Zum Hauptinhalt springen

Quantitative estimates and extrapolation for multilinear weight classes.

Nieraeth, Bas
In: Mathematische Annalen, Jg. 375 (2019-10-01), Heft 1/2, S. 453-507
Online academicJournal

Quantitative estimates and extrapolation for multilinear weight classes 

In this paper we prove a quantitative multilinear limited range extrapolation theorem which allows us to extrapolate from weighted estimates that include the cases where some of the exponents are infinite. This extends the recent extrapolation result of Li, Martell, and Ombrosi. We also obtain vector-valued estimates including ℓ ∞ spaces and, in particular, we are able to reprove all the vector-valued bounds for the bilinear Hilbert transform obtained through the helicoidal method of Benea and Muscalu. Moreover, our result is quantitative and, in particular, allows us to extend quantitative estimates obtained from sparse domination in the Banach space setting to the quasi-Banach space setting. Our proof does not rely on any off-diagonal extrapolation results and we develop a multilinear version of the Rubio de Francia algorithm adapted to the multisublinear Hardy–Littlewood maximal operator. As a corollary, we obtain multilinear extrapolation results for some upper and lower endpoints estimates in weak-type and BMO spaces.

Keywords: 42B25; 42B20

Introduction

An essential tool in the theory of singular operators is extrapolation. In one of its forms, the classical extrapolation theorem of Rubio de Francia [[14]] says that if an operator T satisfies Lq(w) boundedness for a fixed q(1,) and for all weights w in the Muckenhoupt class Aq , then T is in fact bounded on Lp(w) for all p(1,) and all wAp .

Many variations of Rubio de Francia's extrapolation theorem have appeared over the years adapted to various situations. A multilinear version of the extrapolation result was found by Grafakos and Martell [[17]]. Another version provided by Auscher and Martell [[1]] dealt with operators bounded only for a limited range of p rather than for all p(1,) . Combining these approaches, it was shown by Cruz-Uribe and Martell [[7]] that if there are 0rj<sj and qj[rj,sj] , qj0, , such that an m-linear operator T satisfies

1.1 T(f1,...,fm)Lq(wq)cj=1mfjLqj(wjqj)

Graph

for all weights wjqj in the restricted Muckenhoupt and Reverse Hölder class Aqj/rjRH(sj/qj) , where w=j=1mwj , 1q=j=1m1qj , then T satisfies the same boundedness for all pj(rj,sj) and all wjpjApj/rjRH(sj/pj) , as well as certain vector-valued bounds.

In the linear setting for operators satisfying weighted bounds, it need not be the case that they are bounded on L , as is the case, for example, for the Hilbert transform. In particular, it is impossible to extrapolate estimates to this endpoint. This is in contrast to what happens in the multilinear setting, where it may very well occur that singular integral operators satisfy boundedness as in (1.1), but with some of the qj being equal to . This brings an interest to the question whether it is possible to extrapolate to bounds that include these endpoint cases pj= , starting from an initial weighted estimate where the qj are also allowed to be infinite. In this work we develop a method that does include these cases based on a multilinear Rubio de Francia algorithm. To facilitate this we give a natural extension in the definition of the weight classes to include these cases, see Definition 2.1 below. We point out that it is also possible to obtain these endpoint cases through off-diagonal extrapolation methods [[32]].

As an application for the theory, one can consider the bilinear Hilbert transform BHT given by

BHT(f1,f2)(x):=p.v.Rf1(x-t)f2(x+t)dtt,

Graph

which plays a central role in the theory of time-frequency analysis. It was shown by Lacey and Thiele [[25]] that BHT is bounded Lp1×Lp2Lp with 1p=1p1+1p2 if 1<p1,p2 and 23<p< . Through the helicoidal method of Benea and Muscalu [[2], [4]], vector-valued bounds of the form Lp1(q1)×Lp2(q2)Lp(q) were established in this range of p1 , p2 , p for various choices of 1<q1,q2 , 23<q< with 1q=1q1+1q2 . However, they left open the problem whether one can obtain vector-valued bounds for all q1 , q2 , q in the same range as Lacey and Thiele's theorem, i.e., for all 1<q1,q2 with 23<q< . While BHT satisfies weighted bounds as well as more general sparse bounds, see [[3], [8]], the extrapolation result by Cruz-Uribe and Martell [[7]] does not allow one to cover the full range of exponents. In particular, their result cannot retrieve any of the vector-valued bounds involving spaces. Such bounds also fall outside of the extrapolation result of Lorist and the author [[35]] where vector valued extensions of multilinear operators were considered in the setting of UMD Banach spaces, since does not satisfy the UMD property. The problem seems to be that the multilinear nature of the problem is not completely utilized when one imposes individual conditions on the weights rather than involving an interaction between the various weights.

In the recent work [[33]] by Li, Martell, and Ombrosi an extrapolation result was presented where they work with a limited range version of the multilinear weight condition introduced by Lerner, Ombrosi, Pérez, Torres, and Trujillo-González [[30]] which also appears in [[3]] and, in the bilinear case, in [[8]]. Indeed, such weight classes are characterized by boundedness of the multi-sublinear Hardy–Littlewood maximal operator as well as by boundedness of sparse forms, meaning the theory can be applied to important operators such as multilinear Calderón-Zygmund operators as well as the bilinear Hilbert transform. They introduced the weight class Ap,r where p=(p1,...,pm) , r=(r1,...,rm+1) and 1rjpj< and rm+1>p with 1p=j=1m1pj and w=(w1,...,wm)Ap,r if

1.2 [w]Ap,r:=supQa cube1|Q|Q(j=1mwjppj)rm+1rm+1-pdx1p-1rm+1j=1m1|Q|Qwjrjrj-pjdx1rj-1pj<.

Graph

They showed that if (1.1) holds for a q with 1rjqj< , rm+1>q and all (w1q1,...,wmqm)Aq,r , then T satisfies the same boundedness for all p and (w1p1,...,wmpm)Ap,r with rj<pj< and rm+1>p . Furthermore, their result extends and reproves some of the vector-valued bounds found by Benea and Muscalu [[4]] for BHT . This class of weights does seem to be adapted to the situation even when pj= , but one needs to be careful in how the constant is interpreted in this case. Similar to the proof of the extrapolation result of Cruz-Uribe and Martell, their proof of this extrapolation result is based upon an off-diagonal extrapolation result, but in their work they left open exactly what happens in the case that some of the exponents are infinite. They announced a paper in which these cases were treated which had not appeared yet when our paper was first posted, but is available now [[32]]. Here they show that, as a feature of off-diagonal extrapolation, it is also possible to obtain estimates that include the cases of infinite exponents.

In this work we again prove an extrapolation result using the multilinear weight classes, and our result includes these endpoint cases which, in particular, include the possibility of extrapolating from the cases where in the initial assumption the exponents can be infinite. Our proof is new and does not rely on any off-diagonal extrapolation result. Rather, we generalize the Rubio de Francia algorithm to a multilinear setting adapted to the multi-sublinear Hardy–Littlewood maximal operator. As a corollary, we are able to obtain vector-valued extensions of operators to spaces including spaces. Thus, applying this to BHT allows us to recover these endpoint bounds that were obtained earlier through the helicoidal method [[4]].

Our construction is quantitative in the sense that it allows us to track the dependence of the bounds on the weight constants. Such quantitative versions of extrapolation results were first formalized by Dragičević, Grafakos, Pereyra, and Petermichl in the linear setting in [[11]], but are completely new in the multilinear setting. In the linear setting this result is based on Buckley's sharp weighted bound for the Hardy–Littlewood maximal operator. This bound has been generalized to the multi-sublinear Hardy–Littlewood maximal operator by Damián, Lerner, and Pérez [[10]] to a sharp estimate in the setting of a mixed type Ap-A estimates and a sharp Ap bound is found in [[34]]. We give a different proof of this result for the limited range version of this maximal operator by generalizing a proof of Lerner [[28]].

Finally, we also show how our quantitative extrapolation result recovers and extends a bound obtained for multi-(sub)linear sparsely dominated operators, generalizing the bound of Hytönen's A2 Theorem [[20]]. More precisely, sparse domination yields sharp bounds for an operator for exponents p1,...,pm only if 1p=j=1m1pm1 so that we may appeal to duality. Our extrapolation result allows us to show that this same control in terms of the weight also holds when 1p>1 .

Symmetry in Muckenhoupt weight classes

To facilitate our results, we heavily rely on the symmetric structure of the Muckenhoupt classes.

For p(1,) , a standard method of obtaining weighted Lp estimates with a weight w is by using the duality (Lp(w))=Lp(w1-p) given through the integral pairing

f,g=Rnfgdx.

Graph

Moreover, the Muckenhoupt Ap class is defined through these two weights w and w1-p through

[w]Ap:=supQ1|Q|Qwdx1|Q|Qw1-pdxp-1

Graph

where the supremum is taken over all cubes QRn . One way to understand this definition better is by noting that we can relate the weights w and w1-p through w1p(w1-p)1p=1 . One can also make sense of this condition if p=1 through

[w]A1:=supQ1|Q|QwdxessinfxQw(x)-1,

Graph

and one usually defines A:=p[1,)Ap .

When we replace the weight w by the weight wp we find, using the averaging notation hq,Q:=1|Q|Q|h|qdx1q , that

[wp]Ap1p=supQwp,Qw-1p,Q

Graph

for p(1,) . The symmetry in this condition is much more prevalent and this condition seems to be more naturally adapted to the weighted Lp theory. Indeed, defining

[w]p:=[wp]Ap1p,

Graph

we note that [w]p=[w-1]p . If we denote the Hardy–Littlewood maximal operator by M and if we define the bi-sublinear Hardy–Littlewood maximal operator M(1,1) by

M(1,1)(f1,f2)(x):=supQxf11,Qf21,Q,

Graph

then we have the remarkable equivalences

1.3 M(1,1)Lp(wp)×Lp(w-p)L1,MLp(wp)Lp,(wp)MLp(w-p)Lp,(w-p)[w]p,

Graph

where the implicit constant depends only on the dimension, see Proposition 2.7 and Proposition 2.14 below.

Another way of thinking of these equivalences is by setting w1:=w , w2:=w-1 and p1:=p , p2:=p so that we have the relations

1.4 w1w2=1,1p1+1p2=1.

Graph

Then one can impose a symmetric weight condition

[(w1,w2)](p1,p2):=supQw1p1,Qw2p2,Q<

Graph

and note that

[(w1,w2)](p1,p2)=[w1]p1=[w2]p2.

Graph

The equivalences in (1.3) can now be thought of as

M(1,1)Lp1(w1p1)×Lp2(w2p2)L1,[(w1,w2)](p1,p2),MLp1(w1p1)Lp1,(w1p1)[w1]p1,MLp2(w2p2)Lp2,(w2p2)[w2]p2.

Graph

We can even make sense of these expressions when p1=1 and p2= or p1= and p2=1 , given that we use the correct interpretation and this is what allows us to extrapolate using such classes. Indeed, one can think of fLp(wp) as the condition fwLp< , which makes sense even when p= by requiring that the function fw is essentially bounded. Using the interpretation h,Q=esssupxQ|h(x)| , we see that the condition [w1]1< is equivalent to the usual A1 condition imposed on the weight w1=w , while the condition [w1]< is equivalent to the condition w2=w-1A1 . We emphasize here that our condition [w]< is not equivalent to the condition wA=p[1,)Ap and these notions should not be confused. The condition w-1A1 seems to be a natural upper endpoint condition and one can show that this is equivalent to the boundedness

(Mf)wLcfwL,

Graph

see Proposition 2.14 below. It also turns out that this condition allows us to extrapolate away from weighted L estimates. We point out that this idea has already been used in the endpoint extrapolation result of Harboure, Macías and Segovia [[19], Theorem 3].

We wish to view our symmetric weight condition in the context of extrapolation. In proving Rubio de Francia's extrapolation theorem, one usually starts with a pair of functions (h, f) and assumes that one has the inequality

1.5 hLq(wq)cfLq(wq)

Graph

for some q[1,] and all weights w satisfying [w]q< . The idea is then that given a p(1,) and a weight w satisfying [w]p< , one can construct a weight W, possibly depending on f, h, and w, so that W satisfies [W]q< as well as some additional properties to ensure that we can use (1.5) with W to conclude that

1.6 hLp(wp)c~fLp(wp).

Graph

Applying this with h=Tf then gives the desired boundedness for an operator T. For the proof one usually splits into two cases, namely the case where p<q and the case where p>q . In the former case one can apply Hölder's inequality to move from Lp to Lq and in the latter case one uses duality and a similar trick to move from Lp to Lq . The point is that both of these cases are essentially the same, but due to the notation we use we have to deal with the cases separately. Here, we wish to come up with a formalization to avoid this redundancy.

The extrapolation theorem is essentially a consequence of to the following proposition:

Proposition

Suppose we are given p1,p2(1,) satisfying 1p1+1p2=1 and weights w1,w2 satisfying w1w2=1 and [(w1,w2)](p1,p2)< . Moreover, assume we have two functions f1Lp1(w1p1) and f2Lp2(w2p2) and q1,q2[1,] with 1q1+1q2=1 . Then there are weights W1 , W2 satisfying W1W2=1 ,

f1W1Lq1f2W2Lq22f1w1Lp1f2w2Lp2

Graph

and

[(W1,W2)](q1,q2)C[(w1,w2)](p1,p2)maxp1q1,p2q2.

Graph

Indeed, the result of the extrapolation theorem follows by applying the proposition with f1:=f , q1:=q , q2:=q , p1:=p , p2:=p , w1:=w , w2=w-1 and W1:=W , W2:=W-1 so that, by (1.5), we have

|h,f2|hWLqf2W-1LqcfWLqf2W-1Lq2cfwLpf2w-1Lp.

Graph

Thus, by duality, we obtain (1.6), as desired.

The proof of the proposition uses the classical construction using the Rubio de Francia algorithm and the novelty here is our symmetric formulation. A proof can be found in this work, as it is a special case of Theorem 3.1. The case p<q in the proposition takes the form p1<q1 and p2>q2 while the case p>q takes the form p1>q1 and p2<q2 . The fact that the proposition is formulated completely symmetrically in terms of the parameters indexed over {1,2} , where we note that [(w1,w2)](p1,p2)=[(w2,w1)](p2,p1) , means that these respective cases can be proven using precisely the same argument, up to a permutation of the indices. Thus, without loss of generality, one only needs to prove one of the two cases.

These symmetries become especially important in the m-linear setting where we are dealing with parameters indexed over {1,...,m+1} and the amount of cases we have to consider increases. Thanks to our formulation, we will be able to reduce these multiple cases back to a single case in our arguments again by permuting the indices.

We wish to point out here that to facilitate our symmetric formulation and to use the duality argument involving the Rubio de Francia algorithm as above, we need to essentially restrict ourselves to the Banach range 1p1 . However, in the m-linear setting one also has to deal with the quasi-Banach range 1p>1 . This means that to employ our multilinear Rubio de Francia algorithm, we must first reduce to the case where j=1m1pj=1p1. In this case we can set 1pm+1:=1-1p0 and j=1m+11pj=1 , which places us in the setting of Theorem 3.1. This is not a problem however, as reducing to this case is facilitated by the rescaling properties of the multilinear weight classes, see also Remark 2.3. In conclusion, even though our multilinear Rubio de Francia algorithm is applied in the Banach range 1p1 , our result also includes the quasi-Banach range 1p>1 .

This article is organized as follows:

  • In Sect. 2 we state our main result and give an overview of the multilinear weight classes, proving some important properties as well as proving new quantitative estimates with respect to the multisublinear maximal operator as well as sparse forms.
  • In Sect. 3 we prove the main result.
  • In Sect. 4 we apply the extrapolation result for weak type bounds and certain BMO type bounds as well as for vector-valued bounds. Moreover, we give an application of our results to the bilinear Hilbert transform.
Multilinear weight classes

Setting and main result

We work in Rn equipped with the Lebesgue measure dx . This is mostly for notational convenience and our results also hold in the more general setting of spaces of doubling quasimetric measure spaces, provided one uses the right notion of dyadic cubes in this setting, see [[21]]. For a measurable set E we denote its Lebesgue measure by |E|. A measurable function w:Rn(0,) is called a weight. We can identify w with a measure by w(E):=Ewdx . For p(0,] , a weight w, and a measurable function f on Rn we say that fLp(wp) provided that fLp(wp):=fwLp< . Moreover, for a measurable set ERn with 0<|E|< we write

fp,E:=1|E|E|f|pdx1p

Graph

when 0<p< and f,E:=esssupxE|f(x)| .

We will use the notation AB if there is a constant c>0 , independent of the important parameters, such that AcB . Moreover, we write AB if AB and BA .

Let mN and let r1,...,rm(0,) , s(0,] . For p1,...,pm(0,] , writing r=(r1,...,rm) and similarly for p , we write rp if rjpj for all j{1,...,m} . Moreover, we write (r,s)p if rp and ps , where p is defined by

1p=j=1m1pj.

Graph

Similarly, we write r<p if rj<pj for all j{1,...,m} and we write (r,s)<p if r<p and p<s .

Definition 2.1

Let r1,...,rm(0,) , s(0,] , and p1,...,pm(0,] with (r,s)p . Let w1,...,wm be weights and write w=j=1mwj , w=(w1,...,wm) . We say that wAp,(r,s) if

[w]p,(r,s):=supQj=1mwj-111rj-1pj,Qw11p-1s,Q<,

Graph

where the supremum is taken over all cubes QRn .

As a point of comparison, we note here that, for finite pj , our condition wAp,(r,s) is equivalent to the condition (w1p1,...,wmpm)Ap,(r1,...,rm,s) , where the latter condition considers the weight class of Li, Martell, and Ombrosi defined in (1.2). Thus, in this range their extrapolation result [[33]] consider the same weights as we do.

Our main theorem is as follows:

Theorem 2.2

(Quantitative multilinear limited range extrapolation) Let (f1,...,fm,h) be an m+1 -tuple of measurable functions and let r1,...,rm(0,) , s(0,] . Suppose that for some q1,...,qm(0,] with q(r,s) there is an increasing function ϕq such that

2.1 hLq(wq)ϕq([w]q,(r,s))j=1mfjLqj(wjqj)

Graph

for all wAq,(r,s) .

Then for all p1...,pm(0,] with p>(r,s) there is an increasing function ϕp,q,r,s such that

2.2 hLp(wp)ϕp,q,r,s([w]p,(r,s))j=1mfjLpjwjpj

Graph

for all wAp,(r,s) . More explicitly, we can take

2.3 ϕp,q,r,s(t)=2m2rϕqCp,q,r,strmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s1r,

Graph

where 1r=j=1m1rj .

We note that if there is equality in one of the components in q(r,s) , i.e., if q=s or qj=rj for some j{1,...,m} , then we may also include the respective cases with p=s or pj=rj to the conclusion of the extrapolation result. In this case one should respectively use the interpretation 1q-1s1p-1s=1 or 1rj-1qj1rj-1pj=1 . To see this, one need only note that the proof we give of the theorem already accounts for the respective cases when 1p=1q or 1pj=1qj .

Our result is stronger than that in [[33]] in the sense that we do not have to restrict our exponents to the case where they are finite, i.e., in the initial assumption we include all the cases where qj= and in the conclusion we similarly obtain all the cases where pj= , see also [[32]]. We emphasize here that we use the interpretation fjLqj(wjqj)=fjwjL in the case where qj= and we need to impose the weight condition from Definition 2.1 with 1qj=0 . For example, in the case m=1 , r=1 and q=s= , one has to use the condition wA,(1,) in the initial estimate (2.1) which, following our definition, is equivalent to the condition w-1A1 . This stronger result is possible due to our use of a multilinear Rubio de Francia algorithm, fully utilizing the multilinear nature of the problem. Our result also implies vector valued estimates in these ranges and we refer the reader to Sect. 4 where we elaborate on this further.

Next we make some remarks on the quantitative result (2.3).

Usually in applications, the increasing function will be of the form ϕq(t)=ctα for some c,α>0 . Then we find from (2.3) that

ϕp,q,r,s(t)=c~tαmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s.

Graph

In the case m=1 , r=1 , s= , this means that we have

2.4 ϕp,q,1,(t)=c~tαmaxpq,pq,

Graph

and this coincides with the bound obtained in [[11]]. This result was used in Hytönen's A2 theorem [[20]] to reduce proving the sharp estimate

2.5 TfLp(wp)[wp]Apmax(p,p)pfLp(wp)

Graph

for Calderón-Zygmund operators T to only having to prove the linear A2 bound

TfL2(w2)[w2]A2fL2(w2).

Graph

Indeed, noting that [w]p,(1,)=[wp]Ap1p , we find that (2.5) follows from (2.4) by taking α=2 and q=2 .

The fact that we need to extrapolate from q=2 to obtain the sharp bounds for Calderón-Zygmund operators speaks to their nature as operators revolving around their properties in L2 . As a contrast, we note that the estimate

(Tf)wL[w-1]A1fwL

Graph

is central, for example, for when T is the Hardy–Littlewood maximal operator M. Indeed, by (2.4) with q= and α=1 and by noting that [w],(1,)=[w-1]A1 , this estimate extrapolates to the estimate

TfLp[wp]ApppfLp(wp)

Graph

for p(1,] , which is precisely Buckley's sharp bound obtained for M. We point out here that this argument is actually circular for when T=M , since the proof of the quantitative estimate in the extrapolation result makes use of Buckley's sharp bound. Nonetheless, we think this example is heuristically interesting, since it exhibits how one can extrapolate away from weighted L estimates. Multilinear versions of Buckley's sharp bound have been found in [[10], [34]] and can be recovered in a similar way, see also Theorem 4.12.

The remainder of this section will be dedicated to a discussion on the quantitative properties of the multilinear weight classes. We split this into two separate cases. In the first case we adopt the symmetric notation from the introduction and think in terms of m+1 -tuples of weights and parameters satisfying a symmetric relation. In the second case we adopt the more classical approach of thinking in terms of m-tuples and we prove some key results for our main theorem.

Quantitative properties of multilinear weight classes: the m+1-tuple case

Let r1,...,rm(0,) , s(0,] , and p1,...,pm(0,] and let w1,...,wm be weights, with w:=j=1mwj . In terms of symmetries, the definition of the weight class

[w]p,(r,s)=supQj=1mwj-111rj-1pj,Qw11p-1s,Q

Graph

seems to be best suited to the case where 1p1 . Indeed, if we set 1pm+1:=1-1p0 , 1rm+1:=1-1s and wm+1:=w-1 , then we have

j=1m+11pj=1,j=1m+1wj=1.

Graph

The condition (r,s)p is equivalent to rjpj for all j{1,...,m+1} and the constant for the weight class now takes the form

[w]p,(r,s)=supQj=1m+1wj-111rj-1pj,Q=[(w1,...,wm+1)](p1,...,pm+1),((r1,...,rm+1),),

Graph

where the last equality follows from the fact that the term involving the product weight in the m+1 -linear weight class is equal to 1. The symmetry of this last expression also emphasizes a certain permutational invariance. Indeed, if πSm+1 is a permutation, then, since

j=1m+11pπ(j)=j=1m+11pj=1,j=1m+1wπ(j)=j=1m+1wj=1,

Graph

we have

[w]p,(r,s)=[(wπ(1),...,wπ(m))](pπ(1),...,pπ(m)),((rπ(1),...,rπ(m)),rπ(m+1)),

Graph

and this will be used in the proof of our extrapolation theorem.

Remark 2.3

While we restrict ourselves to the Banach range 1p1 in this section, we do point out that our main results do also apply in the cases where 1p>1 . This is facilitated by the rescaling property

[w]pα,(rα,sα)1α=w11α,...,wm1αp,(r,s),

Graph

which, in our arguments, allows us to reduce back to the case where 1p1 , see also the proof of Theorem 2.2.

It will sometimes also be useful to redefine vj:=wj-11rj-1pj for j{1,...,m+1} so that

[(w1,...,wm+1)](p1,...,pm+1),((r1,...,rm+1),)=supQj=1m+1vj1,Q1rj-1pj.

Graph

These weight classes are governed by a certain maximal operator, see also [[30]].

Definition 2.4

Given r1,...,rm(0,) , we define the m-sublinear Hardy–Littlewood maximal operator

Mr(f1,...,fm)(x):=supQxj=1mfjrj,Q

Graph

for fjLlocrj , where the supremum is taken over all cubes QRn containing x. Moreover, for a dyadic grid D we define

MrD(f1,...,fm)(x):=supQxQDj=1mfjrj,Q

Graph

for fjLlocrj .

For the relevant definitions and results regarding dyadic grids we refer the reader to [[29]]. A property we need is the fact that there exist 3n dyadic grids (Dα)α=13n such that for each cube QRn there is an α and a cube Q~Dα such that QQ~ and |Q~|6n|Q| . This implies the following:

Lemma 2.5

Let r1,...,rm(0,) . Then there exist 3n dyadic grids (Dα)α=13n such that

Mrα=13nMrDα.

Graph

See also [[29]].

Definition 2.6

A collection of cubes S in a dyadic grid is called sparse if there is a pairwise disjoint collection of measurable sets (EQ)QS such that EQQ and |Q|2|EQ| .

Given r1,...,rm(0,) , for a sparse collection of cubes S we define the sparse operator

Ar,S(f1,...,fm):=QSj=1mfjrj,QχQ,

Graph

and the sparse form

Λr,S(f1,...,fm):=QSj=1mfjrj,Q|Q|.

Graph

We point out here that the sparsity constant 2 appearing in the estimate |Q|2|EQ| is not too important and in most situations it can be replaced by any other constant greater than 1. Note however, that we will be considering the form supSΛr,S and here it is important that one only considers sparse collections in this supremum with the same sparsity constant. See [[29]] for further properties and results regarding sparse collections of cubes.

Since this section contains results involving both m-tuples and m+1 -tuples with the same parameters, it is convenient to separate these notationally. We will use the following convention: for m+1 parameters α1,...,αm+1 we shall use the boldface notation α=(α1,...,αm+1) for m+1 -tuples while we will use the arrow notation α=(α1,...,αm) for m-tuples.

The main result for this section is the following:

Proposition 2.7

Let r1,...,rm+1(0,) , p1,...,pm+1(0,] satisfy 1pj<1rj for all j{1,...,m+1} and j=1m+11pj=1 . Moreover, let w1,...,wm+1 be weights satisfying j=1m+1wj=1 . Then the following are equivalent:

  • wAp,(r,) ;
  • MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,< ;
  • MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1< ;
  • supSΛr,SLp1(w1p1)××Lpm+1(wm+1pm+1)R< .

Moreover, we have

2.6 MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,[w]p,(r,),

Graph

2.7 MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1supSΛr,SLp1(w1p1)××Lpm+1(wm+1pm+1)R

Graph

where the implicit constants depend only on the dimension, and

2.8 supSΛr,SLp1(w1p1)××Lpm+1(wm+1pm+1)Rcp,r[w]p,(r,)maxj=1,...,m+11rj1rj-1pj,

Graph

where the implicit constant depends on the dimension and

cp,r=j=1m+11rj1rj-1pj1rj.

Graph

Remark 2.8

We again point out that the condition wAp,(r,) is equivalent to the condition wAp,(r,rm+1) , with equal constants. Moreover, the results containing the sparse forms are formulated with the supremum taken inside of the norm. One can equivalently put the supremum outside of the norm which follows from the fact that there is a single sparse form that dominates all the other sparse forms, see [[27], Section 4].

In the case m=1 , r1=r2=1 , the equivalence (2.6) takes the more familiar form

M(1,1)Lp(wp)×Lp(w-p)L1,[wp]Ap1p

Graph

which appeared in the introduction.

We note that the estimate (2.8) was already obtained in [[8]] in the case m=2 .

For r1=r , r2=s the estimate (2.8) takes the form

2.9 supSΛ(r,s),SLp(wp)×Lp(w-p)Rw1p-1s-1A1r-1s1p-1smax1p-1s1r-1p1r,1s

Graph

and when r=1 and s= we reobtain the sharp bound from the A2 theorem. We wish to compare (2.9) to the bound obtained in [[5]]. For their main result they prove that

2.10 supSΛ(r,s),SLp(w)×Lp(w1-p)R([w]Apr[w]RHsp)max1p-r,s-1s-p,

Graph

Our result implies that

supSΛ(r,s),SLp(w)×Lp(w1-p)R([w]Apr[w]RHsp)spmax1p-1s1r-1p1r,1s,

Graph

see also [[24]], and this recovers the estimate (2.10).

Finally, we point out here that the estimate (2.8) already appears in [[33], p. 12] for the particular choice 1pj=1rj1j=1m+11rj , and it seems like this choice of pj is central for the theory of these sparse forms, see also the proof of Corollary 4.2.

For the proof of the proposition we will require several preparatory lemmata.

Lemma 2.9

Let 0<r1,...,rm< . Then for each dyadic grid D and all fjLrj there is a sparse collection SD such that

MrD(f1,...,fm)2n+1rQSj=1mfjrj,QχEQ

Graph

pointwise almost everywhere, where 1r=j=1m1rj . In particular we have

MrD(f1,...,fm)2n+1rAr,S(f1,...,fm)

Graph

pointwise almost everywhere.

The proof is essentially the same as the well-known result in the case m=1 , r=1 .

Proof

For kZ we define

Ωk:={xRn:MrD(f1,...,fm)(x)>2n+1rk}.

Graph

By taking the maximal cubes Q in Ωk we obtain a pairwise disjoint collection QkD such that Ωk=QQkQ and

2.11 2n+1rk<j=1mfjrj,Q2n+1r(k+1)21r

Graph

for all QQk . We define S:=kZQk and claim that S is a sparse collection of cubes. Indeed, for QQk it follows from (2.11) that for any QQk+1 we have

j=1mfjrj,Q>21r2n+1r(k+1)21r21rj=1mfjrj,Q.

Graph

Thus, by maximality of Qk and Hölder's inequality with j=1mrrj=1 , we have

|Ωk+1Q|=QQk+1QQ|Q|12j=1mfjrj,QrQQk+1QQ|Q|j=1mfjrj,Qr=|Q|2QQk+1QQj=1mQ|fj|rjdxrrjj=1mQ|fj|rjdxrrj|Q|2j=1mΩk+1Q|fj|rjdxrrjj=1mQ|fj|rjdxrrj|Q|2.

Graph

Thus, defining EQ:=Q\Ωk+1 , we have |Q|2|EQ| .

To conclude that S is sparse, it remains to check that (EQ)QS is pairwise disjoint. Let Q,QS such that EQEQ . If QQk and QQk , we have EQΩk\Ωk+1 and EQΩk\Ωk+1 . Since (Ωk\Ωk+1)kZ is pairwise disjoint, this means that we must have k=k . Since QQ , it follows from maximality of Qk that Q=Q , as desired.

Finally, if xRn and MrD(f1,...,fm)(x)0 , then there is a unique kZ such that 2n+1rk<MrD(f1,...,fm)(x)2n+1r(k+1) . Hence, xΩk\Ωk+1 and thus there is a cube QQk so that xQ\Ωk+1=EQ and

MrD(f1,...,fm)(x)2n+1r2n+1rk<2n+1rj=1mfjrj,Q=2n+1rQSj=1mfjrj,QχEQ(x).

Graph

This proves the assertion.

The following result is a reformulation of the definition of the weight class.

Lemma 2.10

Let r1,...,rm+1(0,) , p1,...,pm+1(0,] satisfy 1pj<1rj for all j{1,...,m+1} and j=1m+11pj=1 . Moreover, let w1,...,wm+1 be weights satisfying j=1m+1wj=1 and define vj:=wj-11rj-1pj . Then wAp,(r,) if and only if v1,...,vm+1 are locally integrable and there is a constant c>0 such that for all cubes Q we have

j=1m+1vj1,Q1rj|Q|cj=1m+1vj(Q)1pj.

Graph

In this case, the optimal constant c in this inequality is given by [w]p,(r,) .

The following lemma allows us to deal with weighted estimates involving sparse forms.

Lemma 2.11

Let r1,...,rm+1(0,) , p1,...,pm+1(0,] satisfy 1pj<1rj for all j{1,...,m+1} and j=1m+11pj=1 . Moreover, let w1,...,wm+1 be weights satisfying j=1m+1wj=1 with wAp,(r,) and define vj:=wj-11rj-1pj . Let Q be a cube and let EQ such that |Q|2|E| . Then

2.12 j=1m+1vj1,Q1rj|Q|[w]p,(r,)maxj=1,...,m+11rj1rj-1pjj=1m+1vj(E)1pj.

Graph

Remark 2.12

Having Lemma 2.10 in mind, it seems that the larger power of the weight constant in (2.12) comes from the fact that we are passing from the weighted measure of the set Q to the measure of the smaller set E. In fact, it seems like we are only using the full weight condition wAp,(r,) once and we are left with an estimate of the form

j=1m+1vj(Q)1pjj=1m+1vj(E)1pj,

Graph

where the implicit constant depends on the weights. This estimate seems to only require the weaker Fujii-Wilson A condition satisfied by the weight vj , but we do not pursue this further here. We refer the reader to [[22]] where quantitative estimates involving this condition first appeared. We also point out that estimates of this type for the limited range sparse operator in the case m=1 have been studied in [[13], [31]]. This condition has also been considered in the multilinear case in [[10]].

Proof

We set γ:=maxj=1,...,m+11rj1rj-1pj and

βj:=1rj-1rj-1pjγ,

Graph

so that βj0 for all j{1,...,m+1} . Thus, since vj1,E2vj1,Q by the assumptions on E, we have vj1,Qβj2-βjvj1,Eβj . Then

2.13 j=1m+1vj1,Q1rj|Q|=j=1m+1vj1,Q1rj-1pjγj=1m+1vj1,Qβj|Q|[w]p,(r,)γj=1m+1vj1,Qβj|Q|[w]p,(r,)γj=1m+1vj1,Eβj|E|=[w]p,(r,)γj=1m+1vj(E)βj|E|1-j=1m+1βj.

Graph

Next, set α:=j=1m+11rj-1pj>0 and kj:=α1rj-1pj-1 . Then

j=1m+11kj=1αj=1m+11rj-1pj=1

Graph

and

1-j=1m+1βj=j=1m+11pj-j=1m+11rj+γj=1m+11rj-1pj=(γ-1)α

Graph

so that

1-j=1m+1βjkj=1rj-1pj(γ-1)=1pj-βj.

Graph

Thus, since j=1m+1vj1rj-1pj=j=1m+1wj=1 , it follows from Hölder's inequality that

|E|1-j=1m+1βj=Ej=1m+1vj1α1rj-1pjdx1-j=1m+1βjj=1m+1vj(E)1-j=1m+1βjkj=j=1m+1vj(E)1pj-βj.

Graph

By combining this estimate with (2.13), we obtain (2.12). The assertion follows.

Proof of Proposition 2.7

We set vj:=wj-11rj-1pj for j{1,...,m+1} .

The strategy for the proof will be as follows: We will prove the equivalence of (i) and (ii) by proving (2.6) and we will prove the equivalence of (iii) and (iv) by proving (2.7). Then, noting that the implication (iii) (ii) is clear, we conclude the proof by showing that (i) (iv) through (2.8).

For (2.6), for the first inequality we note that it follows from Lemma 2.5 that it suffices to consider the estimate for MrD for a dyadic grid D . First consider a finite collection FD . Let λ>0 , fjLpj(wjpj) and, defining MrF as MrD but with the supremum taken over all QF , we set

ΩλF:=MrF(f1,...,fm+1)>λ

Graph

and similarly for ΩλD .

Let P denote the collection of those cubes QF such that j=1m+1fjrj,Q>λ that have no dyadic ancestors in F . Using the rule

hr,Q=hu-1rr,Quu1,Q1r,

Graph

where hr,Qu:=1u(Q)Q|h|rudx1r , it follows from Lemma 2.10 and the fact that P gives a decomposition of ΩλF , that

λ|ΩλF|=QPλ|Q|QPj=1m+1fjrj,Q|Q|=QPj=1m+1fjvj-1rjrj,Qvjvj1,Q1rj|Q|[w]p,(r,)QPj=1m+1fjvj-1rjrj,Qvjvj(Q)1pj[w]p,(r,)QPj=1m+1Q|fj|pjvjpj1pj-1rjdx1pj[w]p,(r,)j=1m+1fjLpjwjpj,

Graph

where in the fourth step we used Hölder's inequality with rjpj and in the last step we used Hölder's inequality on the sum.

By considering an exhaustion of D of finite sets it follows from monotonicity of the measure and by taking a supremum over λ>0 that

MrDLp1(w1p1)××Lpm+1(wm+1pm+1)L1,[w]p,(r,).

Graph

For the converse inequality, fix a cube Q. Assuming for the moment that the vj are locally integrable, we let 0<λ<j=1mvj1,Q1rj . Setting fj:=vj1rjχQ , we obtain

Mr(f1,...,fm+1)(x)j=1m+1fjrj,Q=j=1m+1vj1,Q1rj>λ

Graph

for all xQ so that Q{Mr(f1,...,fm+1)>λ} . Thus,

λ|Q|λ|{Mr(f1,...,fm+1)>λ}|MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,j=1m+1fjLpjwjpj=MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,j=1m+1vj(Q)1pj.

Graph

Taking a supremum over such λ , we conclude that

2.14 j=1m+1vj1,Q1rj|Q|MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,j=1m+1vj(Q)1pj.

Graph

Thus, it follows from Lemma 2.10 that

[w]p,(r,)MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1,,

Graph

proving (2.6). To prove our initial assumption that the vj are locally integrable, we repeat the above argument with the weights replaced by (vj-1+ε)-1 for ε>0 . As these weights are bounded, they are locally integrable. An appeal to the Monotone Convergence Theorem as ε0 after a rearrangement of (2.14) yields the desired conclusion.

For (2.7), let fjLpj(wjpj) and let D be a dyadic grid. By Lemma 2.9 there exists a sparse collection SD such that

MrD(f1,...,fm+1)L1Ar,S(f1,...,fm+1)L1Λr,S(f1,...,fm+1).

Graph

Thus, it follows from Lemma 2.5 that

MrLp1(w1p1)××Lpm+1(wm+1pm+1)L1supSΛr,SLp1(w1p1)××Lpm+1(wm+1pm+1)R.

Graph

For the converse inequality, we estimate

Λr,S(f1,...,fm+1)2QSj=1m+1fjrj,Q|EQ|2QSEQMr(f1,...,fm+1)dx2Mr(f1,...,fm,g)L1.

Graph

As this estimate is uniform in S , this proves (2.7) and thus the equivalence of (iii) and (iv).

To prove (2.8) and thus the implication (i) (iv), we note that it follows from Lemma 2.11 that for a sparse collection S in a dyadic grid D and for γ=maxj=1,...,m+11rj1rj-1pj we have

Λr,S(f1,...,fm+1)=QSj=1m+1fjrj,Q|Q|=QSj=1m+1fjvj-1rjrj,Qvjvj1,Q1rj|Q|[w]p,(r,)γQSj=1m+1fjvj-1rjrj,Qvjvj(EQ)1pj[w]p,(r,)γQPj=1m+1EQMrjvj,D(fjvj-1rj)pjvjdx1pj[w]p,(r,)γj=1m+1Mrjvj,D(fjvj-1rj)Lpj(vj)cp,r[w]p,(r,)γj=1m+1fjLpjwjpj,

Graph

where in the last step we used the fact that the weighted dyadic maximal operator Mru,Dh:=supQDhr,QuχQ is bounded on Lq(u) for q>r with constant bounded by 1r1r-1q1r , uniformly in the weight u. As this estimate is uniform in the sparse collection S , this proves (2.8). The assertion follows.

Quantitative properties of multilinear weight classes: the m-tuple case

It is sometimes convenient to emphasize this separation of the parameter s from the rj , as it often plays a different role from the other parameters in the proofs. The following lemma provides a way to deal with this parameter.

Lemma 2.13

(Translation lemma) Let r1,...,rm(0,) , s(0,] and p1,...,pm(0,] with (r,s)p and let w1,...,wm be weights with w=j=1mwj . Then wAp,(r,s) if and only if there are 1s1,...1sm satisfying 1sj1pj , j=1m1sj=1s , and wAp(s),(r(s),) , where

p(s)=11p1-1s1,...,11pm-1sm,r(s)=11r1-1s1,...,11rm-1sm.

Graph

Moreover, in this case we have

2.15 [w]p,(r,s)=[w]p(s),(r(s),).

Graph

Proof

We have

1p(s):=j=1m1pj-1sj=1p-1s.

Graph

it remains to note that

j=1mwj-111rj-1pj,Qw11p-1s,Q=j=1mwj-111rj-1sj-1pj-1sj,Qwp(s),Q.

Graph

Taking a supremum over all cubes Q yields (2.15), proving the assertion.

We point out that the choice of the 1sj in the lemma is not necessarily unique if m1 . One could, for example, take 1sj=ppj1s , but a different choice will be made later in the proof of the main result. We also note that this lemma can be used even if 1s=0 . In this case it can occur that some of the 1sj are negative, but this does not seem to cause any problems.

Having reduced to the case where s= , the following proposition is the main result for this subsection.

Proposition 2.14

Let r1,...,rm(0,) , p1,...,pm(0,] with (r,)p and let w1,...,wm be weights with w=j=1mwj . Then the following are equivalent:

  • wAp,(r,) ;
  • MrLp1(w1p1)××Lpm(wmpm)Lp,(wp)< .

In this case we have

2.16 MrLp1(w1p1)××Lpm(wmpm)Lp,(wp)[w]p,(r,).

Graph

Moreover, if r<p , then (i) and (ii) are equivalent to

  • MrLp1(w1p1)××Lpm(wmpm)Lp(wp)<

and we have

2.17 MrLp1(w1p1)××Lpm(wmpm)Lp(wp)cp,r[w]p,(r,)maxj=1,...,m1rj1rj-1pj,

Graph

where the implicit constant depends on the dimension and

cp,r=j=1m1rj1rj-1pj1rj.

Graph

Moreover, the power of the weight constant in (2.17) is the smallest possible one.

Remark 2.15

The equivalence (2.16) is also contained in the full range version in [[30], Theorem 3.3] in the case where the pj are finite, and the limited range version is proven in [[3], Proposition 21], but here the cases pj= are only treated when wj=1 .

For our result here we use the interpretation that for q= and a weight u we have hLq(uq)=hLq,(uq)=huL .

Remark 2.16

The estimate (2.17) is a generalization of Buckley's sharp weighted bound for the Hardy–Littlewood maximal operator. It can be proven using the sparse domination we obtained in Lemma 2.9, but we present an altogether different proof which generalizes an approach due to Lerner [[28]]. This construction is important, as it turns out to be key for our multilinear Rubio de Francia algorithm.

In the case r1==rm=1 , the sharp bound (2.17) recovers the sharp bound obtained by Li, Moen, and Sun in [[34]] where sparse domination techniques were used. To see this, note given weights w1,...,wm and setting vw:=j=1mwjppj , the multilinear weight constant they used is defined as

2.18 [w]Ap:=supQ1|Q|Qvwdxj=1m1|Q|Qwj1-pjdxppj.

Graph

Writing 1=(1,...,1) , the sharp result they prove is

2.19 M1Lp1(w1)××Lpm(wm)Lp(vw)[w]Apmaxj=1,...,mpjp.

Graph

for all p1,...,pm(1,) and wAp . To compare this to our result, we replace the wj by wjpj and note that vw=j=1m(wjpj)ppj=wp , [(w1p1,...,wmpm)]Ap=[w]p,(1,)p. Thus, (2.19) coincides with our bound found in (2.17) when r=1 .

Lemma 2.17

Let r1,...,rm(0,) , p1,...,pm(0,] with r<p and let w1,...,wm be weights with w=j=1mwj and wAp,(r,) . Then there exist sublinear operators Npj,rj,w:Lpj(wjpj)Lpj(wjpj) so that for any fjLpj(wjpj) we have

2.20 Mr(f1,...,fm)[w]p,(r,)maxj=1,...,m1rj1rj-1pjj=1mNpj,rj,w(fj).

Graph

Moreover, Npj,rj,w satisfies

Npj,rj,wLpjwjpjLpjwjpj1rj1rj-1pj1rj.

Graph

Proof

We first prove this result for the dyadic maximal operator MrD for a dyadic grid D to obtain the appropriate operators Npj,rj,wD . Then it follows from Lemma 2.5 that

2.21 Mr(f)α=13nj=1mNpj,rj,wDα(fj)j=1mα=13nNpj,rj,wDα(fj).

Graph

The result then follows by setting

Npj,rj,w:=cα=13nNpj,rj,wDα,

Graph

where c is an appropriate constant determined by the implicit constant in (2.21).

Now, fix a dyadic grid D . Let γ:=maxj=1,...,m1rj1rj-1pj , let QD , and set vj:=wj-11rj-1pj . Since j=1mwj-1w1pj1p=j=1mwj-1w=1 , it follows from Hölder's inequality that

1=11j=1m1rj,Qγ-1j=1mwj-1w1pj1prj,Qγ-1=j=1mwj-1w1pj1prj,Qγ-1rj1rj-1pjwj-1w1pj1prj,Q1pj1rj-1pjj=1mwj-111rj-1pjw1pj1ppj,Qγ-1rj1rj-1pjwj-1w1pj1prj,Q1pj1rj-1pj=j=1mvj1,Q1rj-1pjwp1,Q1pjγ-1rj1rj-1pjj=1mwj-rjw1pj1prj1,Q1pj1rj1rj-1pj.

Graph

This implies that

j=1mvj1,Q1rj[w]p,(r,)γj=1mvj1,Q1rj-1pjγ-1rjwp,Qγ=[w]p,(r,)γj=1mvj1,Q1rj-1pjwp1,Q1pjγ-1rj1rj-1pjj=1m1wp1,Q1pj1rj1rj-1pj[w]p,(r,)γj=1mwj-rjw1pj1prj1,Qwp1,Q1pj1rj1rj-1pj.

Graph

Thus, for fjLpj(wjpj) and any xQ , we have

2.22 j=1mfjrj,Q=j=1mfjvj-1rjrj,Qvjvj1,Q1rj[w]p,(r,)γj=1minfyQMrjvj,Dfjvj-1rj(y)1rj-1pj1pj1rjwj-rjw1pj1prj1,Qwp1,Q1pj1rj1rj-1pj[w]p,(r,)γj=1mM1rj-1pj1pj1rjwp,DMrjvj,Dfjvj-1rjvj1pjw-1pj1p(x).

Graph

Setting

Npj,rj,wD(fj):=M1rj-1pj1pj1rjwp,DMrjvj,Dfjvj-1rjvj1pjw-1pj1pw1pj1pwj-1

Graph

and by taking a supremum over all Q containing x in (2.22) we have proven (2.20) in the dyadic case. We remark here that in the case that 1pj=0 , we use the interpretation

N,rj,wD(fj)=Mrjvj,Dfjvj-1rjLwj-1.

Graph

Noting that

M1r-1q1q1ru,D(h)Lq(u)qr1q1r1r-1qhLq(u)=elogq-logrq-rhLq(u)e1rhLq(u),

Graph

for the case 1pj>0 , we compute

Npj,rj,wD(fj)Lpjwjpj=M1rj-1pj1pj1rjwp,DMrjvj,Dfjvj-1rjvj1pjw-ppjLpj(wp)Mrjvj,Dfjvj-1rjvj1pjw-ppjLpj(wp)=Mrjvj,Dfjvj-1rjLpj(vj)1rj1rj-1pj1rjfjvj-1rjLpj(vj)=1rj1rj-1pj1rjfjLpjwjpj,

Graph

and for the case 1pj=0 , we compute

N,rj,wD(fj)wjL=Mrjvj,D(fjvj-1rj)Lfjvj-1rjL=fjwjL.

Graph

The assertion follows.

Proof of Proposition 2.14

We will prove the equivalence of (i) and (ii) by proving (2.16).

For '' , we note that it follows from Lemma 2.5 that it suffices to prove the estimate for MrD for a fixed dyadic grid D . Note that by Hölder's inequality we have fjrj,Qfjwjpj,Qwj-111rj-1pj,Q for a cube Q, so that

j=1mfjrj,Q[w]p,(r,)wp,Q-1j=1mfjwjpj,Q=[w]p,(r,)j=1mfjwjw-ppjpj,Qwp.

Graph

Thus, by Hölder's inequality for weak type Lebesgue spaces, we have

MrD(f1,...,fm)Lp,(wp)[w]p,(r,)j=1mMpjwp,Dfjwjw-ppjLp,(wp)[w]p,(r,)j=1mMpjwp,Dfjwjw-ppjLpj,(wp)[w]p,(r,)j=1mfjLpjwjpj,

Graph

where we used the fact that the weighted dyadic maximal operator Mqu,D is bounded Lq(u)Lq,(u) with constant uniform in q and the weight u. Thus, we have shown that

MrLp1(w1p1)××Lpm(wmpm)Lp,(wp)[w]p,(r,).

Graph

For the converse inequality, fix a cube Q and let fjLpj(wjpj) . Letting 0<λ<j=1mfjrj,Q , we have

Mr(f1,...,fm)(x)j=1mfjrj,Q>λ

Graph

for all xQ so that Q{Mr(f1,...,fm)>λ} . Hence,

λwp,Q|Q|-1pλ(wp({Mr(f1,...,fm)>λ}))1pMrLp1(w1p1)××Lpm(wmpm)Lp,(wp)j=1m|Q|-1pjfjLpj(wpj).

Graph

Taking a supremum over such λ and by replacing fj with χQfj , we conclude that

2.23 j=1mfjrj,Qwp,QMrLp1(w1p1)××Lpm(wmpm)Lp,(wp)j=1mfjwjpj,Q.

Graph

Now set fj=wj-1rj1rj-1pj and assume for the moment that fjrj=fjpjwjpj=wj-11rj-1pj is locally integrable. Then the product on the right-hand side of (2.23) is positive and finite so that we may take it to the left-hand side. This yields

2.24 j=1mwj-111rj-1pj,Qwp,QMrLp1(w1p1)××Lpm(wmpm)Lp,(wp)

Graph

and taking a supremum over all cubes Q yields (2.16). To prove that wj-11rj-1pj is indeed locally integrable, we choose fj such that fjpjwjpj=(wj11rj-1pj+ε)-1 for ε>0 , the latter expression being bounded and thus locally integrable. Again taking the product on the right-hand side of (2.23) to the left, an appeal to the Monotone Convergence Theorem as ε0 yields (2.24). The assertion follows.

Since the implication (iii) (ii) is clear, we may finish the proof of the equivalences by showing (i) (iii) through (2.17).

By Lemma 2.17, it follows from Hölder's inequality that

Mr(f1,...,fm)Lp(wp)[w]p,(r,)maxj=1,...,m1rj1rj-1pjj=1mNpj,rj,wfjLpjwjpjcp,r[w]p,(r,)maxj=1,...,m1rj1rj-1pjj=1mfjLpjwjpj,

Graph

as desired.

Finally, we prove optimality of (2.17). Let α0 denote the smallest possible constant in the estimate

Mr(f1,...,fm)Lp(wp)[w]p,(r,)αj=1mfjLpjwjpj.

Graph

We have shown that αmaxj=1,...,m1rj1rj-1pj and it remains to prove the lower bound. We assume that we are in dimension n=1 , the general case following mutatis mutandis. Moreover, we assume without loss of generality that the maximum maxj=1,...,m1rj1rj-1pj is attained for j=1 , the other cases following similarly by permuting the indices. For 0<ε<1 we define

w1(x):=|x|(1-ε)1r1-1p1,wj(x):=1forj{2,...,m},f1(x):=|x|-1-εr1χ(0,1)(x),fj(x):=|x|-1-εpjχ(0,1)(x)forj{2,...,m}.

Graph

Then, by Hölder's inequality and a computation, we have

[w]p,(r,)[w1]p1,(r1,)ε1p1-1r1.

Graph

Moreover, one computes

j=1mfjLpjwjpj=ε-1p

Graph

and

j=1mfjrj,[-|x|,|x|]ε-1r1f1(x)j=2m1rj1rj-(1-ε)1pj1rjfj(x).

Graph

Setting f(x):=j=1mfj(x)wj(x)=|x|-1-εpχ(0,1)(x) , we find that

Mr(f1,...,fm)Lp(wp)ε-1r1fLp=ε-1r1-1p

Graph

and

Mr(f1,...,fm)Lp(wp)[w]p,(r,)αj=1mfjLpjwjpjεα1p1-1r1-1p

Graph

Letting ε0 shows that we must have α1p1-1r1-1p-(-1r1-1p)0 , i.e.,

α1r11r1-1p1=maxj=1,...,m1rj1rj-1pj.

Graph

Thus, we have α=maxj=1,...,m1rj1rj-1pj and the assertion follows.

Remark 2.18

We point out here that in the unweighted case we actually have an equivalence MrLp1××LpmLpcp,r , which follows from a similar calculation as above, with fj(x):=|x|-1-εpjχ(0,1)(x) for all j{1,...,m} .

Proof of the main result

The proof of the main theorem essentially follows from the theorem below. In this theorem we deal with m+1 -tuples as well as m-tuples of the same parameters, which can be notationally confusing. To circumvent this problem, we shall use the earlier established convention that for m+1 parameters α1,...,αm+1 we shall use the boldface notation α=(α1,...,αm+1) for m+1 -tuples while we will use the arrow notation α=(α1,...,αm) for m-tuples, see also Sect. 2.2.

We again point out that even though this result is formulated for the Banach range 1p1 , it can be used to obtain results in the range including the cases 1p>1 , see also Remark 2.3 and the proof of Theorem 2.2.

Theorem 3.1

Let 1r1,...,1rm+1(0,1] and suppose we are given 1p1,...,1pm+1[0,1] satisfying 1pj<1rj for all j{1,...,m+1} and j=1m+11pj=1 . Assume moreover that we are given weights w1,...wm+1 satisfying j=1m+1wj=1 and wAp,(r,) .

Suppose we are given functions fjLpj(wjpj) and 1q1,...,1qm+1[0,1] satisfying 1qj1rj and j=1m+11qj=1 . Then there are weights W1,...,Wm+1 satisfying j=1m+1Wj=1 and WAq,(r,) such that

3.1 j=1m+1fjLqj(Wjqj)2m2j=1m+1fjLpjwjpj

Graph

and

3.2 [W]q,(r,)Cp,q,r[w]p,(r,)maxj=1,...,m+11rj-1qj1rj-1pj.

Graph

The proof of this theorem relies on a multilinear generalization of the Rubio de Francia algorithm.

Lemma 3.2

(Multilinear Rubio de Francia algorithm) Let r1,...,rm,p1,...,pm(0,) with r<p . Then for each wAp,(r,) there exist operators Rpj,rj,w:LpjwjpjLpjwjpj satisfying

  • |fj|Rpj,rj,wfj ;
  • Rpj,rj,wfjLpj(wjpj)2fjLpj(wjpj) ;
  • j=1mRpj,rj,wfjrj,Qcp,r[w]p,(r,)maxj=1,...,m1rj1rj-1pjinfyQj=1mRpj,rj,wfj(y) for all cubes Q, where the implicit constant depends on the dimension and
  • cp,r=j=1m1rj1rj-1pj1rj.

Graph

Proof

Letting Npj,rj,w be as in Lemma 2.17, we define

Rpj,rj,wfj:=k=0Npj,rj,wk(fj)2kNpj,rj,wLpjwjpjLpjwjpjk,

Graph

where Npj,rj,w0(fj):=|fj| and Npj,rj,wk(fj):=Npj,rj,w(Npj,rj,wk-1(fj)) .

To prove property (i), it suffices to note that the k=0 term in the sum is equal to |fj| .

For (ii) we have

Rpj,rj,wfjLpjwjpjk=0Npj,rj,wk(fj)Lpjwjpj2kNpj,rj,wLpjwjpjLpjwjpjkk=0fjLpjwjpj2k=2fjLpjwjpj.

Graph

To prove (iii), we first note that

Npj,rj,w(Rpj,rj,wfj)k=0Npj,rj,wk+1(fj)2kNpj,rj,wLpjwjpjLpjwjpjk2Npj,rj,wLpjwjpjLpjwjpjRpj,rj,wfj.

Graph

Thus, it follows from Lemma 2.17 that

Mr(Rp1,r1,wf1,...,Rpm,rm,wfm)[w]p,(r,)maxj=1,...,m1rj1rj-1pjj=1mNpj,rj,w(Rpj,rj,wfj)2mcp,r[w]p,(r,)maxj=1,...,m1rj1rj-1pjj=1mRpj,rj,wfj,

Graph

as desired. The assertion follows.

Proof of Theorem 3.1

The proof will consist of two steps. In the first step we prove the result for very specific q . In the second step we iterate the first step to obtain the desired result.

Step 1. In this step we assume that there is some j0{1,...,m+1} such that

1pj0<1qj0,1pj1qjforjj0.

Graph

Since none of the statements in the formulation of the proposition depend on the order of the indices, we may assume without loss of generality that j0=m+1 . More precisely, we can let πSm+1 be the transposition given by π(j)=j for jj0,m+1 and π(j0)=m+1 , π(m+1)=j0 . Replacing the index j by π(j) everywhere then indeed allows us to reduce to the case j0=m+1 .

We define 1s:=1-1rm+10 , 1p:=1-1pm+1>0 , 1q:=1-1qm+10 , and w:=wm+1-1 so that w=j=1mwj . For an m+1 -tuple (α1,...,αm+1) we will use the notation α=(α1,...,αm) so that the arrow notation will always refer to an m-tuple. Thus, we have now reduced the problem to proving that there exist m weights WAq,(r,s) such that fjLqj(Wjqj) , fm+1Lq(W-q) , where W:=j=1mWj , with

3.3 j=1mfjLqj(Wjqj)fm+1Lq(W-q)2mj=1mfjLpjwjpjfm+1Lp(w-p),

Graph

and

3.4 [W]q,(r,s)Cp,q,r,s[w]p,(r,s)maxj=1,...,m1rj-1qj1rj-1pj.

Graph

Indeed, the result then follows by setting Wm+1:=W-1 and by noting that

[W]q,(r,)=[W]q,(r,s),[w]p,(r,)=[w]p,(r,s).

Graph

The construction of the m weights W1,...,Wm relies on the multilinear Rubio de Francia algorithm as well as a clever usage of the translation lemma to deal with the parameter s. Setting

1sj:=1p-1s1qj-1q-1s1pj1p-1q,

Graph

we have

1sj1p-1s1qj-1q-1s1qj1p-1q=1qj

Graph

with equality if and only if 1qj=1pj and so that 1sj1qj1pj , and

j=1m1sj=1p-1s1q-1q-1s1p1p-1q=1s.

Graph

We set

1pj(s):=1pj-1sj,1qj(s):=1qj-1sj,1rj(s):=1rj-1sj

Graph

and 1p(s):=j=1m1pj(s)=1p-1s , p(s):=(p1(s),...,pm(s)) , and similarly for 1q(s) , q(s) , and r(s) .

We emphasize here that 1pj(s)=0 if and only if 1pj=1qj and we encourage the reader to verify that the remaining steps in this proof remain valid in this particular case.

We may compute

3.5 1pj-1qj=1p(s)-1q(s)1p(s)1pj(s),1qj(s)=1q(s)1p(s)1pj(s).

Graph

We set gj:=|fj|1pj(s)1pjwj-1sj1pj so that

gjLpj(s)(wjpj(s))=fjLpjwjpj1pj(s)1pj

Graph

and, using the notation from Lemma 3.2, we set

Wj:=(Rpj(s),rj(s),w(gj))-1p(s)-1q(s)1p(s)wj1q(s)1p(s).

Graph

Unwinding the definitions, it follows from (3.5) and property (i) of our multilinear Rubio de Francia algorithm that

3.6 fjLqj(Wj)=gj1pj1qj(Rpj(s),rj(s),w(gj))-1pj-1qj1qjLpj(s)(wjpj(s))1qj1pj(s)gjLpj(s)(wjpj(s))1qj1pj(s)=fjLpjwjpj1qj1pj.

Graph

Next, it follows from (3.5), Hölder's inequality, and property (ii) that

fm+1Lq(W-q)fm+1w-1LpW-1wL11p-1q=fm+1Lp(w-p)j=1mRpj(s),rj(s),w(gj)1p(s)-1q(s)1p(s)w1p(s)-1q(s)1p(s)L11p(s)-1q(s)=fm+1Lp(w-p)j=1mRpj(s),rj(s),w(gj)Lp(s)(wp(s))1p(s)-1q(s)1p(s)fm+1Lp(w-p)j=1mRpj(s),rj(s),w(gj)Lpj(s)(wjpj(s))1p(s)-1q(s)1p(s)2mfm+1Lp(w-p)j=1mfjLpjwjpj1pj-1qj1pj.

Graph

By combining this estimate with (3.6), we have proven (3.3).

Finally, we prove (3.4). Noting that

1rj-1qj=1p(s)-1q(s)1p(s)1rj(s)+1q(s)1p(s)1rj-1pj,

Graph

it follows from Hölder's inequality and (iii) that for a cube Q we have

3.7

Graph

Moreover, we have

infyQj=1mRpj(s),rj(s),w(gj)(y)1p(s)-1q(s)1p(s)W11q-1s,Qw1q(s)1p(s)q(s),Q=w11p-1s,Q1q(s)1p(s).

Graph

By combining this with (3.7) we find that

3.8 j=1mWj-111rj-1qj,QW11q-1s,Qcp(s),r(s)[w]p(s),(r(s),)maxj=1,...,m1rj(s)1rj(s)-1pj(s)1p(s)-1q(s)1p(s)[w]p,(r,s)1q(s)1p(s).

Graph

By the translation lemma, Lemma 2.13, we have [w]p(s),(r(s),)=[w]p,(r,s) and, moreover, by using (3.5) we compute

1rj(s)1rj(s)-1pj(s)1p(s)-1q(s)1p(s)+1q(s)1p(s)=1pj(s)-1qj(s)1rj(s)+1rj(s)-1pj(s)1qj(s)1rj-1pj1pj(s)=1rj-1qj1rj-1pj,

Graph

which we interpret as being equal to 1 when 1qj=1pj=1rj , so that

maxj=1,...,m1rj(s)1rj(s)-1pj(s)1p(s)-1q(s)1p(s)+1q(s)1p(s)=maxj=1,...,m1rj-1qj1rj-1pj.

Graph

Hence, (3.4) follows by taking a supremum over all cubes Q in (3.8). This concludes Step 1.

Step 2. Now suppose q is arbitrary. For each j we either have 1pj<1qj or 1pj1qj . Assume without loss of generality that there is a j1{1,...,m} such that

3.9 1pj1qjifj{1,...,j1},1pj<1qjifj{j1+1,...,m+1}.

Graph

Indeed, if this is not the case then, just as in Step 1, we may permute the indices to reduce back to this case.

The strategy will be to construct the m+1 weights W in m-j1+1 steps through repeated application of Step 1.

We define

θk:=j=m-k+2m+11qj-1pjj=j1+1m+11qj-1pjifk{1,...,m-j1+1};0ifk=0,

Graph

so that 0=θ0θ1θm-j1+1=1 . Thus, defining,

1qjk:=1qj+θk1pj-1qj,

Graph

we have

1qj=1qj01qj11qjm-j11qjm-j1+1=1pj.

Graph

Now, we define

q1:=q11,...,qj11,qj1+1,...,qm,pm+1q2:=q12,...,qj12,qj1+1,...,qm-1,pm,pm+1qm-j1:=q1m-j1,...,qj1m-j1,qj1+1,pj1+2,...,pm+1.

Graph

First we will check that the reciprocals of the coordinates of these m+1 -tuples sum to 1. Indeed, using j=1m+11pj=j=1m+11qj=1 , we have

j=1j11qjk=j=1j11qj+θkj=1j11pj-1qj=j=1j11qj+θk1-j=j1+1m+11pj-θk1-j=j1+1m+11qj=j=1j11qj+j=m-k+2m+11qj-1pj=1-j=j1+1m-k+11qj-j=m-k+2m+11pj

Graph

so that

j=1j11qjk+j=j1+1m-k+11qj+j=m-k+2m+11pj=1,

Graph

as desired.

Now, for k{1,...,m-j1+1} we define

γk:=maxj=1,...,j11rj-1qjk-11rj-1qjk,

Graph

which should be interpreted as being equal to 1 when 1qjk=1rj , and we write qk=(q1k,...,qmk) for the m-tuple given by the first m coordinates of qk , with 1qk:=j=1m1qjk .

We may apply Step 1 with j0=j1+1 to obtain weights Wm-j1=(W1m-j1,...,Wm+1m-j1) such that

3.10 j=1m+1fjLqjm-j1((Wjm-j1)qjm-j1)2mj=1m+1fjLpjwjpj

Graph

and

3.11 [Wm-j1]qm-j1,(r,)Cp,q,r[w]p,(r,)γm-j1+1.

Graph

Next we apply Step 1 with j0=j1+2 to obtain weights Wm-j1-1 with

j=1m+1fjLqjm-j1-1((Wjm-j1-1)qjm-j1-1)2mj=1m+1fjLqjm-j1((Wjm-j1)qjm-j1)

Graph

and

[Wm-j1-1]qm-j1-1,(r,)Cp,q,r[Wm-j1]qm-j1,(r,)γm-j1.

Graph

Combining these estimates with (3.10) and (3.11) we obtain

j=1m+1fjLqjm-j1-1((Wjm-j1-1)qjm-j1-1)(2m)2j=1m+1fjLpjwjpj

Graph

and

[Wm-j1-1]qm-j1-1,(r,)Cp,q,r[w]p,(r,)γm-j1γm-j1+1.

Graph

Continuing this process, applying Step 1 with j0=j1+k for k=3,...,m-j1+1 , we conclude, setting W:=W0 , that

3.12 j=1m+1fjLqj(Wjqj)=j=1m+1fjLqj0((Wj0)qj0)(2m)m-j1+1j=1m+1fjLpjwjpj

Graph

and

3.13 [W]q,(r,)=[W0]q0,(r,)Cp,q,r[w]p,(r,)k=1m-j1+1γk.

Graph

Since (2m)m-j1+12m2 , we note that (3.1) now follows from (3.12). Finally, we note that (3.2) follows from (3.13), provided we can show that

3.14 k=1m-j1+1γk=maxj=1,...,m+11rj-1qj1rj-1pj.

Graph

Note that by our initial assumption (3.9), this maximum is attained at some j2{1,...,j1} .

We claim that

γk=1rj2-1qj2k-11rj2-1qj2k

Graph

for all k{1,...,m-j1+1} . Assuming for the moment that the claim is true, we find that

k=1m-j1+1γk=k=1m-j1+11rj2-1qj2k-11rj2-1qj2k=1rj2-1qj201rj2-1qj2m-j1+1=1rj2-1qj21rj2-1pj2,

Graph

proving (3.14).

To prove the claim, we compute

1rj-1qjk=1rj-1qj-θk1rj-1qj+θk1rj-1pj=1rj-1pj(1-θk)1rj-1qj1rj-1pj+θk

Graph

so that

1rj-1qjk-11rj-1qjk=(1-θk-1)1rj-1qj1rj-1pj+θk-1(1-θk)1rj-1qj1rj-1pj+θk=fk1rj-1qj1rj-1pj,

Graph

where

fk(x)=(1-θk-1)x+θk-1(1-θk)x+θk.

Graph

We note that proving the claim is equivalent to proving the equality

maxj=1,...,m+1fk1rj-1qj1rj-1pj=fkmaxj=1,...,m+11rj-1qj1rj-1pj.

Graph

The inequality

fkmaxj=1,...,m+11rj-1qj1rj-1pj=fk1rj2-1qj21rj2-1pj2maxj=1,...,m+1fk1rj-1qj1rj-1pj

Graph

is clear. To prove the converse inequality, it suffices to show that fk is an increasing function for all k{1,...,m-j1+1} . Computing

fk(x)=(1-θk-1)((1-θk)x+θk)-(1-θk)((1-θk-1)x+θk-1)((1-θk)x+θk)2=θk-θk-1((1-θk)x+θk)20,

Graph

we have proven the desired result. This concludes Step 2. The assertion follows.

Proof of Theorem 2.2

The result essentially follows from an application of Theorem 3.1. However, in order to use this result we must reduce to a case where 1p1 so that we may set 1pm+1=1-1p0 . To reduce to this case, we employ a general rescaling trick that also appears in the proof of the case m=1 given by Auscher and Martell in [[1], Theorem 4.9].

First we will show that we may assume that 1r:=j=1m1rj=1 . Indeed, assuming we have shown the result for 1r=1 , we consider the m+1 -tuple (|f1|r,...,|fm|r,|h|r) . Then, since

[w]qr,(rr,sr)1r=(w11r,...,wm1r)q,(r,s),

Graph

we find that for all wAqr,(rr,sr) we have

|h|rLqr(wqr)=hLq((w1r)q)rϕq(w11r,...,wm1rq,(r,s))rj=1mfjLqj((wj1r)qj)r=ϕq[w]qr,(rr,sr)1rrj=1m|fj|rLqjr(wjqjr).

Graph

Thus, since j=1mrrj=1 , applying the extrapolation result with r replaced by rr , q replaced by qr , and s replaced by sr , we find that for any pr with pr>rr and pr<sr , or equivalently, for all p>(r,s) , we have

hLp(wp)=|h|rLpr((wr)pr)1rϕpr,qr,rr,srw1r,...,wmrpr,(rr,sr)1rj=1m|fj|rLpjr((wjr)pjr)1r=ϕpr,qr,rr,sr[w]p,(r,s)r1rj=1mfjLpjwjpj,

Graph

for all wAp,(r,s) , with

ϕpr,qr,rr,sr([w]p,(r,s)r)1r=2m2rϕqCp,q,r,s[w]p,(r,s)rmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s1r

Graph

as desired.

Now that we have reduced to the case where 1r=1 , we have 1s1pj=1m1rj=1 . Thus, we may set 1pm+1:=1-1p0 , 1qm+1:=1-1q0 , 1rm+1:=1-1s0 and wm+1:=w-1 .

Let fm+1Lpm+1(wm+1pm+1) and let W=(W1,...,Wm+1) be the weights obtained from Theorem 3.1. Setting W=(W1,...,Wm) and W:=j=1mWj we find, using the assumption (2.2) and property (3.1) of W , that

3.15 |h,fm+1|hLq(Wq)fm+1Lqm+1(Wm+1qm+1)ϕq([W]q,(r,s))j=1m+1fjLqj(Wjqj)2m2ϕq([W]q,(r,s))j=1m+1fjLpjwjpj.

Graph

Moreover, it follows from (3.2) that

[W]q,(r,s)=[W]q,(r,)Cp,q,r[w]p,(r,)maxj=1,...,m+11rj-1qj1rj-1pj=Cp,q,r,s[w]p,(r,)max1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s.

Graph

By combining this estimate with (3.15) and by noting that

hLp(wp)=supfm+1Lpm+1(wm+1pm+1)=1|h,fm+1|,

Graph

the assertion follows.

Applications of the extrapolation theorem

In applying extrapolation theorems, one can obtain further results by making appropriate choices in the m+1 -tuples. We provide some applications in this section.

Boundedness of operators through extrapolation

Given an operator T defined on m-tuples of functions, one can apply the extrapolation result to the m+1 -tuples (f1,...,fm,T(f1,...,fm)) to obtain the following extension result:

Theorem 4.1

Let T be an m-linear or a positive valued m-sublinear operator and suppose that there exist r1,...,rm(0,) , s(0,] and q1,...,qm(0,] with q(r,s) and an increasing function ϕq such that

4.1 TLq1(w1q1)××Lqm(wmqm)Lq(wq)ϕq([w]q,(r,s))

Graph

for all wAq,(r,s) .

Then for all p1,...,pm(0,] with p>(r,s) and all weights wAp,(r,s) the operator T extends to a bounded operator Lp1(w1p1)××Lpm(wmpm)Lp(wp) . Moreover, T satisfies the bound

TLp1(w1p1)××Lpm(wmpm)Lp(wp)2m2rϕqCp,q,r,s[w]p,(r,s)rmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s1r,

Graph

where 1r=j=1m1rj .

Proof

Let f1,...,fm be simple functions. By (4.1) we have

T(f1,...,fm)Lq(wq)ϕq([w]q,(r,s))j=1mfjLqj(wjqj)

Graph

for all wAq,(r,s) . Thus, by applying Theorem 2.2 to the m+1 -tuple (f1,...,fm,T(f1,...,fm)) we find that for all p1,...,pm(0,] with p>(r,s) and all weights wAp,(r,s) we have

T(f1,...,fm)Lp(wp)ϕp,q,r,s([w]p,(r,s))j=1mfjLpjwjpj

Graph

with ϕp,q,r,s given by (2.3). Since this estimate holds for all simple functions f1,...,fm , the assumptions on T allow us to conclude the results through density.

The initial estimate (4.1) is often obtained through sparse domination. Once we have an estimate of the form

|T(f1,...,fm),g|supSΛ(r1,...,rm,s),S(f1,...,fm,g),

Graph

it follows from duality and Proposition 2.7 that for p1,...,pm(0,] with p>(r,s) and 1p<1 , we have

4.2 TLp1(w1p1)××Lpm(wmpm)Lp(wp)[w]p,(r,s)max1r11r1-1p1,...,1rm1rm-1pm,1s1s-1p.

Graph

We are, however, still missing the cases outside of the reflexive range 1p<1 . One can reach these cases through extrapolation, see [[32]]. The novelty in our result is that we also obtain a quantitative weighted bound in this range through Theorem 4.1.

Corollary 4.2

Let T be an m-linear or a positive valued m-sublinear operator and suppose that there exist r1,...,rm(0,) , s[1,] such that for all bounded compactly supported f1,...,fm,g we have

|T(f1,...,fm),g|supSΛ(r1,...,rm,s),S(f1,...,fm,g),

Graph

where the supremum runs over all sparse collections S with a fixed sparsity constant. Then for all p1,...,pm(0,] with p>(r,s) and all weights wAp,(r,s) the operator T extends to a bounded operator Lp1(w1p1)××Lpm(wmpm)Lp(wp) . Moreover, T satisfies the bound

4.3 TLp1(w1p1)××Lpm(wmpm)Lp(wp)[w]p,(r,s)max1r11r1-1p1,...,1rm1rm-1pm,1-1s1p-1s.

Graph

Proof

We set 1τ:=1s+j=1m1rj . Assuming the set of p satisfying p>(r,s) is non-empty, we have τ<1 . Indeed, for such a p we have

1τ>1p+j=1m1pj=1.

Graph

Setting 1qj:=τrj<1rj , we have

1q=11s+j=1m1rjj=1m1rj=1-τs

Graph

so that

1r11r1-1q1==1rm1rm-1qm=1s1s-1q=11-τ.

Graph

Then by (4.2) with this specific choice of qj we obtain

TLq1(w1q1)××Lqm(wmqm)Lq(wq)[w]q,(r,s)11-τ

Graph

for all wAq,(r,s) . Thus, it follows from (4.1) that for all p1,...,pm(0,] with p>(r,s) and all weights wAp,(r,s) we have

4.4 TLp1(w1p1)××Lpm(wmpm)Lp(wp)[w]p,(r,s)11-τmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s.

Graph

Noting that

1rj-1qj1rj-1pj=(1-τ)1rj1rj-1pj,1q-1s1p-1s=(1-τ)1-1s1p-1s,

Graph

the estimate (4.3) now follows from (4.4).

Remark 4.3

We note that in particular the quantitative bound (4.3) extends the bound (4.2) obtained from the sparse form, even though the proof only used the sparse bound for the particular values 1pj=τrj with 1τ=1s+j=1m1rj . It seems that these values are, in some sense, central for the sparse form and the quantitative bound for these values has already appeared in [[33]], but giving a quantitative bound for the whole range of p>(r,s) is new. In the case m=1 this value becomes p=r(1s+1r)=1+rs which is the value central in the main theorem of [[5]]. In particular when r=1 , s= we have p=2 which is central in the theory of Calderón-Zygmund operators.

In the full-range case, i.e., when r1==rm=1 , s= , the particular case we consider becomes p1==pm=m+1 and in [[10]] a bound in this case for multilinear Calderón-Zygmund operators was found. Using the sparse domination result of [[10]], this result was extended by Li, Moen, and Sun in [[34]] to the range of pj(1,) with 1p1 . They showed that for a multilinear Calderón-Zygmund operator T, all p1,...,pm(1,) with 1p1 and all weights wAp we have

4.5 TLp1(w1)××Lpm(wm)Lp(vw)[w]Apmaxp1p,...,pmp,1,

Graph

where the class Ap is defined through the constant in (2.18), and vw:=j=1mwjppj . They proved that this same bound holds even in the case 1p>1 for multilinear sparse operators, leading them to conjecture that the bound (4.5) should also extend to the case 1p>1 . This conjecture was independently proven to be true by Conde-Alonso and Rey [[6]] and Lerner and Nazarov [[29]] for kernels satisfying log –Dini conditions. We also refer the reader to [[23], [26]], where the weaker Dini condition was considered in the linear case. The Dini condition was used in the multilinear setting by Damián, Hormozi and Li [[9]] where, in addition, quantitative mixed multilinear Ap A bounds were considered.

Our results yields another proof of the extension of the bound to the case 1p>1 . To see this, we note that by replacing the wj by wjpj we have vw=j=1m(wjpj)ppj=wp and

w1p1,...,wmpmAp=[w]p,(1,)p.

Graph

Thus, the result (4.5) takes the equivalent form

TLp1(w1p1)××Lpm(wmpm)Lp(wp)[w]p,(1,)maxp1,...,pm,p,

Graph

which precisely corresponds to the bound (4.2). By applying our extrapolation result we can now extend (4.5), proving the following:

Corollary 4.4

Let T be an m-linear Calderón-Zygmund operator. Then for all p1,...,pm(1,) we have

TLp1(w1)××Lpm(wm)Lp(vw)[w]Apmaxp1p,...,pmp,1.

Graph

As in Corollary 4.2, our result actually yields weighted bounds for multilinear Calderón-Zygmund operators in the more general case p1,...,pm(1,] with 1p>0 .

Vector-valued extrapolation

By Fubini's Theorem we are able to extend the extrapolation theorem to a vector-valued setting. In the following result we are considering spaces of the form Lp(wp;Lq(Ω)) for p,q(0,] , a weight w, and Ω a σ -finite measure space. Such spaces consist of functions f:RnLq(Ω) such that the function fLq(Ω) lies in Lp(wp) , with fLp(wp;Lq(Ω)):=fLq(Ω)Lp(wp) . In the case when p=q , we can use Fubini's Theorem to find that

fLq(wq;Lq(Ω))=fLq(wq)Lq(Ω),

Graph

valid for any q(0,] , allowing us to carry over scalar-valued estimates to the vector-valued setting.

Theorem 4.5

(Vector-valued extrapolation) Let r1,...,rm(0,) , s(0,] . Let Ω be a σ -finite measure space, let q1,...,qm(0,] and q(r,s) , and let (f1,...,fm,h) be an m+1 -tuple of measurable functions on Rn×Ω . Assume that there is an increasing function ϕq,r,s such that the inequality

4.6 hLq(wq)ϕq,r,s([w]q,(r,s))j=1mfjLqj(wjqj)

Graph

holds pointwise a.e. in Ω for all wAq,(r,s) .

Then for all p1...,pm(0,] with p>(r,s) there is an increasing function ϕp,q,r,s such that

hLp(wp;Lq(Ω))ϕp,q,r,s([w]p,(r,s))j=1mfjLpj(wjpj;Lqj(Ω))

Graph

for all wAp,(r,s) . More explicitly, we can take

ϕp,q,r,s(t)=2m2rϕq,r,sCp,q,r,strmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s1r,

Graph

where 1r=j=1m1rj .

Proof

Set f~j:=fjLqj(Ω) , h~:=hLq(Ω) , which, by Fubini's Theorem, are measurable functions on Rn . Then by Fubini's Theorem, the assumption (4.6), and Hölder's inequality, we have

h~Lq(wq)=hLq(wq)Lq(Ω)ϕq,r,s([w]q,(r,s))j=1mfjLqj(wjqj)Lq(Ω)ϕq,r,s([w]q,(r,s))j=1mfjLqj(wjqj)Lqj(Ω)=ϕq,r,s([w]q,(r,s))j=1mf~jLqj(wjqj).

Graph

Thus, we may apply Theorem 2.2 to the m+1 -tuple (f~1,...,f~m,h~) , proving the result.

By iterated uses of Fubini's Theorem, a similar argument also allows us to extrapolate to vector-valued bounds with iterated Lq -spaces which were considered by Benea and Muscalu through their helicoidal method [[4]], but we do not detail this here.

We emphasize here that our extrapolation result goes through even if we have qj= for some j{1,...,m} in (4.6). The conclusion of our result then yields vector-valued estimates in the mixed normed spaces Lpj(L) .

If we take Ω=N with the counting measure, we obtain vector-valued bounds for q -spaces. Given an m-linear operator T and sequences of measurable functions (fk1)kN,...,(fkm)kN , we may define

4.7 T((fk1)kN,...,(fkm)kN):=(T(fk1,...,fkm))kN.

Graph

By combining the vector-valued extrapolation theorem with Corollary 4.2, we obtain the following:

Corollary 4.6

Let T be an m-linear operator and suppose that there exist r1,...,rm(0,) , s[1,] such that for all bounded compactly supported f1,...,fm,g we have

|T(f1,...,fm),g|supSΛ(r1,...,rm,s),S(f1,...,fm,g),

Graph

where the supremum runs over all sparse collections S with a fixed sparsity constant.

Then for all p1...,pm,q1,...,qm(0,] with p,q>(r,s) , the operator T has a bounded extension Lp1(w1p1;q1)××Lpm(wmpm;qm)Lp(wp;q) given by (4.7). Moreover, there is an increasing function ϕp,q,r,s such that

TLp1(w1p1;q1)××Lpm(wmpm;qm)Lp(wp;q)ϕp,q,r,s([w]p,(r,s))

Graph

for all wAp,(r,s) . More explicitly, we can take

4.8 ϕp,q,r,s(t)tmax1r11r1-1q1,...,1rm1rm-1qm,1-1s1q-1s·max1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s.

Graph

Proof

For each j{1,...,m} , let (fkj)kN be a sequence of simple functions with at most finitely many non-zero entries. Setting fj(x,k):=fkj(x) and h(x,k):=T(fk1,...,fkm)(x) , it follows from Corollary 4.2 that (4.6) is satisfied with

ϕq,r,s(t)tmax1r11r1-1q1,...,1rm1rm-1qm,1-1s1q-1s.

Graph

The assertion now follows from Theorem 4.5 and density.

Remark 4.7

If one can use an argument where extrapolation is only required once, then we may be able to replace the exponent in (4.8) by the smaller exponent

max1r11r1-1p1,...,1rm1rm-1pm,1-1s1p-1s

Graph

which no longer depends on the exponents of the qj spaces. One way of doing this is by considering a vector-valued sparse domination rather than a scalar one. Such a sparse domination for the bilinear Hilbert transform is obtained in [[3]]. See also [[18]] where such ideas are used for vector-valued Calderón-Zygmund operators.

The bilinear Hilbert transform

The bilinear Hilbert transform

BHT(f1,f2)(x):=p.v.Rf1(x-t)f2(x+t)dtt

Graph

is an integral operator falling outside of the theory of bilinear Calderón-Zygmund operators. It was introduced by A. Calderón and he wanted to know if it was bounded as an operator from L2×L to L2 . This question was answered by Lacey and Thiele and they showed that BHT is bounded Lp1×Lp2Lp for all p1,p2(1,] with 23<p< , 1p=1p1+1p2 , see [[25]]. It is an open problem whether one can remove the condition 1p<32 or not. However, in this range several weighted bounds and vector-valued extensions have been obtained, some of which we detail here.

Let r1,r2,s(1,) . Then, under certain conditions on r1 , r2 , and s, the sparse domination

|BHT(f1,f2),g|supSΛ(r1,r2,s),S(f1,f2,g)

Graph

was shown in [[8]]. These conditions can be formulated in the following equivalent ways:

Lemma 4.8

Let r1,r2,s(1,) . Then the following conditions are equivalent:

  • We have max1r1,12+max1r2,12+max1s,12<2;
  • There exist θ1,θ2,θ3[0,1) with θ1+θ2+θ3=1 so that
  • 1r1<1+θ12,1r2<1+θ22,1s>1-θ32.

Graph

The sparse domination in terms of characterization (i) was obtained by Culiuc, Di Plinio and Ou in [[8]] and characterization (ii) was used in [[3]] where more general vector-valued sparse domination results were obtained.

Note that if we have r1,r2,s(1,) satisfying one of the equivalent conditions (i) or (ii) and we have p1,p2(1,] with p>(r,s) , then

1p=1p1+1p2<max1r1,12+max1r2,12<2-max1s,1232

Graph

so that we are still in the range of Lacey and Thiele.

From the sparse domination result for BHT , it was deduced in [[8]] that we have the weighted bounds BHT:Lp1(w1p1)×Lp1(w1p1)Lp(wp) for all p1,p2(1,) with p>(r,s) in the Banach range p>1 and for all wAp,(r,s) . These weighted bounds were used in [[7]] to obtain weighted and vector-valued estimates in the range p1 through extrapolation using products of Ap classes. This result was extended in [[33]] where the full multilinear weight classes were used, but only the cases for finite pj were treated. However, their methods can be used to also obtain the cases with pj= [[32]]. By applying Corollaries 4.2 and 4.6 we obtain the following result:

Corollary 4.9

Let r1,r2,s(1,) satisfy one of the equivalent conditions in Lemma 4.8. Then for all p1,p2(1,] with p>(r,s) we have

BHTLp1(w1p1)×Lp2(w2p2)Lp(wp)[w]p,(r,s)max1r11r1-1p1,1r21r2-1p2,1-1s1p-1s.

Graph

for all wAp,(r,s) .

Moreover, for all p1,p2,q1,q2(1,] with p,q>(r,s) there is an increasing function ϕp,q,r,s such that

4.9 BHTLp1(w1p1;q1)×Lp2(w2p2;q2)Lp(wp;q)ϕp,q,r,s([w]p,(r,s))

Graph

for all wAp,(r,s) .

While Corollary 4.6 gives us an expression for the increasing function ϕp,q,r,s in (4.9), this estimate will not be sharp in general, see also Remark 4.7. Rather, a better quantitative estimate can be obtained if one applies our extrapolation result to weighted bounds that can be obtained from the vector-valued sparse domination result obtained in [[3], Theorem 1], but we do not pursue this further here.

Our result should be compared with [[33], Corollary 2.17] and [[4], Theorem 3]. Qualitatively, we completely recover the results on weighted boundedness in [[33], Corollary 2.17] and extend it in the sense that we also include the cases where either p1 or p2 is equal to and where either q1 or q2 is equal to , but this can also be done through their methods [[32]]. If, for example p1= , then we have p2=p and our scalar bound takes the form

BHT(f1,f2)Lp(wp)[w](,p),(r,s)max1r21r2-1p,1-1s1p-1sf1w1Lf2Lp(w2p)

Graph

for all p(r2,s) and all weights w1 , w2 satisfying

[w](,p),(r,s)=supQw1-1r1,Qw2-111r2-1pw1w211p-1s<.

Graph

This is also slightly more general than the weighted bounds in [[3], Corollary 3] in this endpoint case since they only formulate their result in the case p1= when w1=1 (or more generally, pj= when wj=1 ), but their methods do allow for this more general case.

The result [[4], Theorem 3] asserts that if p1,p2,q1,q2(1,] satisfy p,q>(r,s) for r1,r2,s(1,) satisfying one of the equivalent properties of Lemma 4.8, then we have

4.10 BHTLp1(q1)×Lp2(q2)Lp(q)<.

Graph

This result is completely recovered in Corollary 4.9 in the unweighted version of (4.9).

By again extrapolating from the weighted vector-valued bounds we can also consider iterated q spaces in our results. For example, by applying Theorem 4.5 to the weighted vector valued bounds (4.9), one can obtain

BHT:Lp1(2())×Lp2((2))Lp(2(2))

Graph

for all p1,p2(1,] with 23<p< . Such bounds were already obtained in [[2]] through the helicoidal method, but could not be obtained through earlier extrapolation results. More precisely, to obtain this result through extrapolation one needs to be able to extrapolate away from weighted L estimates which is one of our novelties. These type of multiple vector-valued bounds can be applied to prove boundedness results of operators such as the tensor product of BHT and paraproducts and we refer the reader to [[2]] for an overview of such operators.

Endpoint extrapolation results

Finally, we shall discuss some of the endpoint estimates one can extrapolate from.

The following is an extrapolation result involving weak-type estimates. The trick used to obtain this result is well-known and can be found already in [[17]].

Theorem 4.10

(Weak type extrapolation) Let (f1,...,fm,h) be an m+1 -tuple of measurable functions and let r1,...,rm(0,) , s(0,] . Suppose that for some q1,...,qm(0,] with q(r,s) there is an increasing function ϕq such that

4.11 hLq,(wq)ϕq([w]q,(r,s))j=1mfjLqj(wjqj)

Graph

for all wAq,(r,s) .

Then for all p1...,pm(0,] with p>(r,s) there is an increasing function ϕp,q,r,s such that

4.12 hLp,(wp)ϕp,q,r,s([w]p,(r,s))j=1mfjLpjwjpj

Graph

for all wAp,(r,s) . More explicitly, we can take

4.13 ϕp,q,r,s(t)=2m2rϕqCp,q,r,strmax1r1-1q11r1-1p1,...,1rm-1qm1rm-1pm,1q-1s1p-1s1r,

Graph

where 1r=j=1m1rj .

Proof

Let λ>0 and set Eλ:={xRn:|h(x)|>λ} . We define

hλ:=λχEλ

Graph

and note that by (4.11) we have

hλLq(wq)=λ(wq(Eλ))1qhLq,(wq)ϕq([w]q,(r,s))j=1mfjLqj(wjqj)

Graph

Thus, by applying Theorem 2.2 to the m+1 -tuple (f1,...,fm,hλ) we conclude that for all p1...,pm(0,] with p>(r,s) there is an increasing function ϕp,q,r,s such that

hλLp(wp)ϕp,q,r,s([w]p,(r,s))j=1mfjLpjwjpj

Graph

for all wAp,(r,s) , with ϕp,q,r,s given by (4.13). As λ>0 was arbitrary, noting that supλ>0hλLp(wp)=hLp,(wp) proves (4.12). The assertion follows.

As a consequence we can extrapolate from weak lower endpoint estimates in cases where strong bounds are not available. Passing to the full-range case where r1==rm=1 and s= , writing 1 for the vector consisting of m components all equal to 1, we obtain the following corollary:

Corollary 4.11

Let (f1,...,fm,h) be an m+1 -tuple of measurable functions and suppose that there is an increasing function ϕ such that

hL1m,(w1m)ϕ([w]1,(1,))j=1mfjL1(wj)

Graph

for all wA1,(1,) .

Then for all p1...,pm(1,] with 1p>0 there is an increasing function ϕp such that

hLp,(wp)ϕp([w]p,(1,))j=1mfjLpjwjpj

Graph

for all wAp,(1,) . More explicitly, we can take

ϕp(t)=2m3ϕCptpm.

Graph

On the other hand, we can also extrapolate from the upper endpoints. An application of Theorem 2.2 in the s= case with q1==qm= , writing for the vector consisting of m components all equal to , yields the following:

Theorem 4.12

(Upper endpoint extrapolation) Let (f1,...,fm,h) be an m+1 -tuple of measurable functions and let r1,...,rm(0,) . Suppose that there is an increasing function ϕ such that

hwLϕ([w],(r,))j=1mfjwjL

Graph

for all wA,(r,) .

Then for all p1...,pm(0,] with p>r , there is an increasing function ϕp,r such that

hLp(wp)ϕp,r([w]p,(r,))j=1mfjLpjwjpj

Graph

for all wAp,(r,) . More explicitly, we can take

ϕp,r(t)=2mrϕCp,rtrmaxj=1,...,m1rj1rj-1pj1r,

Graph

where 1r=j=1m1rj .

An interesting application is related to the space BMO of functions of bounded mean oscillation. We define the sharp maximal operator M# by

M#f=supQ|f-f1,Q|1,QχQ

Graph

for locally integrable functions f, where the supremum is taken over all cubes QRn . The classical definition of BMO can be given in terms of M# by saying a measurable function f is in BMO if M#fL , with fBMO:=M#fL . The way we have dealt with weighted estimates in L so far suggests the following definition of a weighted version of the BMO space:

Definition 4.13

Given a weight w, we define the space BMO(w) as those locally integrable functions f such that

fBMO(w):=(M#f)wL<.

Graph

Weighted BMO spaces also appeared in the work of Muckenhoupt and Wheeden in [[36]], and they showed that the estimate

4.14 TfBMO(w)fwL,

Graph

with an explicit constant depending on w, is satisfied when T is the Hilbert transform, if and only if w-1A1 . We recall here that the condition w-1A1 is equivalent to our condition wA,(1,) with [w],(1,)=[w-1]A1 . Later it was shown by Harboure, Macías and Segovia in [[19]] that one can extrapolate from the estimate (4.14) for an operator T to obtain that T is bounded on Lp(wp) for all wpAp . As a consequence of Theorem 4.12 we obtain a qualitative multilinear version of this result.

Corollary 4.14

(Extrapolation from BMO estimates) Let T be an m-(sub)linear operator and let r1,...,rm(0,) . Suppose that there is an increasing function ϕ such that

T(f1,...,fm)BMO(w)ϕ([w],(r,))j=1mfjwjL

Graph

for all wA,(r,) and all fj with fjwjL .

Then for all p1...,pm(0,] with p>r , there is an increasing function ϕp,r such that

T(f1,...,fm)Lp(wp)ϕp,r([w]p,(r,))j=1mfjLpjwjpj

Graph

for all wAp,(r,) and all fjLpj(wjpj) , whenever the left-hand side is finite.

Proof

We apply Theorem 4.12 to the m+1 -tuples (f1,...,fm,M#(T(f1,...,fm))) . Then we find that for all p1...,pm(0,] with p>r , there is an increasing function ϕp,r such that

4.15 M#(T(f1,...,fm))Lp(wp)ϕp,r([w]p,(r,))j=1mfjLpjwjpj

Graph

for all wAp,(r,) and all fjLpj(wjpj) . By the classical Fefferman-Stein inequality for the sharp maximal operator, see [[12]], we find that

T(f1,...,fm)Lp(wp)M#(T(f1,...,fm))Lp(wp),

Graph

for p>1 , with implicit constant depending on the A constant of wp , which is bounded by an increasing function of [w]p,(r,) , where 1r=j=1m1rj , see also [[15], Chapter 7]. Since [w]p,(r,)[w]p,(r,s) by Hölder's inequality, the result for p>1 follows from (4.15). By extrapolating again, we also obtain the cases p1 , proving the assertion.

Examples of multilinear operators satisfying weak-type and BMO endpoint estimates are multilinear Calderón-Zygmund operators, see also [[16], Section 7.4.1]. Weighted estimates in these situations can be found in [[30]].

Acknowledgements

The author would like to thank Cristina Benea and Dorothee Frey for their feedback on the draft. In particular the author is grateful to Cristina for her thoroughness in helping to provide context regarding what is known for the bilinear Hilbert transform and multilinear extrapolation. Finally, the author would like to thank the anonymous referee for their helpful suggestions.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References 1 Auscher P, Martell JM. Weighted norm inequalities, off-diagonal estimates and elliptic operators. I. General operator theory and weights. Adv. Math.. 2007; 212; 1: 225-2762319768. 10.1016/j.aim.2006.10.0021213.42030 2 Benea C, Muscalu C. Multiple vector-valued inequalities via the helicoidal method. Anal. PDE. 2016; 9; 8: 1931-19883599522. 10.2140/apde.2016.9.19311361.42005 3 Benea, C., Muscalu, C.: Sparse domination via the helicoidal method. arXiv:1707.05484v2 (2017) 4 Benea, C., Muscalu, C.: The Helicoidal method. arXiv:1801.10071 (2018) 5 Bernicot F, Frey D, Petermichl S. Sharp weighted norm estimates beyond Calderón–Zygmund theory. Anal. PDE. 2016; 9; 5: 1079-11133531367. 10.2140/apde.2016.9.10791344.42009 6 Conde-Alonso JM, Rey G. A pointwise estimate for positive dyadic shifts and some applications. Math. Ann.. 2016; 365; 3–4: 1111-11353521084. 10.1007/s00208-015-1320-y1353.42011 7 Cruz-Uribe D, Martell JM. Limited range multilinear extrapolation with applications to the bilinear Hilbert transform. Math. Ann.. 2018; 371; 1–2: 615-6533788859. 10.1007/s00208-018-1640-906871963 8 Culiuc A, Di Plinio F, Ou Y. Domination of multilinear singular integrals by positive sparse forms. J. Lond. Math. Soc. (2). 2018; 98; 2: 369-3923873113. 10.1112/jlms.121391402.42013 9 Damián W, Hormozi M, Li K. New bounds for bilinear Calderón–Zygmund operators and applications. Rev. Mat. Iberoam.. 2018; 34; 3: 1177-12103850284. 10.4171/RMI/10211406.42017 Damián W, Lerner AK, Pérez C. Sharp weighted bounds for multilinear maximal functions and Calderón–Zygmund operators. J. Fourier Anal. Appl.. 2015; 21; 1: 161-1813302105. 10.1007/s00041-014-9364-z1321.42036 Dragičević O, Grafakos L, Pereyra MC, Petermichl S. Extrapolation and sharp norm estimates for classical operators on weighted Lebesgue spaces. Publ. Mat.. 2005; 49; 1: 73-9121402001081.42007 Fefferman C, Stein EM. Hp\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$H^{p}$$\end{document} spaces of several variables. Acta Math.. 1972; 129; 3–4: 137-193447953. 10.1007/BF023922150257.46078 Frey D, Nieraeth B. Weak and strong type A1-A∞\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$A_1-A_\infty $$\end{document} estimates for sparsely dominated operators. J. Geom. Anal.. 2019; 29; 1: 247-2823897012. 10.1007/s12220-018-9989-207024065 García-Cuerva J, Rubio de Francia JL. Weighted Norm Inequalities and Related Topics. 1985104: Amsterdam; North-Holland Publishing Co. Grafakos L. Classical Fourier Analysis. 20143: New York; Springer Grafakos L. Modern Fourier Analysis. 20143: New York; Springer Grafakos L, Martell JM. Extrapolation of operators of many variables and applications. J. Geom. Anal.. 2004; 14; 1: 19-462030573. 10.1007/BF029218641049.42007 Hänninen TS, Hytönen TP. The A2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$A_2$$\end{document} theorem and the local oscillation decomposition for Banach space valued functions. J. Oper. Theory. 2014; 72; 1: 193-218. 10.7900/jot.2012nov21.19721313.42037 Harboure E, Macías RA, Segovia C. Extrapolation results for classes of weights. Am. J. Math.. 1988; 110; 3: 383-397944321. 10.2307/23746160676.42020 Hytönen TP. The sharp weighted bound for general Calderón–Zygmund operators. Ann. Math.. 2012; 175; 3: 1473-15062912709. 10.4007/annals.2012.175.3.91250.42036 Hytönen TP, Kairema A. Systems of dyadic cubes in a doubling metric space. Colloq. Math.. 2012; 126; 1: 1-332901199. 10.4064/cm126-1-11244.42010 Hytönen TP, Pérez C. Sharp weighted bounds involving A∞\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$A_\infty $$\end{document}. Anal. PDE. 2013; 6; 4: 777-8183092729. 10.2140/apde.2013.6.7771283.42032 Hytönen TP, Roncal L, Tapiola O. Quantitative weighted estimates for rough homogeneous singular integrals. Israel J. Math.. 2017; 218; 1: 133-1643625128. 10.1007/s11856-017-1462-61368.42015 Johnson R, Neugebauer CJ. Change of variable results for Ap\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$A_p$$\end{document}- and reverse Hölder RHr\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$${\rm RH}_r$$\end{document}-classes. Trans. Am. Math. Soc.. 1991; 328; 2: 639-666 Lacey M, Thiele C. On Calderón's conjecture. Ann. Math. (2). 1999; 149; 2: 475-4961689336. 10.2307/1209710934.42012 Lacey MT. An elementary proof of the A2\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$A_2$$\end{document} bound. Israel J. Math.. 2017; 217; 1: 181-1953625108. 10.1007/s11856-017-1442-x1368.42016 Lacey MT, Mena D. The sparse T1 theorem. Houston J. Math.. 2017; 43; 1: 111-12736479351378.42010 Lerner AK. An elementary approach to several results on the Hardy-Littlewood maximal operator. Proc. Am. Math. Soc.. 2008; 136; 8: 2829-28332399047. 10.1090/S0002-9939-08-09318-01273.42019 Lerner, A.K., Nazarov, F.: Intuitive dyadic calculus: the basics. arXiv:1508.05639 (2015) Lerner AK, Ombrosi S, Pérez C, Torres RH, Trujillo-González R. New maximal functions and multiple weights for the multilinear Calderón–Zygmund theory. Adv. Math.. 2009; 220; 4: 1222-12642483720. 10.1016/j.aim.2008.10.0141160.42009 Li K. Two weight inequalities for bilinear forms. Collect. Math.. 2017; 68; 1: 129-1443591468. 10.1007/s13348-016-0182-21365.42012 Li, K., Martell, J.M., Martikainen, H., Ombrosi, S., Vuorinen, E.: End-point estimates, extrapolation for multilinear Muckenhoupt classes, and applications. arXiv:1902.04951 (2019) Li, K., Martell, J.M., Ombrosi, S.: Extrapolation for multilinear Muckenhoupt classes and applications to the bilinear Hilbert transform. arXiv:1802.03338 (2018) Li K, Moen K, Sun W. The sharp weighted bound for multilinear maximal functions and Calderón–Zygmund operators. J. Fourier Anal. Appl.. 2014; 20; 4: 751-7653232584. 10.1007/s00041-014-9326-51318.42022 Lorist, E., Nieraeth, B.: Vector-valued extensions of operators through multilinear limited range extrapolation. J. Fourier Anal. Appl. arXiv:1712.08157 (2017) Muckenhoupt, B., Wheeden, R.L.: Weighted bounded mean oscillation and the Hilbert transform. Stud. Math. 54(3), 221–237 (1975/1976)

By Bas Nieraeth

Reported by Author

Titel:
Quantitative estimates and extrapolation for multilinear weight classes.
Autor/in / Beteiligte Person: Nieraeth, Bas
Link:
Zeitschrift: Mathematische Annalen, Jg. 375 (2019-10-01), Heft 1/2, S. 453-507
Veröffentlichung: 2019
Medientyp: academicJournal
ISSN: 0025-5831 (print)
DOI: 10.1007/s00208-019-01816-5
Schlagwort:
  • 42B20
  • 42B25
Sonstiges:
  • Nachgewiesen in: DACH Information
  • Sprachen: English
  • Document Type: Article
  • Author Affiliations: 1 = Delft Institute of Applied Mathematics, Delft University of Technology, P.O. Box 5031, 2600 GA, Delft, The Netherlands
  • Full Text Word Count: 18131

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -