to create a particularly slick argument).

It then remained to improve the lower bound construction to eliminate the {\log\log\log x} losses in the exponents. By deconstructing the proof of the upper bound, it became natural to consider something like the set of natural numbers {n} that had at most {(\log\log n)^{1/2}} prime factors. This construction actually worked for some scales {x} – namely those {x} for which {(\log\log x)^{1/2}} was a natural number – but there was some strange “discontinuities” in the analysis that prevented me from establishing the boundedness of (2) for arbitrary scales {x}. The basic problem was that increasing the number of permitted prime factors from one natural number threshold {k} to another {k+1} ended up increasing the density of the set by an unbounded factor (of the order of {k}, in practice), which heavily disrupted the task of trying to keep the ratio (2) bounded. Usually the resolution to these sorts of discontinuities is to use some sort of random “average” of two or more deterministic constructions – for instance, by taking some random union of some numbers with {k} prime factors and some numbers with {k+1} prime factors – but the numerology turned out to be somewhat unfavorable, allowing for some improvement in the lower bounds over my previous construction, but not enough to close the gap entirely. It was only after substantial trial and error that I was able to find a working deterministic construction, where at a given scale one collected either numbers with at most {k} prime factors, or numbers with {k+1} prime factors but with the largest prime factor in a specific range, in which I could finally get the numerator and denominator in (2) to be in balance for every {x}. But once the construction was written down, the verification of the required properties ended up being quite routine.

Many modern mathematical proofs are a combination of conceptual arguments and technical calculations. There is something of a tradeoff between the two: one can add more conceptual arguments to try to reduce the technical computations, or vice versa. (Among other things, this leads to a Berkson paradox-like phenomenon in which a negative correlation can be observed between the two aspects of a proof; see this recent Mastodon post of mine for more discussion.)

In a recent article, Heather Macbeth argues that the preferred balance between conceptual and computational arguments is quite different for a computer-assisted proof than it is for a purely human-readable proof. In the latter, there is a strong incentive to minimize the amount of calculation to the point where it can be checked by hand, even if this requires a certain amount of ad hoc rearrangement of cases, unmotivated parameter selection, or otherwise non-conceptual additions to the arguments in order to reduce the calculation. But in the former, once one is willing to outsource any tedious verification or optimization task to a computer, the incentives are reversed: freed from the need to arrange the argument to reduce the amount of calculation, one can now describe an argument by listing the main ingredients and then letting the computer figure out a suitable way to combine them to give the stated result. The two approaches can thus be viewed as complementary ways to describe a result, with neither necessarily being superior to the other.

In this post, I would like to illustrate this computation-outsourced approach with the topic of zero-density theorems for the Riemann zeta function, in which all computer verifiable calculations (as well as other routine but tedious arguments) are performed “off-stage”, with the intent of focusing only on the conceptual inputs to these theorems.

Zero-density theorems concern upper bounds for the quantity {N(\sigma,T)} for a given {1/2 \leq \sigma \leq 1} and large {T}, which is defined as the number of zeroes of the Riemann zeta function in the rectangle {\{ \beta+i\gamma: \sigma \leq \beta \leq 1; 0 \leq \gamma \leq T \}}. (There is also an important generalization of this quantity to {L}-functions, but for simplicity we will focus on the classical zeta function case here). Such quantities are important in analytic number theory for many reasons, one of which is through explicit formulae such as the Riemann-von Mangoldt explicit formula

\displaystyle  \sum_{n \leq x}^{\prime} \Lambda(n) = x - \sum_{\rho:\zeta(\rho)=0} \frac{x^\rho}{\rho} - \log(2\pi) - \frac{1}{2} \log(1-x^{-2}) \ \ \ \ \ (1)

relating the prime numbers to the zeroes of the zeta function (the “music of the primes”). The better bounds one has on {N(\sigma,T)}, the more control one has on the complicated term {\sum_{\rho:\zeta(\rho)=0} \frac{x^\rho}{\rho}} on the right-hand side.

Clearly {N(\sigma,T)} is non-increasing in {\sigma}. The Riemann-von Mangoldt formula, together with the functional equation, gives us the asymptotic

\displaystyle  N(1/2,T) \asymp T \log T

in the {\sigma=1/2} case, while the prime number theorem tells us that

\displaystyle  N(1,T) = 0. \ \ \ \ \ (2)

The various zero free regions for the zeta function can be viewed as slight improvements to (2); for instance, the classical zero-free region is equivalent to the assertion that {N(\sigma,T)} vanishes if {\sigma > 1 - c/\log T} for some small absolute constant {c>0}, and the Riemann hypothesis is equivalent to the assertion that {N(\sigma,T)=0} for all {\sigma>1/2}.

Experience has shown that the most important quantity to control here is the exponent {A(\sigma)}, defined as the least constant for which one has an asymptotic

\displaystyle  N(\sigma,T) = T^{A(\sigma)(1-\sigma)+o(1)}

as {T \rightarrow \infty}. Thus, for instance,

\displaystyle  A(1/2) = 2, \ \ \ \ \ (3)

{A(1) = 0}, and {A(\sigma)(1-\sigma)} is a non-increasing function of {\sigma}, so we obtain the trivial “von Mangoldt” zero density theorem

\displaystyle  A(\sigma) \leq \frac{1}{1-\sigma}.

Of particular interest is the supremal value {\|A\|_\infty := \sup_{1/2 \leq \sigma \leq 1} A(\sigma)} of {A}, which has to be at least {2} thanks to (3). The density hypothesis asserts that the maximum is in fact exactly {2}, or equivalently that

\displaystyle  A(\sigma) \leq 2, \ \ \ \ \ (4)

for all {1/2 \leq \sigma \leq 1}. This is of course implied by the Riemann hypothesis (which clearly implies that {A(\sigma)=0} for all {\sigma>1/2}), but is a more tractable hypothesis to work with; for instance, the hypothesis is already known to hold for {\sigma \geq 25/32 = 0.78125} by the work of Bourgain (building upon many previous authors). The quantity {\|A\|_\infty} directly impacts our understanding of the prime number theorem in short intervals; indeed, it is not difficult using (1) (as well as the Vinogradov-Korobov zero-free region) to establish a short interval prime number theorem

\displaystyle  \sum_{x \leq n \leq x + x^\theta} \Lambda(n) = (1+o(1)) x^\theta

for all {x \rightarrow \infty} if {1 - \frac{1}{\|A\|_\infty} < \theta < 1} is a fixed exponent, or for almost all {x \rightarrow \infty} if {1 - \frac{2}{\|A\|_\infty} < \theta < 1} is a fixed exponent. Until recently, the best upper bound for {\|A\|_\infty} was {12/5 = 2.4}, thanks to a 1972 result of Huxley; but this was recently lowered to {30/13=2.307\ldots} in a breakthrough work of Guth and Maynard.

In between the papers of Huxley and Guth-Maynard are dozens of additional improvements on {A(\sigma)}, though it is only the Guth-Maynard paper that actually lowered the supremum norm {\|A\|_\infty}. A summary of most of the state of the art before Guth-Maynard may be found in Table 2 of this recent paper of Trudgian and Yang; it is complicated, but it is easy enough to get a computer to illustrate it with a plot:

(For an explanation of what is going on under the assumption of the Lindelöf hypothesis, see below the fold.) This plot represents the combined effort of nearly a dozen papers, each one of which claims one or more components of the depicted piecewise smooth curve, and is written in the “human-readable” style mentioned above, where the argument is arranged to reduce the amount of tedious computation to human-verifiable levels, even if this comes the cost of obscuring the conceptual ideas. (For an animation of how this bound improved over time, see here.) Below the fold, I will try to describe (in sketch form) some of the standard ingredients that go into these papers, in particular the routine reduction of deriving zero density estimates from large value theorems for Dirichlet series. We will not attempt to rewrite the entire literature of zero-density estimates in this fashion, but focus on some illustrative special cases.

Read the rest of this entry »

The Salem prize was established in 1968 and named in honor of Raphaël Salem (1898-1963), a mathematician famous notably for his deep study of the links between Fourier series and number theory and for pioneering applications of probabilistic methods to these fields. It was not awarded from 2019-2022, due to both the COVID pandemic and the death of Jean Bourgain who had been almost single-handedly administering the prize, but is now active again, being administered by Akshay Ventakesh and the IAS. I chair the scientific committee for this prize, whose other members are Guy David and Mikhail Sodin. Last year, the prize was awarded to Sarah Peluse and Julian Sahasrabudhe.

Nominations for the 2024 Salem Prize are now open until September 1st. Nominations should include a CV of the nominee and a nomination letter explaining the significance of the nominee’s work. Supplementary documentation, such as supporting letters of recommendation or key publications, can additionally be provided, but are not required.

Nominees may be individuals from any country or institution. Preference will be given to nominees who have received their PhD in the last ten years, although this rule may be relaxed if there are mitigating personal circumstances, or if there have been few Salem prize winners in recent years.  Self-nominations will not be considered, nor are past Prize winners or Scientific Committee members eligible.

The prize does not come with a direct monetary award, but winners will be invited to visit the IAS and to give a lecture associated with the award of the prize.

See also the previous year’s announcement of the Salem prize nomination period.

Archives