Given any finite collection of elements in some Banach space
, the triangle inequality tells us that
However, when the all “oscillate in different ways”, one expects to improve substantially upon the triangle inequality. For instance, if
is a Hilbert space and the
are mutually orthogonal, we have the Pythagorean theorem
For sake of comparison, from the triangle inequality and Cauchy-Schwarz one has the general inequality
for any finite collection in any Banach space
, where
denotes the cardinality of
. Thus orthogonality in a Hilbert space yields “square root cancellation”, saving a factor of
or so over the trivial bound coming from the triangle inequality.
More generally, let us somewhat informally say that a collection exhibits decoupling in
if one has the Pythagorean-like inequality
for any , thus one obtains almost the full square root cancellation in the
norm. The theory of almost orthogonality can then be viewed as the theory of decoupling in Hilbert spaces such as
. In
spaces for
one usually does not expect this sort of decoupling; for instance, if the
are disjointly supported one has
and the right-hand side can be much larger than when
. At the opposite extreme, one usually does not expect to get decoupling in
, since one could conceivably align the
to all attain a maximum magnitude at the same location with the same phase, at which point the triangle inequality in
becomes sharp.
However, in some cases one can get decoupling for certain . For instance, suppose we are in
, and that
are bi-orthogonal in the sense that the products
for
are pairwise orthogonal in
. Then we have
giving decoupling in . (Similarly if each of the
is orthogonal to all but
of the other
.) A similar argument also gives
decoupling when one has tri-orthogonality (with the
mostly orthogonal to each other), and so forth. As a slight variant, Khintchine’s inequality also indicates that decoupling should occur for any fixed
if one multiplies each of the
by an independent random sign
.
In recent years, Bourgain and Demeter have been establishing decoupling theorems in spaces for various key exponents of
, in the “restriction theory” setting in which the
are Fourier transforms of measures supported on different portions of a given surface or curve; this builds upon the earlier decoupling theorems of Wolff. In a recent paper with Guth, they established the following decoupling theorem for the curve
parameterised by the polynomial curve
For any ball in
, let
denote the weight
which should be viewed as a smoothed out version of the indicator function of
. In particular, the space
can be viewed as a smoothed out version of the space
. For future reference we observe a fundamental self-similarity of the curve
: any arc
in this curve, with
a compact interval, is affinely equivalent to the standard arc
.
Theorem 1 (Decoupling theorem) Let
. Subdivide the unit interval
into
equal subintervals
of length
, and for each such
, let
be the Fourier transform
of a finite Borel measure
on the arc
, where
. Then the
exhibit decoupling in
for any ball
of radius
.
Orthogonality gives the case of this theorem. The bi-orthogonality type arguments sketched earlier only give decoupling in
up to the range
; the point here is that we can now get a much larger value of
. The
case of this theorem was previously established by Bourgain and Demeter (who obtained in fact an analogous theorem for any curved hypersurface). The exponent
(and the radius
) is best possible, as can be seen by the following basic example. If
where is a bump function adapted to
, then standard Fourier-analytic computations show that
will be comparable to
on a rectangular box of dimensions
(and thus volume
) centred at the origin, and exhibit decay away from this box, with
comparable to
On the other hand, is comparable to
on a ball of radius comparable to
centred at the origin, so
is
, which is just barely consistent with decoupling. This calculation shows that decoupling will fail if
is replaced by any larger exponent, and also if the radius of the ball
is reduced to be significantly smaller than
.
This theorem has the following consequence of importance in analytic number theory:
Corollary 2 (Vinogradov main conjecture) Let
be integers, and let
. Then
Proof: By the Hölder inequality (and the trivial bound of for the exponential sum), it suffices to treat the critical case
, that is to say to show that
We can rescale this as
As the integrand is periodic along the lattice , this is equivalent to
The left-hand side may be bounded by , where
and
. Since
the claim now follows from the decoupling theorem and a brief calculation.
Using the Plancherel formula, one may equivalently (when is an integer) write the Vinogradov main conjecture in terms of solutions
to the system of equations
but we will not use this formulation here.
A history of the Vinogradov main conjecture may be found in this survey of Wooley; prior to the Bourgain-Demeter-Guth theorem, the conjecture was solved completely for , or for
and
either below
or above
, with the bulk of recent progress coming from the efficient congruencing technique of Wooley. It has numerous applications to exponential sums, Waring’s problem, and the zeta function; to give just one application, the main conjecture implies the predicted asymptotic for the number of ways to express a large number as the sum of
fifth powers (the previous best result required
fifth powers). The Bourgain-Demeter-Guth approach to the Vinogradov main conjecture, based on decoupling, is ostensibly very different from the efficient congruencing technique, which relies heavily on the arithmetic structure of the program, but it appears (as I have been told from second-hand sources) that the two methods are actually closely related, with the former being a sort of “Archimedean” version of the latter (with the intervals
in the decoupling theorem being analogous to congruence classes in the efficient congruencing method); hopefully there will be some future work making this connection more precise. One advantage of the decoupling approach is that it generalises to non-arithmetic settings in which the set
that
is drawn from is replaced by some other similarly separated set of real numbers. (A random thought – could this allow the Vinogradov-Korobov bounds on the zeta function to extend to Beurling zeta functions?)
Below the fold we sketch the Bourgain-Demeter-Guth argument proving Theorem 1.
I thank Jean Bourgain and Andrew Granville for helpful discussions.
— 1. Initial reductions —
The claim will proceed by an induction on dimension, thus we assume henceforth that (the
case being immediate from the Pythagorean theorem) and that Theorem 1 has already been proven for smaller values of
. This has the following nice consequence:
Proposition 3 (Lower dimensional decoupling) Let the notation be as in Theorem 1. Suppose also that
, and that Theorem 1 has already been proven for all smaller values of
. Then for any
, the
exhibits decoupling in
for any ball
of radius
.
Proof: (Sketch) We slice the ball into
-dimensional slices parallel to the first
coordinate directions. On each slice, the
can be interpreted as functions on
whose Fourier transform lie on the curve
, where
. Applying Theorem 1 with
replaced by
, and then integrating over all slices using Fubini’s theorem and Minkowski’s inequality (to interchange the
norm and the square function), we obtain the claim.
The first step, needed for technical inductive purposes, is to work at an exponent slightly below . More precisely, given any
and
, let
denote the assertion that
whenever ,
,
and
are as in Theorem 1. Theorem 1 is then clearly equivalent to the claim
holding for all
. This turns out to be equivalent to the following variant:
Proposition 4 Let
, and assume Theorem 1 has been established for all smaller values of
. If
is sufficiently close to
, then
holds for all
.
The reason for this is that the functions and
all have Fourier transform supported on a ball of radius
, and so there is a Bernstein-type inequality that lets one replace the
norm of either function by the
norm, losing a power of
that goes to zero as
goes to
. (See Corollary 6.2 and Lemma 8.2 of the Bourgain-Demeter-Guth paper for more details of this.)
Using the trivial bound (1) we see that holds for large
(e.g.
). To reduce
, it suffices to prove the following inductive claim.
Proposition 5 (Inductive claim) Let
, and assume Theorem 1 has been established for all smaller values of
. If
is sufficiently close to
, and
holds for some
, then
holds for some
.
Since the set of for which
holds is clearly a closed half-infinite interval, Proposition 5 implies Proposition 4 and hence Theorem 1.
Henceforth we fix as in Proposition 5. We fix
and use
to denote any quantity that goes to zero as
, keeping
fixed. Then the
hypothesis reads
The next step is to reduce matters to a “multilinear” version of the above estimate, in order to exploit a multilinear Kakeya estimate at a later stage of the argument. Let be a large integer depending only on
(actually Bourgain, Demeter, and Guth choose
). It turns out that it will suffice to prove the multilinear version
whenever are families of disjoint subintervals on
of length
that are separated from each other by a distance of
, and where
denotes the geometric mean
We have the following nice equivalence (essentially due to Bourgain and Guth, building upon an earlier “bilinear equivalence” result of Vargas, Vega, and myself, and discussed in this previous blog post):
Proposition 6 (Multilinear equivalence) For any
, the estimates (2) and (3) are equivalent.
Proof: The derivation of (3) from (2) is immediate from Hölder’s inequality. To obtain the converse implication, let denote the best constant in (2), thus
is the smallest constant such that
The idea is to prove an inequality of the form
for any fixed integer (with the implied constant in the
notation independent of
); by choosing
large enough one can then prove
by an inductive argument.
We partition the intervals in (2) into
classes
of
consecutive intervals, so that
can be expressed as
where
. Observe that for any
, one either has
for some (i.e. one of the
dominates the sum), or else one has
for some with the transversality condition
. This leads to the pointwise inequality
Bounding the supremum by
and then taking
norms and using (3), we conclude that
On the other hand, applying an affine rescaling to (4) one sees that
and the claim follows. (A more detailed version of this argument may be found in Theorem 4.1 of this paper of Bourgain and Demeter.)
It thus suffices to show (3).
The next step is to set up some intermediate scales between and
, in order to run an “induction on scales” argument. For any scale
, any exponent
, and any function
, let
denote the local
average
where denotes the volume of
(one could also use the equivalent quantity
here if desired). For any exponents
,
, and
(independent of
), let
denote the least exponent for which one has the local decoupling inequality
for as in (3), where the
-length intervals in
have been covered by a family
of finitely overlapping intervals of length
, and
. It is then not difficult to see that the estimate (3) is equivalent to the inequality
(basically because when , there is essentially only one
for each
, and
is basically
; also; the averaging
is essentially the identity when
since all the
and
here have Fourier support on a ball of radius
). To put it another way, our task is now to show that
On the other hand, one can establish the following inequalities concerning the quantities , arranged roughly in increasing order of difficulty to prove.
Proposition 7 (Inequalities on
) Throughout this proposition it is understood that
,
, and
.
- (i) (Hölder) The quantity
is convex in
, and monotone nondecreasing in
.
- (ii) (Minkowski) If
, then
is monotone non-decreasing in
.
- (iii) (Stability) One has
. (In fact,
is Lipschitz in
uniformly in
, but we will not need this.)
- (iv) (Rescaled decoupling hypothesis) If
and
, then one has
.
- (v) (Lower dimensional decoupling) If
and
, then
.
- (vi) (Multilinear Kakeya) If
and
, then
.
We sketch the proof of the various parts of this proposition in later sections. For now, let us show how these properties imply the claim (6). In the paper of Bourgain, Demeter, and Guth, the above properties were iterated along a certain “tree” of parameters , relying in (v) to increase the
parameter (which measures the amount of decoupling) and (vi) to “inflate” or increase the
parameter (which measures the spatial scale at which decoupling has been obtained), and (i) to reconcile the different choices of
appearing in (v) and (vi), with the remaining properties (ii), (iii), (iv) used to control various “boundary terms” arising from this tree iteration. Here, we will present an essentially equivalent “Bellman function” formulation of the argument which replaces this iteration by a carefully (but rather unmotivatedly) chosen inductive claim. More precisely, let
be a small quantity (depending only on
and
) to be chosen later. For any
, let
denote the claim that for every
, and for all sufficiently small
, one has the inequality
From Proposition 7 (i), (ii), (iv), we see that holds for some small
. We will shortly establish the implication
for some independent of
; this implies upon iteration that
holds for arbitrarily large values of
. Applying (9) with
for a sufficiently large
and a sufficiently small
, and combining with Proposition 7(iii), we obtain the claim (6).
We now prove the implication (10). Thus we assume (7) holds for , sufficiently small
, and
obeying (8), and also (9) for
and we wish to improve this to
for the same range of and for sufficiently small
, and also
By Proposition 7(i) it suffices to show this for the extreme values of , thus we wish to show that
We begin with (13). The case of this estimate is
But since , we see that
if
is small enough, so the right-hand side of (16) is greater than
and the claim follows from Proposition 7(iv) (with a little bit of room to spare). Now we look at the
cases of (13). By Proposition 7(vi), we have
For close to
,
lies between
and
, so from (7) one has
Since , one has
for small enough depending on
, and (13) follows (if
is small enough depending on
but not on
).
The same argument applied with gives
Since , we thus have
if are sufficiently small depending on
(but not on
). This, together with Proposition 7(i), gives (15).
Finally, we establish (14). From Proposition 7(v) (with replaced by
) we have
In the case, this gives
and the claim (14) follows from (15) in this case. Now suppose . Since
is close to
,
lies between
and
, and so we may apply (7) to conclude that
and hence (after simplifying)
which gives (14) for small enough (depending on
, but not on
).
— 2. Rescaled decoupling —
The claims (i), (ii), (iii) of Proposition 7 are routine applications of the Hölder and Minkowski inequalities (and also the Bernstein inequality, in the case of (iii)); we will focus on the more interesting claims (iv), (v), (vi).
Here we establish (iv). The main geometric point exploited here is that any segment of the curve is affinely equivalent to
itself, with the key factor of
in the bound
coming from this affine rescaling.
Using the definition (5) of , we see that we need to show that
for balls of radius
. By Hölder’s inequality, it suffices to show that
for each . By Minkowski’s inequality (and the fact that
), the left-hand side is at most
so it suffices to show that
for each . From Fubini’s theorem one has
so we reduce to showing that
But this follows by applying an affine rescaling to map to
, and then using the hypothesis
with
replaced by
. (The ball
gets distorted into an ellipsoid, but one can check that this ellipsoid can be covered efficiently by finitely overlapping balls of radius
, and so one can close the argument using the triangle inequality.)
— 3. Lower dimensional decoupling —
Now we establish (v). Here, the geometric point is the one implicitly used in Proposition 3, namely that the -dimensional curve
projects down to the
-dimensional curve
for any
.
Let be as in Proposition 7(v). From (5), it suffices to show that
for balls of radius
. It will suffice to show the pointwise estimate
for any , or equivalently that
where . Clearly this will follow if we have
for each . Covering the intervals in
by those in
, it suffices to show that
for each . But this follows from Proposition 3.
— 4. Multidimensional Kakeya —
Finally, we establish (vi), which is the most substantial component of Proposition 7, and the only component which truly takes advantage of the reduction to the multilinear setting. Let and
be such that
. From (5), it suffices to show that
for balls of radius
. By averaging, it suffices to establish the bound
for balls of radius
. If we write
, the right-hand side simplifies to
so it suffices to show that
At this point it is convenient to perform a dyadic pigeonholing (giving up a factor of ) to normalise, for each
, all of the quantities
to be of comparable size, after reducing the sets
so some appropriate subset
. (The contribution of those
for which this quantity is less than, say,
of the maximal value, can be safely discarded by trivial estimates.) By homogeneity we may then normalise
for all surviving , so the estimate now becomes
Since is close to
,
is less than
, so we can estimate
and so it suffices to show that
or, on raising to the power ,
Localising to balls of radius
, it suffices to show that
The arc is contained in a box of dimensions roughly
, so by the uncertainty principle
is essentially constant along boxes of dimensions
(this can be made precise by standard methods, see e.g. the discussion in the proof of Theorem 5.6 of Bourgain-Demeter-Guth, or my general discussion on the uncertainty principle in this previous blog post). This implies that
, when restricted to
, is essentially constant on “plates”, defined as the intersection of
with slabs that have
dimensions of length
and the remaining
dimensions infinite (and thus restricted to be of length about
after restriction to
). Furthermore, as
varies (and
is constrained to be in
, the orientation of these slabs varies in a suitably “transverse” fashion (the precise definition of this is a little technical, but can be verified for
; see the BDG paper for details). After rescaling, the claim then follows from the following proposition:
Proposition 8 (Multilinear Kakeya) For
, let
be a collection of “plates” that have
dimensions of length
, and
dimensions that are infinite, and for each
let
be a non-negative number. Assume that the families of plates
obey a suitable transversality condition. Then
for any ball
of radius
.
The exponent here is natural, as can be seen by considering the example where each
consists of about
parallel disjoint plates passing through
, with
for all such plates.
For (where the plates now become tubes), this result was first obtained by Bennett, Carbery, and myself using heat kernel methods, with a rather different proof (also capturing the endpoint case) later given using algebraic topological methods by Guth (as discussed in this previous post. More recently, a very short and elementary proof of this theorem was given by Guth, which was initially given for
but extends to general
. The scheme of the proof can be described as follows.
- When all the plates
in a each family
are parallel, the claim follows from the Loomis-Whitney inequality (when
) or a more general Brascamp-Lieb inequality of Bennett, Carbery, Christ, and myself (for general
). These inequalities can be proven by a repeated applications of the Hölder inequality and Fubini’s theorem.
- Perturbing this, we can obtain the proposition with a loss of
for any
and
, provided that the plates in each
are within
of being parallel, and
is sufficiently small depending on
and
. (For the case of general
, this requires some uniformity in the result of Bennett, Carbery, Christ, and myself, which can be obtained by hand in the specific case of interest here, but was recently established in general by Bennett, Bez, Flock, and Lee.
- A standard “induction on scales” argument shows that if the proposition is true at scale
with some loss
, then it is also true at scale
with loss
. Iterating this, we see that we can obtain the proposition with a loss of
uniformly for all
, provided that the plates are within
of being parallel and
is sufficiently small depending now only on
(and not on
).
- A finite partition of unity then suffices to remove the restriction of the plates being within
of each other, and then sending
to zero we obtain the claim.
The proof of the decoupling theorem (and thus the Vinogradov main conjecture) are now complete.
Remark 9 The above arguments extend to give decoupling for the curve
in
for every
. As it turns out (Bourgain, private communication), a variant of the argument also handles the range
, and the range
can be covered from an induction on dimension (using the argument used to establish Proposition 3).
59 comments
Comments feed for this article
10 December, 2015 at 1:14 pm
sylvainjulien
Dear Terry,
Following the strengthening of the triangle inequality you consider at the beginning of this post and which involves the square root of the cardinality of a set, can the optimal error term in the prime number theorem, which is a classical reformulation of RH, be interpreted as some kind of orthogonality of basis vectors coming from prime numbers in some Hilbert space? If so, is there any link with Selberg’s orthogonality conjecture for the Selberg class where primitive elements thereof replace prime numbers?
10 December, 2015 at 2:06 pm
Anonymous
In theorem 1, which properties are needed for the measure
(e.g. positiveness, absolute continuity wrt the arc length measure, etc.) ?
[The measure needs to be finite Borel, this has now been added. -T.]
10 December, 2015 at 4:25 pm
Anonymous
The way w_B is defined it looks more like a smoothed out dirac measure of x_0 rather than a smoothed out indicator function of B. Can someone explain?
10 December, 2015 at 5:16 pm
Terence Tao
The function is normalised to have a sup norm comparable to 1 (like that of an indicator function) rather than to have total mass comparable to 1 (like that of a delta function).
10 December, 2015 at 11:44 pm
Anonymous
In addition of being measurable and having sup norm comparable to 1, are there more properties needed for
?
11 December, 2015 at 8:21 am
Terence Tao
As I said in the post, one should think of
as being a smoothed out version of
, which is localised to
and has the local constancy property
when
(
being the radius of
).
11 December, 2015 at 3:10 am
Anonymous
The word immediately before (1) is “inequailty”, and between (9) and (10) there is “\eqreF{impl}” which I think should be rendered as a link to (9).
[Corrected, thanks – T.]
11 December, 2015 at 4:51 pm
The two-dimensional case of the Bourgain-Demeter-Guth proof of the Vinogradov main conjecture | What's new
[…] this blog post, I would like to specialise the arguments of Bourgain, Demeter, and Guth from the previous post to the two-dimensional case of the Vinogradov main conjecture, […]
16 December, 2015 at 11:42 am
Alastair Irving
A very small correction: when you state the existing bounds due to Wooley you use k instead of n in a couple of equations.
[Corrected, thanks – T.]
16 December, 2015 at 9:46 pm
Anonymous
The
defined in Theorem 1 are defined as an integral over
yet immediately afterwards, the
are measures on
, so shouldn’t the integral definition of
be over
instead?
[Corrected, thanks – T.]
18 December, 2015 at 1:43 am
Anonymous
Why the exponent
in the definition of
is chosen so large? What seems to be its minimal possible value?
18 December, 2015 at 4:11 am
Anonymous
As the text explains, $w_B$ is meant to be a smoothed replacement of $1_B$, so if you want to be completely rigorous, the exponent in the definition should be a variable, and in the end you should take a limit where that variable goes to infinity. So I don’t think that a question about the “minimal possible value” of the exponent makes any sense, actually.
18 December, 2015 at 4:44 am
Anonymous
It is not clear why
was chosen? (if e.g.
is sufficient ) and which constraints for this exponent are really needed.
18 December, 2015 at 7:50 am
Terence Tao
Basically one wants
, as well as any of the fractional powers of
which may arise in various invocations of Holder’s inequality, to be absolutely integrable. One can certainly dig through the proof to see exactly which fractional powers of
show up and thence work out the minimal exponent here, but this value is not terribly important to the rest of the argument and so it is generally preferable to just set this exponent to a conservatively large value of no particular significance, such as
. (In general, whenever nice round numbers such as
,
, or
appear in an analysis argument, they are probably used as a convenient choice of a large constant whose exact value is of little importance to the rest of the argument, so long as it is large enough.)
19 December, 2015 at 6:36 pm
Anonymous
Are you sure that the application of Holder in the first sentence of the proof of Corollary 2 is correct? It seems that if I use the trivial bound on the exponential sum and the case of
, we have
and so the
norm is
which is worse than either term in the statement of Corollary 2.
20 December, 2015 at 9:11 am
Terence Tao
Holder’s inequality gives
, rather than the first inequality asserted in your comment.
20 December, 2015 at 1:44 am
benjdewantara
Reblogged this on benjdewantara.
20 December, 2015 at 10:41 pm
Anonymous
Dear Tao, reading the prepring in Arxiv, in Lemma 7.2, instead of
I get
, that seems better. The
factor appears when I rescale like in Lemma 6.3. Am I misunderstanding something?
20 December, 2015 at 11:06 pm
Anonymous
I should say, I am rescaling the lower dimensional curve, not the original one.
23 December, 2015 at 4:25 pm
Terence Tao
I think the
factor is simply discarded by Bourgain-Guth-Demeter as it is not needed for their argument (in lower dimensions it ends up that the scale parameter just gets raised to the power
, so it really doesn’t matter too much exactly what it is.)
25 December, 2015 at 10:50 pm
Anonymous
Can you say a bit more about the derivation of (2) from (3) in the proof of Proposition 6? What are you taking the
to be when you try to prove (2)?
26 December, 2015 at 12:51 am
Terence Tao
See the second paragraph after (4). If
is a multiple of
, one can take
to be the intervals
; when
is not a multiple of
, one can round the endpoints of this string of consecutive intervals to the nearest integer.
26 December, 2015 at 11:59 am
Anonymous
There seems to be a typo in the first sentence of the proof of Proposition 6, (2) and (3) should be swapped then. However, I do not see how Holder’s inequality can show (3) from (2) since it seems that (3) is true for any collection of subintervals.
[Corrected, thanks -T.]
26 December, 2015 at 10:37 pm
Anonymous
By Holder’s inequality, the left hand side of (3) is
, and so to prove (3), one needs (2) with the
replaced by
. Is this still true?
27 December, 2015 at 8:50 am
Terence Tao
Yes. If the intervals in
are of the form
for a natural number value of
, then
is a subset of the
intervals considered in Theorem 1, and one can simply add some “dummy”
(that are equal to zero when the interval
is not of the form considered in Theorem 1) to conclude.
If the intervals
are not all of the above form (or equivalently, if some of the
are non-integer reals), one can use the intervals in Theorem 1 to cut each of the
into at most two subintervals, and each
then gets split into at most two pieces, each of which has an
norm controlled by the
norm of the original function
by the
boundedness of the Hilbert transform. Each interval in Theorem 1 is now associated to at most two of these pieces, and so
can be broken up into at most two sums, each of which can be controlled adequately by (2) and the boundedness of the Hilbert transform.
(Another approach, avoiding the use of the boundedness of the Hilbert transform, uses the “1/3 translation trick”, which I believe is originally due to Michael Christ: every interval of side length
can be covered by an interval of the form
or by
for some integer
. Using this trick and the triangle inequality, one can often reduce questions about intervals of a given length to intervals with (three times) the given length, and whose endpoints are integer multiples of that length (after exploiting some translation invariance to get rid of the shift).
In any event, this direction of the implication in Proposition 6 is not actually used in the proof of the decoupling conjecture, and is only present to give some context to the relative strength of the linear and multilinear formulations of the decoupling theorem.
28 December, 2015 at 3:17 am
Anonymous
Which properties of the integrand in corollary 2 are needed (and used) by the decoupling method, and is there a more general family of similar multidimentional integrals which (potentially) can be estimated by this new method ?
28 December, 2015 at 8:09 am
Terence Tao
The decoupling argument would apply to any curve
that was maximally curved in the sense that at any point of the curve, the first
derivatives of the curve parameterisation
were linearly independent. (For curves in three dimensions, this amounts to the curvature and torsion being both non-zero throughout the curve.) To get estimates of the form in Corollary 2, though, one needs some integrality properties of the frequency points being sampled, since one has to exploit some periodicity in the proof of the corollary. Jean Bourgain has informed me (private communication) that the method used to reprove Corollary 2 can be modified for instance to prove a classical inequality of Hua.
Also, you are indeed correct that Corollary 2 extends to the case where
is non-integer.
28 December, 2015 at 4:52 am
Anonymous
In the formulation of corollary 2, is it really necessary of
to be an integer?
28 December, 2015 at 3:57 pm
Anonymous
Very first word of the post: “\iGiven” –> “Given”
[Corrected, thanks – T.]
10 January, 2016 at 10:26 am
Anonymous
Is it possible to combine the B-D-G decoupling technique with the (apparently very different but still closely related) efficient congruencing technique of Wooley?
[I believe people are looking at this right now, but I do not know of any publicly available work connecting the two thus far. -T.]
18 January, 2016 at 3:48 pm
Anonymous
Dear Terry,
Is the Theorem 5.5 of the Bourdain-Demeter-Guth paper directly follows from Prop. 5.2 , Lemma 5.3 and Prop. 5.4 in the same section ? It seems to me that conditions of 5.5 just ensure that the previous three propositions apply. But I do not understand that why we have an extra factor R^{\epsilon} in (7) since we are only dealing with parallel plates in each family. In the
one-sentence “proof”, the authors said that this is proved in Guth [14]. Do they mean that we also need to apply Guth’s arguments with 5.2 , 5.3 and 5.4 to get Theorem 5.5?
Maybe I am totally wrong here. Can you he me to clarify this a little bit? Many thanks.
22 January, 2016 at 2:37 pm
Terence Tao
Yes, one needs to apply the arguments from [14] with Prop 5.2, Lemma 5.3, and Prop 5.4 inserted in those arguments in the appropriate places. Guth’s argument, based on an induction on scales, loses the
factor. I would recommend reading through the argument in [14] (which treats the k=n-1 case) first, as it should become clearer after doing so how the argument is to be modified.
26 January, 2016 at 11:52 am
Tom Church
You credit Bourgain and Demeter for Theorem 1, but shouldn’t it be Bourgain-Demeter-Guth? Currently Guth’s name doesn’t appear anywhere until much farther down the post. (Perhaps just a typo, since you link to the BDG paper.)
[Corrected, thanks. Bourgain and Demeter have an earlier series of papers on decoupling that precede the BDG paper. -T]
26 January, 2016 at 10:21 pm
Anonymous
In corollary 2, is it possible to give an effective estimate for the
loss in terms of
?
27 January, 2016 at 5:20 am
Terence Tao
In principle, the arguments are effective, but the dependence of the implied constants on
are somewhat poor (in particular, it does not seem that these results immediately imply any substantial improvement on the Vinogradov-Korobov zero-free region).
27 January, 2016 at 11:57 pm
Anonymous
What seems to be (in terms of fixed
and large
)
taken over the RHS in corollary 2 ?
31 January, 2016 at 7:37 am
Polymath10-post 4: Back to the drawing board? | Combinatorics and more
[…] Bourgain, Ciprian Demeter, Larry Guth of Vinogradov’s main conjecture. You can read about it here and […]
23 February, 2016 at 12:37 pm
Anonymous
It seems that there is a small misprint in the chain of inequalities for the biorthogonal decoupling in L^4. The power 1/2 in the last three lines should not be there. Otherwise the inequalities don’t scale properly.
[Corrected, thanks – T.]
5 September, 2016 at 2:11 am
Zak
Could you elaborate on your comment immediately following Theorem 1 regarding orthogonality? I think I am missing something.
[Can you elaborate on your request for elaboration? -T.]
5 September, 2016 at 2:24 am
Zak
Also, in Corollary 2, immediately following “We can rescale as” the exponent of
should be
.
[Corrected, thanks – T.]
6 September, 2016 at 4:52 pm
Zak
I am not sure how to show the
are orthogonal (or almost orthogonal)
6 September, 2016 at 5:11 pm
Terence Tao
One has to compare the weight
with a weight
that has a compactly supported Fourier transform (actually it is convenient to ensure that it is
that has compactly supported Fourier transform, so that the
remain orthogonal in
). See for instance Proposition 6.1 of http://arxiv.org/abs/1604.06032 . (Note that the lower bound in that proposition is not quite correct as stated; I think the notes will be revised soon to address this small issue.)
31 March, 2017 at 1:51 pm
El triunfo del análisis. Conjetura de Vinogradov – Blog del Instituto de Matemáticas de la Universidad de Sevilla
[…] Blog de Terry Tao […]
7 May, 2017 at 6:06 am
Anonymous
Dear Prof. Tao,
in the most recent paper https://arxiv.org/abs/1604.06032 (let’s call it paper A) by Bourgain and Demeter on decoupling the authors write that this paper should serve as a “warm up” for unterstanding the proof of the Vinogradov conjecture by presenting a more streamlined proof of their
decoupling theorem from https://arxiv.org/pdf/1403.5335.pdf (paper B). But to me it looks like the result in paper A is much weaker (and actually different) than the result in paper B.
In paper A the authors establish a decoupling theorem for functions of the form
where
is the lift of the Lebesgue measure from
to the paraboloid in the exponent range
. Whereas in paper B the authors prove a decoupling theorem for any Schwartz function whose Fourier transform has support in a certain neighborhood of the paraboloid for exponents
.
How can the result in paper A be seen as the same as or being comparable to the result in paper B?
10 May, 2017 at 10:40 am
Terence Tao
The decoupling estimate in Paper B for
implies a decoupling estimate for
also (see equation (3) of that paper). If one can decouple Schwartz functions Fourier supported in a neighbourhood of the paraboloid, one can then decouple non-Schwartz functions Fourier supported exactly on the paraboloid by using approximations to the identity and a limiting argument. (One can also use uncertainty principle type arguments to go in the reverse direction, if the thickness of the neighbourhood is comparable to the scale at which the uncertainty principle applies.)
10 May, 2017 at 5:41 pm
Anonymous
Thank you for your answer!
spaces. But I guess the weights can be omitted if one deals with Schwartz functions.) Is this correct?
by using the machinery developed in paper A (maybe by running a similar iteration scheme as in section 10 of paper A but with a different value for
)? Or does the usage of the multilinear Kakeya inequality rather than the multilinear restriction theorem limit oneself to the case
?
If I understand correctly, theorem 5.1 in paper A says that the result proven in paper A is essentially equivalent to the decoupling theorem in paper B in the case of the paraboloid. (Ignoring the fact that in Thm. 5.1 one deals with weighted
Furthermore, is it possible to prove the decoupling theorem as stated in paper B for
13 May, 2017 at 8:45 am
Zane Li
I can try to answer both your questions:
in paper A are slightly different from the ones in paper B.
1. Yes, Theorem 5.1 basically says that if I know how big decoupling constant is in paper A, then I know how big the decoupling constant is in paper B. Remark 5.2 in paper B remarks on the equivalence of various decoupling constants. Note that the
2. I imagine it could be possible, since multilinear Kakeya is used only in the proof of ball inflation (Theorem 9.2 of paper A) and only imposes the requirement that
. However, note that the argument relies on that
(equation (50) in paper A) and as I get closer to
,
gets closer to 1. So you may need to rearrange the argument a bit.
I think the standard argument to get the decoupling constant in the range
is to first prove it for the critical
and then interpolate (via Holder) with the trivial
bound which comes just from Cauchy-Schwarz. Finally, use the uncertainty principle at the end to relate the
sums of the
and
pieces to the
sum of the
piece. (See proposition 10 in Jonathan Hickman’s notes: http://math.uchicago.edu/~j.e.hickman/Decoupling%20notes%205.pdf)
12 June, 2017 at 5:38 am
Anonymous
Dear Prof. Tao,
may I ask you to elaborate on how the dyadic pigeonhole argument in the proof of (vi) in Prop. 7 works?
In particular, I do not understand what the required estimates are to justify that one can discard the contribution of those
for which the quantities
are less than
of the maximum value. In addition, I don’t quite understand how one comes up with the factor
when one normalises the above quantities to be of comparable size.
22 June, 2017 at 5:04 pm
Terence Tao
Let’s normalise the maximum value of
for
to equal
for each
, so the RHS is at least 1. Suppose we have some
and some
for which
. This implies a pointwise estimate of the form
(say). Meanwhile, for all other
and all other
we similarly have
, and this is enough to make the contribution of all such
to the LHS to be negligible.
After the quantities
have been restricted to lie between
and
, there are only
dyadic ranges that they could lie in, so by decomposing into these ranges and using the triangle inequality one can restrict to a single dyadic range for each
(giving up a factor of
).
30 June, 2017 at 5:41 am
Anonymous
Dear Prof. Tao,
I’m currently studying the latest paper of Bourgain and Demeter on decoupling (https://arxiv.org/abs/1604.06032), but there is one issue I just don’t understand. In the proof of Prop. 8.4 in order to apply Thm. 5.1 the authors claim that the Fourier transform of the function
(let’s call this function
for fixed
is supported in the
neighborhood of the parabola
. But for Thm. 5.1 I need that the Fourier support lies in the smaller set
for some constant
.
My question is the following: How is the Fourier transform of
defined in order to give the author’s claim above any sense? Since
does not lie in any
space, the usual definition does not make sense, does it?
If I consider
as a distribution, then I’m able to conclude that it has Fourier support (in the distributional sense) in
. But that doesn’t seem to help me when I want to apply Thm. 5.1.
May I ask you to elaborate on this issue and explain how equation (21) in the above paper can be derived?
24 August, 2017 at 6:33 am
Dmitrii Zhelezov
Dear Terence,
I wonder if some of the decoupling techniques may be applicable in the finite field setting (it’s more of a blind guess rather than suggestion)?
24 August, 2017 at 8:01 am
Terence Tao
A key input of decoupling is the ability to induct on multiple scales, and we don’t seem to have an analogue of this in finite fields (where there are basically only two scales, the scale of a single point, and the scale of the entire field), except possibly in those situations where the finite field has a number of non-trivial subfields. On the other hand, Wooley has a recent preprint at https://arxiv.org/abs/1708.01220 where he develops a nested efficient congruencing method that is in some sense a “p-adic analogue” of decoupling (and in particular provides a second proof of the Vinogradov main conjecture) and which can be applied relatively easily to function field settings.
24 August, 2017 at 11:26 pm
Anonymous
Is it possible to combine both methods (to a possibly stronger method)?
16 March, 2018 at 7:37 am
DZ
In the proof of Proposition 6, should ‘for some
' be read for some
?
[Corrected, thanks – T.]
31 December, 2018 at 11:01 pm
L’impatto matematico di Jean Bourgain | Maddmaths!
[…] La stima della quantità (J_{s,k}(N)), introdotta da Vinogradov, è importante in problemi analitici e combinatorici della teoria dei numeri (problema di Waring, somme di Weyl, etc.). La congettura di Vinogradov dice che, fissati (k,,s), per ogni (epsilon>0) si ha [J_{s,k}(N)ll N^{s+epsilon}+N^{2s-n(n+1)/2+epsilon}] as (Ntoinfty). Risultati parziali era conosciuti, e Wooley aveva risolto i casi (k=1,2,3) e certi intervalli di valori (s) e (k). La soluzione completa è stata ottenuta da Bourgain, in collaborazione con Demeter e Guth. A differenza della dimostrazione di Wooley, basata sul cosiddetto metodo di fare congruenze efficentemente, la loro dimostrazione è basata su nuovi strumenti di analisi armonica. In particolare, in un lavoro precedente con Demeter su Annals del 2015, Bourgain ha dimostrato delle proprietà di disaccoppiamento in (L^{n(n+1)}) (una specie di proprietà di ortogonalità debole, molto più flessibile della classica ortogonalità in (L^2)) per la famiglia di trasformate di Fourier [f_i(x)=int_{gamma(J_i)}e(xcdot xi),dsigma(xi)qquadhbox{with}qquad J_i=gamma([frac{(i-1)} N, frac iN])]associate alle curve di potenza (gamma(t)=(t,t^2,ldots,t^n)), (tin [0,1]). Questa è la nuova idea chiave nella dimostrazione della congettura di Vinogradov. Una descrizione più estesa la potete trovare nel blog di Terence Tao. […]
14 June, 2019 at 2:34 pm
Abstracting induction on scales arguments | What's new
[…] It was shown by Bourgain, Demeter, and Guth that is of subpolynomial size for all in the optimal range , which among other things implies the Vinogradov main conjecture (as discussed in this previous post). […]
20 February, 2020 at 11:35 am
Anonymous
The answer to your question “A random thought – could this allow the Vinogradov-Korobov bounds on the zeta function to extend to Beurling zeta functions?” is no, see this paper
Diamond, Harold G.(1-IL); Montgomery, Hugh L.(1-MI); Vorhauer, Ulrike M. A.(1-KNTS) Beurling primes with large oscillation.
in which they show that 1 – c / \log t is the best zero-free region one can get in the case of Beurling primes (one can construct sets of Beurling primes with zeros beyond this region).
21 February, 2020 at 12:43 pm
Terence Tao
It’s true that this technically answers the question in the negative, but one could still hope that if one imposes stronger conditions on the Beurling primes (for instance, some uniform spacing between Beurling integers) then one could use decoupling methods to improve upon the classical zero free region. I haven’t looked into this carefully, though. [Admittedly, the uniform spacing hypothesis is extremely strong and it may well be that it excludes all examples that aren’t essentially arithmetic in nature, but still it would emphasise the point that the Vinogradov zero free region is not inherently arithmetic in nature.]
13 April, 2020 at 1:04 pm
Adicional 1: Otras introducciones a la teoría de desacople – Restricción y Desacople (Lic) – Teoría de Restricción y Desacople (Doc)
[…] https://terrytao.wordpress.com/2015/12/10/decoupling-and-the-bourgain-demeter-guth-proof-of-the-vino… […]