You are currently browsing the monthly archive for November 2017.
I have just uploaded to the arXiv the paper “An inverse theorem for an inequality of Kneser“, submitted to a special issue of the Proceedings of the Steklov Institute of Mathematics in honour of Sergei Konyagin. It concerns an inequality of Kneser discussed previously in this blog, namely that
whenever are compact non-empty subsets of a compact connected additive group
with probability Haar measure
. (A later result of Kemperman extended this inequality to the nonabelian case.) This inequality is non-trivial in the regime
The connectedness of is essential, otherwise one could form counterexamples involving proper subgroups of
of positive measure. In the blog post, I indicated how this inequality (together with a more “robust” strengthening of it) could be deduced from submodularity inequalities such as
which in turn easily follows from the identity and the inclusion
, combined with the inclusion-exclusion formula.
In the non-trivial regime (2), equality can be attained in (1), for instance by taking to be the unit circle
and
to be arcs in that circle (obeying (2)). A bit more generally, if
is an arbitrary connected compact abelian group and
is a non-trivial character (i.e., a continuous homomorphism), then
must be surjective (as
has no non-trivial connected subgroups), and one can take
and
for some arcs
in that circle (again choosing the measures of these arcs to obey (2)). The main result of this paper is an inverse theorem that asserts that this is the only way in which equality can occur in (1) (assuming (2)); furthermore, if (1) is close to being satisfied with equality and (2) holds, then
must be close (in measure) to an example of the above form
. Actually, for technical reasons (and for the applications we have in mind), it is important to establish an inverse theorem not just for (1), but for the more robust version mentioned earlier (in which the sumset
is replaced by the partial sumset
consisting of “popular” sums).
Roughly speaking, the idea is as follows. Let us informally call a critical pair if (2) holds and the inequality (1) (or more precisely, a robust version of this inequality) is almost obeyed with equality. The notion of a critical pair obeys some useful closure properties. Firstly, it is symmetric in
, and invariant with respect to translation of either
or
. Furthermore, from the submodularity inequality (3), one can show that if
and
are critical pairs (with
and
positive), then
and
are also critical pairs. (Note that this is consistent with the claim that critical pairs only occur when
come from arcs of a circle.) Similarly, from associativity
, one can show that if
and
are critical pairs, then so are
and
.
One can combine these closure properties to obtain further ones. For instance, suppose is such that
. Then (cheating a little bit), one can show that
is also a critical pair, basically because
is the union of the
,
, the
are all critical pairs, and the
all intersect each other. This argument doesn’t quite work as stated because one has to apply the closure property under union an uncountable number of times, but it turns out that if one works with the robust version of sumsets and uses a random sampling argument to approximate
by the union of finitely many of the
, then the argument can be made to work.
Using all of these closure properties, it turns out that one can start with an arbitrary critical pair and end up with a small set
such that
and
are also critical pairs for all
(say), where
is the
-fold sumset of
. (Intuitively, if
are thought of as secretly coming from the pullback of arcs
by some character
, then
should be the pullback of a much shorter arc by the same character.) In particular,
exhibits linear growth, in that
for all
. One can now use standard technology from inverse sumset theory to show first that
has a very large Fourier coefficient (and thus is biased with respect to some character
), and secondly that
is in fact almost of the form
for some arc
, from which it is not difficult to conclude similar statements for
and
and thus finish the proof of the inverse theorem.
In order to make the above argument rigorous, one has to be more precise about what the modifier “almost” means in the definition of a critical pair. I chose to do this in the language of “cheap” nonstandard analysis (aka asymptotic analysis), as discussed in this previous blog post; one could also have used the full-strength version of nonstandard analysis, but this does not seem to convey any substantial advantages. (One can also work in a more traditional “non-asymptotic” framework, but this requires one to keep much more careful account of various small error terms and leads to a messier argument.)
[Update, Nov 15: Corrected the attribution of the inequality (1) to Kneser instead of Kemperman. Thanks to John Griesmer for pointing out the error.]
A basic object of study in multiplicative number theory are the arithmetic functions: functions from the natural numbers to the complex numbers. Some fundamental examples of such functions include
- The constant function
;
- The Kronecker delta function
;
- The natural logarithm function
;
- The divisor function
;
- The von Mangoldt function
, with
defined to equal
when
is a power
of a prime
for some
, and defined to equal zero otherwise; and
- The Möbius function
, with
defined to equal
when
is the product of
distinct primes, and defined to equal zero otherwise.
Given an arithmetic function , we are often interested in statistics such as the summatory function
the logarithmically (or harmonically) weighted summatory function
or the Dirichlet series
In the latter case, one typically has to first restrict to those complex numbers whose real part is large enough in order to ensure the series on the right converges; but in many important cases, one can then extend the Dirichlet series to almost all of the complex plane by analytic continuation. One is also interested in correlations involving additive shifts, such as
, but these are significantly more difficult to study and cannot be easily estimated by the methods of classical multiplicative number theory.
A key operation on arithmetic functions is that of Dirichlet convolution, which when given two arithmetic functions , forms a new arithmetic function
, defined by the formula
Thus for instance ,
,
, and
for any arithmetic function
. Dirichlet convolution and Dirichlet series are related by the fundamental formula
at least when the real part of is large enough that all sums involved become absolutely convergent (but in practice one can use analytic continuation to extend this identity to most of the complex plane). There is also the identity
at least when the real part of is large enough to justify interchange of differentiation and summation. As a consequence, many Dirichlet series can be expressed in terms of the Riemann zeta function
, thus for instance
Much of the difficulty of multiplicative number theory can be traced back to the discrete nature of the natural numbers , which form a rather complicated abelian semigroup with respect to multiplication (in particular the set of generators is the set of prime numbers). One can obtain a simpler analogue of the subject by working instead with the half-infinite interval
, which is a much simpler abelian semigroup under multiplication (being a one-dimensional Lie semigroup). (I will think of this as a sort of “completion” of
at the infinite place
, hence the terminology.) Accordingly, let us define a continuous arithmetic function to be a locally integrable function
. The analogue of the summatory function (1) is then an integral
and similarly the analogue of (2) is
The analogue of the Dirichlet series is the Mellin-type transform
which will be well-defined at least if the real part of is large enough and if the continuous arithmetic function
does not grow too quickly, and hopefully will also be defined elsewhere in the complex plane by analytic continuation.
For instance, the continuous analogue of the discrete constant function would be the constant function
, which maps any
to
, and which we will denote by
in order to keep it distinct from
. The two functions
and
have approximately similar statistics; for instance one has
and
where is the
harmonic number, and we are deliberately vague as to what the symbol
means. Continuing this analogy, we would expect
which reflects the fact that has a simple pole at
with residue
, and no other poles. Note that the identity
is initially only valid in the region
, but clearly the right-hand side can be continued analytically to the entire complex plane except for the pole at
, and so one can define
in this region also.
In a similar vein, the logarithm function is approximately similar to the logarithm function
, giving for instance the crude form
of Stirling’s formula, or the Dirichlet series approximation
The continuous analogue of Dirichlet convolution is multiplicative convolution using the multiplicative Haar measure : given two continuous arithmetic functions
, one can define their convolution
by the formula
Thus for instance . A short computation using Fubini’s theorem shows the analogue
of (3) whenever the real part of is large enough that Fubini’s theorem can be justified; similarly, differentiation under the integral sign shows that
again assuming that the real part of is large enough that differentiation under the integral sign (or some other tool like this, such as the Cauchy integral formula for derivatives) can be justified.
Direct calculation shows that for any complex number , one has
(at least for the real part of large enough), and hence by several applications of (5)
for any natural number . This can lead to the following heuristic: if a Dirichlet series
behaves like a linear combination of poles
, in that
for some set of poles and some coefficients
and natural numbers
(where we again are vague as to what
means, and how to interpret the sum
if the set of poles is infinite), then one should expect the arithmetic function
to behave like the continuous arithmetic function
In particular, if we only have simple poles,
then we expect to have behave like continuous arithmetic function
Integrating this from to
, this heuristically suggests an approximation
for the summatory function, and similarly
with the convention that is
when
, and similarly
is
when
. One can make these sorts of approximations more rigorous by means of Perron’s formula (or one of its variants) combined with the residue theorem, provided that one has good enough control on the relevant Dirichlet series, but we will not pursue these rigorous calculations here. (But see for instance this previous blog post for some examples.)
For instance, using the more refined approximation
to the zeta function near , we have
we would expect that
and thus for instance
which matches what one actually gets from the Dirichlet hyperbola method (see e.g. equation (44) of this previous post).
Or, noting that has a simple pole at
and assuming simple zeroes elsewhere, the log derivative
will have simple poles of residue
at
and
at all the zeroes, leading to the heuristic
suggesting that should behave like the continuous arithmetic function
leading for instance to the summatory approximation
which is a heuristic form of the Riemann-von Mangoldt explicit formula (see Exercise 45 of these notes for a rigorous version of this formula).
Exercise 1 Go through some of the other explicit formulae listed at this Wikipedia page and give heuristic justifications for them (up to some lower order terms) by similar calculations to those given above.
Given the “adelic” perspective on number theory, I wonder if there are also -adic analogues of arithmetic functions to which a similar set of heuristics can be applied, perhaps to study sums such as
. A key problem here is that there does not seem to be any good interpretation of the expression
when
is complex and
is a
-adic number, so it is not clear that one can analyse a Dirichlet series
-adically. For similar reasons, we don’t have a canonical way to define
for a Dirichlet character
(unless its conductor happens to be a power of
), so there doesn’t seem to be much to say in the
-aspect either.
Alice Guionnet, Assaf Naor, Gilles Pisier, Sorin Popa, Dimitri Shylakhtenko, and I are organising a three month program here at the Institute for Pure and Applied Mathematics (IPAM) on the topic of Quantitative Linear Algebra. The purpose of this program is to bring together mathematicians and computer scientists (both junior and senior) working in various quantitative aspects of linear operators, particularly in large finite dimension. Such aspects include, but are not restricted to discrepancy theory, spectral graph theory, random matrices, geometric group theory, ergodic theory, von Neumann algebras, as well as specific research directions such as the Kadison-Singer problem, the Connes embedding conjecture and the Grothendieck inequality. There will be several workshops and tutorials during the program (for instance I will be giving a series of introductory lectures on random matrix theory).
While we already have several confirmed participants, we are still accepting applications for this program until Dec 4; details of the application process may be found at this page.
Recent Comments