A major topic of interest of analytic number theory is the asymptotic behaviour of the Riemann zeta function in the critical strip
in the limit
. For the purposes of this set of notes, it is a little simpler technically to work with the log-magnitude
of the zeta function. (In principle, one can reconstruct a branch of
, and hence
itself, from
using the Cauchy-Riemann equations, or tools such as the Borel-Carathéodory theorem, see Exercise 40 of Supplement 2.)
One has the classical estimate
(See e.g. Exercise 37 from Supplement 3.) In view of this, let us define the normalised log-magnitudes for any
by the formula
informally, this is a normalised window into near
. One can rephrase several assertions about the zeta function in terms of the asymptotic behaviour of
. For instance:
- (i) The bound (1) implies that
is asymptotically locally bounded from above in the limit
, thus for any compact set
we have
for
and
sufficiently large. In fact the implied constant in
only depends on the projection of
to the real axis.
- (ii) For
, we have the bounds
which implies that
converges locally uniformly as
to zero in the region
.
- (iii) The functional equation, together with the symmetry
, implies that
which by Exercise 17 of Supplement 3 shows that
as
, locally uniformly in
. In particular, when combined with the previous item, we see that
converges locally uniformly as
to
in the region
.
- (iv) From Jensen’s formula (Theorem 16 of Supplement 2) we see that
is a subharmonic function, and thus
is subharmonic as well. In particular we have the mean value inequality
for any disk
, where the integral is with respect to area measure. From this and (ii) we conclude that
for any disk with
and sufficiently large
; combining this with (i) we conclude that
is asymptotically locally bounded in
in the limit
, thus for any compact set
we have
for sufficiently large
.
From (iv) and the usual Arzela-Ascoli diagonalisation argument, we see that the are asymptotically compact in the topology of distributions: given any sequence
tending to
, one can extract a subsequence such that the
converge in the sense of distributions. Let us then define a normalised limit profile of
to be a distributional limit
of a sequence of
; they are analogous to limiting profiles in PDE, and also to the more recent introduction of “graphons” in the theory of graph limits. Then by taking limits in (i)-(iv) we can say a lot about such normalised limit profiles
(up to almost everywhere equivalence, which is an issue we will address shortly):
- (i)
is bounded from above in the critical strip
.
- (ii)
vanishes on
.
- (iii) We have the functional equation
for all
. In particular
for
.
- (iv)
is subharmonic.
Unfortunately, (i)-(iv) fail to characterise completely. For instance, one could have
for any convex function
of
that equals
for
,
for
, and obeys the functional equation
, and this would be consistent with (i)-(iv). One can also perturb such examples in a region where
is strictly convex to create further examples of functions obeying (i)-(iv). Note from subharmonicity that the function
is always going to be convex in
; this can be seen as a limiting case of the Hadamard three-lines theorem (Exercise 41 of Supplement 2).
We pause to address one minor technicality. We have defined as a distributional limit, and as such it is a priori only defined up to almost everywhere equivalence. However, due to subharmonicity, there is a unique upper semi-continuous representative of
(taking values in
), defined by the formula
for any (note from subharmonicity that the expression in the limit is monotone nonincreasing as
, and is also continuous in
). We will now view this upper semi-continuous representative of
as the canonical representative of
, so that
is now defined everywhere, rather than up to almost everywhere equivalence.
By a classical theorem of Riesz, a function is subharmonic if and only if the distribution
is a non-negative measure, where
is the Laplacian in the
coordinates. Jensen’s formula (or Greens’ theorem), when interpreted distributionally, tells us that
away from the real axis, where ranges over the non-trivial zeroes of
. Thus, if
is a normalised limit profile for
that is the distributional limit of
, then we have
where is a non-negative measure which is the limit in the vague topology of the measures
Thus is a normalised limit profile of the zeroes of the Riemann zeta function.
Using this machinery, we can recover many classical theorems about the Riemann zeta function by “soft” arguments that do not require extensive calculation. Here are some examples:
Theorem 1 The Riemann hypothesis implies the Lindelöf hypothesis.
Proof: It suffices to show that any limiting profile (arising as the limit of some
) vanishes on the critical line
. But if the Riemann hypothesis holds, then the measures
are supported on the critical line
, so the normalised limit profile
is also supported on this line. This implies that
is harmonic outside of the critical line. By (ii) and unique continuation for harmonic functions, this implies that
vanishes on the half-space
(and equals
on the complementary half-space, by (iii)), giving the claim.
In fact, we have the following sharper statement:
Theorem 2 (Backlund) The Lindelöf hypothesis is equivalent to the assertion that for any fixed
, the number of zeroes in the region
is
as
.
Proof: If the latter claim holds, then for any , the measures
assign a mass of
to any region of the form
as
for any fixed
and
. Thus the normalised limiting profile measure
is supported on the critical line, and we can repeat the previous argument.
Conversely, suppose the claim fails, then we can find a sequence and
such that
assigns a mass of
to the region
. Extracting a normalised limiting profile, we conclude that the normalised limiting profile measure
is non-trivial somewhere to the right of the critical line, so the associated subharmonic function
is not harmonic everywhere to the right of the critical line. From the maximum principle and (ii) this implies that
has to be positive somewhere on the critical line, but this contradicts the Lindelöf hypothesis. (One has to take a bit of care in the last step since
only converges to
in the sense of distributions, but it turns out that the subharmonicity of all the functions involved gives enough regularity to justify the argument; we omit the details here.)
Theorem 3 (Littlewood) Assume the Lindelöf hypothesis. Then for any fixed
, the number of zeroes in the region
is
as
.
Proof: By the previous arguments, the only possible normalised limiting profile for is
. Taking distributional Laplacians, we see that the only possible normalised limiting profile for the zeroes is Lebesgue measure on the critical line. Thus,
can only converge to
as
, and the claim follows.
Even without the Lindelöf hypothesis, we have the following result:
Theorem 4 (Titchmarsh) For any fixed
, there are
zeroes in the region
for sufficiently large
.
Among other things, this theorem recovers a classical result of Littlewood that the gaps between the imaginary parts of the zeroes goes to zero, even without assuming unproven conjectures such as the Riemann or Lindelöf hypotheses.
Proof: Suppose for contradiction that this were not the case, then we can find and a sequence
such that
contains
zeroes. Passing to a subsequence to extract a limit profile, we conclude that the normalised limit profile measure
assigns no mass to the horizontal strip
. Thus the associated subharmonic function
is actually harmonic on this strip. But by (ii) and unique continuation this forces
to vanish on this strip, contradicting the functional equation (iii).
Exercise 5 Use limiting profiles to obtain the matching upper bound of
for the number of zeroes in
for sufficiently large
.
Remark 6 One can remove the need to take limiting profiles in the above arguments if one can come up with quantitative (or “hard”) substitutes for qualitative (or “soft”) results such as the unique continuation property for harmonic functions. This would also allow one to replace the qualitative decay rates
with more quantitative decay rates such as
or
. Indeed, the classical proofs of the above theorems come with quantitative bounds that are typically of this form (see e.g. the text of Titchmarsh for details).
Exercise 7 Let
denote the quantity
, where the branch of the argument is taken by using a line segment connecting
to (say)
, and then to
. If we have a sequence
producing normalised limit profiles
for
and the zeroes respectively, show that
converges in the sense of distributions to the function
, or equivalently
Conclude in particular that if the Lindelöf hypothesis holds, then
as
.
A little bit more about the normalised limit profiles are known unconditionally, beyond (i)-(iv). For instance, from Exercise 3 of Notes 5 we have
as
, which implies that any normalised limit profile
for
is bounded by
on the critical line, beating the bound of
coming from convexity and (ii), (iii), and then convexity can be used to further bound
away from the critical line also. Some further small improvements of this type are known (coming from various methods for estimating exponential sums), though they fall well short of determining
completely at our current level of understanding. Of course, given that we believe the Riemann hypothesis (and hence the Lindelöf hypothesis) to be true, the only actual limit profile that should exist is
(in fact this assertion is equivalent to the Lindelöf hypothesis, by the arguments above).
Better control on limiting profiles is available if we do not insist on controlling for all values of the height parameter
, but only for most such values, thanks to the existence of several mean value theorems for the zeta function, as discussed in Notes 6; we discuss this below the fold.
— 1. Limiting profiles outside of exceptional sets —
In order to avoid an excessive number of extraction of subsequences and discarding of exceptional sets, we now move away from the standard sequential notion of a limit, and instead work with the less popular, but equally valid notion of an ultrafilter limit. Recall that an ultrafilter on a set
is a collection of subsets of
(which we will call the “
-large” sets) which are the sets of full measure with regards to some finitely additive
-valued probability measure on
(with the power set Boolean algebra
). We call a subset of
-small if it is not
-large. Given a function
into a topological space
and a point
, we say that
converges to
along
if
is
-large for every neighbourhood
of
, and then we call
a
-limit of
.
Exercise 8 Let
be a function into a topological space
, and let
be an ultrafilter on
.
- (i) If
is compact, show that
has at least one
-limit.
- (ii) If
is Hausdorff, show that
has at most one
-limit.
- (iii) Conversely, if
fails to be compact (resp. Hausdorff), show that there exists a function
and an ultrafilter
on
such that
has no
-limit (resp. more than one
-limit).
In particular, given an ultrafilter on the non-negative reals
, which is non-principal in the sense that all compact subsets of
are
-small,, there exists a unique normalised limiting profile
that is the limit of
along
, and similarly for
. Because the distributional topology is second countable, such limiting profiles are also limiting profiles of sequences
as in the previous discussion, and so we retain all existing properties of limit profiles such as (i)-(iv). However, in the ultrafilter formalism we can now easily avoid various “small” exceptional sets of
, in addition to the compact sets that have already been excluded. For instance, let us call an ultrafilter
generic if every Lebesgue measurable subset
of
of zero upper density (thus
has Lebesgue measure
as
) is
-small. The existence of generic ultrafilters follows easily from Zorn’s lemma. Define a generic limit profile to be a limit profile that arises from a generic ultrafilter; informally, these capture the possible behaviour of the zeta function outside of a set of heights
of zero density. To see how these profiles are better than arbitrary limit profiles, we recall from Exercise 2 of Notes 6 that
if the are
-separated elements of
and
are arbitrary complex coefficients. If we set
, we can conclude (among other things), that for any constant
, one has
for all outside of a set of measure
(informally: “square root cancellation occurs generically”). Using this, one can for instance show that
for all outside of a set of measure
, which implies that any generic limit profile
vanishes on the critical line, and thus must be
; that is to say, the Lindelöf hypothesis is true “generically”.
One can profitably explore the regime between arbitrary non-principal ultrafilters and generic ultrafilters by introducing the intermediate notion of an -generic ultrafilter for any
, defined as an ultrafilter
with the property that any Lebesgue measurable subset
of
of “dimension at most
” in the sense that
has measure
, is
-small. One can then interpret many existing mean value theorems on the zeta function (or on other Dirichlet series) as controlling the
-generic limit profiles of
, or more generally of the log-magnitude of various Dirichlet series (e.g.
for various exponents
). For instance, the previous argument shows that
for all outside of a set of measure
, which implies that any
-generic limit profile
is bounded above by
on the critical line. One can also recast much of the arguments in Notes 6 in this language (defining limit profiles for various Dirichlet polynomials, and using such profiles and zero-detecting polynomials to establish
-generic zero-free regions), although this is mostly just a change of notation and does not seem to yield any major simplifications to these arguments.
7 comments
Comments feed for this article
2 March, 2015 at 2:29 am
Anonymous
In the mean value inequality for
, it seems (but not stated) that the integral is with the area measure.
[Clarification added, thanks – T.]
2 March, 2015 at 9:03 am
Anonymous
Theorem 3 shows that Lindelof hypothesis implies that the elementary estimate
for the multiplicities of the non-trivial zeros (with large
) can be improved to
. Is it possible to improve the last estimate by letting (in theorem 3)
as
?
2 March, 2015 at 9:50 am
Terence Tao
My feeling is that if one only assumes the standard qualitative form of Lindelof (namely that
as
, with no information about the decay rate in the
exponent) then one cannot improve upon the
bound, basically because smashing together
nearby zeroes to form a single zero of high multiplicity would only increase the magnitude of the zeta function by about
or so at most near these zeroes, so for any
such an “edit” to the zeta function would be undetectable (at least by naive means) to the Lindelof hypothesis. (See also Remark 30 of Supplement 3 for a related example of “editing” the zeta function.)
However, if one assumed a more quantitative form of the Lindelof hypothesis, then one should be able to do better. For instance, it is a classical result of Littlewood that the RH implies
, which implies that any zero has multiplicity at most
. The implied constant was lowered to
by Goldston and Gonek; there has been some further work in this direction but this is still close to the state of the art for what one can say on RH.
2 March, 2015 at 10:31 am
Anonymous
Graph limits (iii)
for 


If :
and
then.
[There was a typo:
is only known for
, not
. This typo has now been fixed. -T.]
24 September, 2015 at 12:17 am
0f33418c7da5cbb5
Sorry for the comment that in some sense does not apply to this post.The zeros of the Riemann zeta function compares integers (zero exists or not).Applying Cantor’s diagonal method is easy to show that the conversion of all real numbers
interval does not exist(for some
and
– constant).Focusing above now we can say that there are an infinite number of real zeros in the half-interval.The function
will adjust the values of
-(defined in the Levinson) for their match with the actual numbers are not zeros of the zeta function.Now we can use the length molifer
instead of
.And following the Levinson can conclude that the number of zeros for some
has an overall principal term at
.
25 September, 2015 at 12:48 am
0f33418c7da5cbb5
Maybe I made a mistake, instead of the
need at least
.

instead 
It means
19 January, 2018 at 4:21 am
The De Bruijn-Newman constant is non-negativ | What's new
[…] we get to assume the Riemann hypothesis, we can sharpen this to , a result of Littlewood (see this previous blog post for a proof). Ideally, we would like to obtain similar control for the other , , as well. […]