arXiv:1609.01090v2 [math.CA] 16 Nov 2016 soitdt tad()bcmseuvln oproving to equivalent becomes (1) and it to associated 2 1 (2) (1) n a ( say The fal1 all If where ae o h prtri question. in operator the for mates Academy. ue e ehd(emdtehlcia ehd o provin for technique, transform method) This Hilbert helicoidal analysis. harmonic the in (termed inequalities valued method new a duced p 1 r 1 ”: 1 ∗ oepeiey let precisely, More h rsn oki aua otnaino u ro artic prior our of continuation natural a is work present The US-AAHVLE NQAIISVATEHELICOIDAL THE VIA INEQUALITIES VALUED QUASI-BANACH + r , . . . , h uhri loaMme fte“iinSolw Institute Stoilow” “Simion the of Member a also is author The L p . . . k Λ 1 R f p , . . . , ≤ ~ T eepc ob sfli te otxsa el ste“linea the is well, as contexts other in useful be tor to expect we eoti ie ometmtsfrΠ for estimates mixed obtain we rdcsΠadfrteblna ibr transform Hilbert bilinear the for and vector quasi-Banach Π and products Banach multiple way this in proving Abstract. rpsto ) nte motn oei lydb h shar the by played is role important norm Another eratorial 8). Proposition h lsia gnrlzdrsrce type) restricted (generalized classical the n h hr siae o h oaie prtra omc norm operatorial localized the for estimates sharp the and usnrsthrough quasinorms W ~ k + omo ( on norm n T ~ n L r , n ftenvlisi h us-aahfaeok(htis, (that framework quasi-Banach the in novelties the of One : n  R j Σ p L P 1 m ( : < aif o vr 1 every for satisfy µ , W p m L k 1 ,µ = ∞  p , | p T ) 1 R satotally a is ( p 1 ) := f n 1 and ; R 1 , eetn h eiodlmto rm[M5 otequasi-Ban the to [BM15] from method helicoidal the extend We W r < k ∈ BHT L g , ; R L k Range , k p < T 1 ˆ Σ k T j R ) ( I W | 0 µ , 1 W r ∞ ≤ ≤ eue oHodrsieult n oevr rcs loca precise very some and H¨older’s inequality to reduces , (  ( ean be RSIABNAADCMLMUSCALU* CAMIL AND BENEA CRISTINA 1 L j f 1 W µ , = ) . . . /r τ < p ∞ ≤ · with , ( µ , σ civdb ulzn its dualizing by achieved , 1 ) T , F  fiiemauesaeadthe and space measure -finite ) ewn opoetevco-audieult f“depth of inequality vector-valued the prove to want We )). ) g , ˆ ∞ × ≤  m Q W m and 1 × τ · , . . . lna prtrs that so operator -linear k n j n 1 T ≤ =1 ~ . . . G r < | ≤ n f 1. ~ × ) r ( m W nteBnc ae h ierzto fteoperator the of linearization the case, Banach the In . 1 a eudrto hog h utlna omΛ form multilinear the through understood be can · m w × L j 1 Introduction METHOD j 1 < p < 1 , H p < , L w , . . . , ⊗ m ′ Q k p ∞ ntewoerneo eegeexponents. Lebesgue of range whole the in Π r m ≤ L j n hc sotie ydaiigthe dualizing by obtained is which , R 1 =1 1 j ; R dualization. L ∞ ≤ Σ ; and n R L ) j m | n (whenever , R r Q ( n m BHT weak- W dµ ( j n r W =1 1 µ , 1 j n L µ , sa meit application, immediate an As . ( + ) µ w ntal eeoe o h bilinear the for developed initially p  j T ) nb ohahee through achieved both be an n . . . n usnrsthrough quasinorms  vle nqaiisfrpara- for inequalities -valued ×  ) -tuples  sdfie as defined is : T → fMteaiso h Romanian the of of L r + L vlaino h op- the of evaluation p n hn0 when p iain fteopera- the of rization” aifissc siae,we estimates, such satisfies − e[M5,weew intro- we where [BM15], le ′ p L r aiu utpevector- multiple various g 1 1 1 m p R j /r × ; n R R = L . . . dµ . . . ; < r < k R L r ′ 1 = ( R j × c context, ach W . ( W L 1 r µ , ) which 1), ( k 1 p weak- µ , w L r , . . . , m ) r 1  ) ) → (see  →  L , 1 p /r L C k n 1 p  esti- l R , with T ~ = n 2 CRISTINABENEAANDCAMILMUSCALU

′ 1 ′ n ′ p R where R = r ,..., r . That is, for every 1 ≤ k ≤ m and every f~k ∈ L k R; L k (W,µ) ~ p′ R R′ W and h ∈ L ; L ( ,µ ) , we want to prove  (3) ~ ~ ~ ~ ~ ~ |Λ (f1,..., fm, h)| . kf1k · . . . · kfmk · khk ′ ′ . T~n Lp1 R;LR1 (W,µ) Lpm R;LRm (W,µ) Lp R;LR (W,µ) Using restricted weak-type interpolation,  we were able in [BM15] to treat also the case 1 ~ m

1 1 1 (4) |Λ (f~ ,..., f~ ,~h)| . |F | p1 · . . . · |F | pm |H| p′ , T~n 1 m 1 m

~ ~ ~ ~ ′ ′ R for all fk, h so that kfkkLRk (W,µ) ≤ 1Fk , khkLR (W,µ) ≤ 1H , where Fk,H ⊆ are sets of finite measure, and H′ ⊆ H is a major subset of H to be constructed in the process. The helicoidal method is a recursive procedure in which vector-valued estimates of depth n corresponding to T~n are proved using localized versions of the (n − 1)-depth vector-valued ′ operator T~n−1. Aiming to prove (4) for fixed sets F1,...,Fm,H,H , we need to exercise great care in evaluating ′ F1,...,Fm,H ~ ~ ~ ′ Λ := Λ ~ f1 · 1F1 ,..., fm · 1Fm , h · 1H . T~n−1;I0 Tn−1;I0

There is another localization associated to the spatial dyadic interval I0, hence the notation Λ . This will be made precise later, but such a localization is natural in the time- T~n−1;I0 frequency analysis setting, where operators are decomposed into wave packets that retain both spatial and frequential information. ′ The estimates needed for ΛF1,...,Fm,H correspond to Lebesgue exponents between 1 and T~n−1;I0 ′ ∞, as a result of the assumption 1 < Rk,R ≤ ∞. So one could estimate the multilinear form as in (3), but as a matter of fact, a sharper result can be obtained by using the fact that the functions f~k are supported inside the sets Fk: (5) ′ ′ F1,...,Fm,H F1,...,Fm,H Λ (f~1,..., f~m,~h) . kΛ k· T~n−1;I0 T~n−1;I0 ~ ~ ~ kf1k ˜ · . . . · kfmk ˜ · khk ′ ˜′ . Lr1 R;LR1 (W,µ) Lrm R;LRm (W,µ) Lr R;LR (W,µ)

′ F1,...,Fm,H ~ ˜    Some information on kfkk Rk W is preserved in the operatorial norm kΛ k, L ( ,µ) T~n−1;I0 which is necessary in the induction step. Obtaining the desired vector-valued inequalities amounts to transforming Lrk estimates into Lpk estimates, and this resembles an extrap- olation principle. If rk ≤ pk, H¨older’s inequality and localizations play an important role, ′ F1,...,Fm,H but in the case rk > pk, the sharp evaluation of kΛ k is essential. The method T~n−1;I0 of the proof is described in greater detail in Section 3, after introducing some necessary definitions. However, if some rj ∈/ [1, ∞) (recall that R = (r1, . . . , rn) corresponds to the target space in (1)), we cannot expect to have an inequality comparable to (5) for the . The difficulty consists, for example, in associating a trilinear form to an operator T : Lp(ℓr1 ) × Lq(ℓr2 ) → Ls(ℓr) when r < 1. The linearization of such an operator is achieved in the quasi-Banach case by dualizing the Lp,∞ through the Lebesgue space Lr (see Section 2). QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 3

We improve the helicoidal method from [BM15] by substituting (5) with

′ F1,...,Fm,H ~ ~ 1 ~ 1 1 ′ ~ ~ ˜ ~ ˜ (6) Tn−1;I0 (f1· F1 ,..., fm· Fm )· H r . kTn−1;I kkf1k r1 R R1 W ·. . .·kfmk rm R Rm W . L 0 L  ;L ( ,µ) L ( ;L ( ,µ))

′ Again, optimal estimates are needed for the operatorial norm kT~ F1,...,Fm,H k, which are in n−1;I0 ′ some sense finer than those for the operatorial norm kΛF1,...,Fm,H k of the multilinear form. T~n−1;I0 All things considered, we are able to prove that whenever (r1, . . . , rm, r) ∈ Range(T ), we have also Range(T~ 1) 6= ∅. Here we write T~ n for the vector-valued operator of depth n asso- ~r R~ ciated to the tuple of vectors R~ = (R1,...,Rm,R). Recursively, whenever (r1, . . . , rm, r) ∈ Range(T~ n−1 ), we are able to give a characterization of Range(T~ n ), where, (R˜1,...,R˜m,R˜) (R1,...,Rm,R) for all 1 ≤ k ≤ m, Rk := (rk, R˜k). We will illustrate how the helicoidal method produces quasi-Banach valued inequalities for two bilinear operators: the paraproduct Π and the bilinear BHT . Then the techniques extend to allow for certain multiple Banach or quasi-Banach valued inequalities (for us, that corresponds to multiple Lp spaces with 0

j j j −−−→n−j (r1, r2, r ) ∈ Range(BHT j+1 n j+1 n j+1 n ). ((r1 ,...,r1 ),(r2 ,...,r2 ),(r ,...,r )) Our main motivation was finding the full range of mixed norm estimates for the operator Π⊗Π, i.e. the biparameter paraproduct. Such estimates imply Leibniz rules in mixed-norm Lp spaces and they can prove useful in the study of nonlinear dispersive PDE (particular cases of these inequalities were used in [Ken04]). In [BM15], we proved that

(7) Π ⊗ Π(f, g) s2 s1 . f p2 p1 g q2 q1 Ly Lx Ly Lx Ly Lx whenever 1 + 1 = 1 , with 1

p1 p2 ∞ q1 q2 2 s1 s2 2 (8) Πx : Lx Ly ℓ × Lx Ly ℓ → Lx Ly ℓ .    4 CRISTINABENEAANDCAMILMUSCALU

Here we extend the helicoidal method to the context of Banach or quasi-Banach spaces, which will allow us to prove estimates as above. Hence, combined with our previous results from [BM15], we have

1 1 1 1 Theorem 1. Let 1 < pi,qi ≤ ∞ and 2 < si < ∞, be so that + = for any index pi qi si i = 1, 2. Then the bi-parameter paraproduct Π ⊗ Π satisfies the following mixed norm estimates:

p1 p2 q1 q2 s1 s2 (9) Π ⊗ Π : Lx Ly × Lx Ly → Lx Ly . The Leibniz rule implied by this theorem, as will be shown in Section 7, can be formulated in the following way: Theorem 2. For any α,β > 0 α β α β α β D D (f · g) . D D f · kgk q1 q2 + kfk p3 p4 · D D g 1 2 s1 s2 1 2 p1 p2 Lx Ly Lx Ly 1 2 q3 q4 Lx Ly Lx Ly Lx Ly

+ Dαf · Dβg + Dβf · Dαg , 1 p5 p6 2 q5 q6 2 p7 p 8 1 q7 q8 Lx Ly Lx Ly Lx Ly Lx Ly whenever 1 < p ,q ≤ ∞, 1 < s < ∞ , max 1 , 1 < s < ∞ , and the indices j j 1+α 1 1+α 1+ β 2 satisfy the natural H¨older-type conditions.  1 1 That s2 should be greater than both 1+α and 1+β is sensible, since this is also the natural condition in the case when s1 = s2, as proved in [MPTT04]. On the other hand, 1 for s1 we only have one constraint: 1+α

r 1/r r1 1/r1 r2 1/r2 (10) Π fk, gk s . fk p · gk q, k k k X   X  X  1 1 1 1 for any p,q,s with 1 < p,q ≤∞,

(8) corresponds to the boundedness of the square and of the maximal function in the one-parameter study of paraproducts. Otherwise, multiple vector-valued inequalities for multilinear Calder´on-Zygmund operators can be obtained as in [GM04] or [CUMP04] by extrapolation, from weighted estimates. The bilinear Hilbert transform is an operator given by dt BHT (f, g)(x) := p.v. f(x − t)g(x + t) . ˆR t p q s 2 1 1 1 It is known ([LT99]) to satisfy L × L → L estimates for 3 , then r1,r2,r corresponds to those (p,q,s) ∈ Range(BHT ) for r2 r 2 r1 2 1 3 1 which 0 ≤ < − . q 2 r1 1 1 1 1 1 (iii) If , ′ ≤ , > , then the range of exponents is similar to the one in ii), with r1 r 2 r2 2 D the roles of r1 and r2 interchanged. That is, r1,r2,r consists of tuples (p,q,s) ∈ 1 3 1 Range(BHT ) for which 0 ≤ < − . p 2 r2 1 1 1 1 1 D (iv) If , ≤ , ′ > , then r1,r2,r corresponds to those (p,q,s) ∈ Range(BHT ) for r1 r2 2 r 2 1 1 1 1 1 1 which 0 ≤ , < + , − < < 1. p q 2 r r s′ 6 CRISTINABENEAANDCAMILMUSCALU

1 1 1 1 1 1 D (v) If > , ≤ and − < ′ < 0, then r1,r2,r corresponds to those (p,q,s) ∈ r1 2 r2 2 2 r 1 3 1 1 3 1 Range(BHT ) for which 0 ≤ < − , ′ < − . q 2 r1 s 2 r1 1 1 1 1 1 1 D (vi) If > , ≤ and − < ′ < 0, then r1,r2,r corresponds to those (p,q,s) ∈ r2 2 r1 2 2 r 1 3 1 1 3 1 Range(BHT ) for which 0 ≤ < − , ′ < − . p 2 r2 s 2 r2 1 1 1 1 1 1 D (vii) If > , > and − < ′ < 0, then r1,r2,r corresponds to those (p,q,s) ∈ r1 2 r2 2 2 r 1 3 1 1 3 1 1 1 Range(BHT ) for which 0 ≤ < − , 0 ≤ < − and ′ < 2 − . p 2 r2 q 2 r1 s r While (i)-(iv) are from [BM15], the rest of the estimates are new and correspond to the 2 quasi-Banach case 3

2. Quasi-Banach spaces: a short review r 1 The quasi-Banach spaces concerned in the present paper are L spaces, with 2

The following observation will be used throughout the paper:

Proposition 7. Let Rn = (r1, . . . , rn), with at least one Lebesgue exponent corresponding rj0 to a quasi-. Then k · kLRn is subadditive, where j0 is any index for which rj0 = min rj. 1≤j≤n

Proof. We prove the above statement by induction. The case n = 1 is trivial, as there is r r only one L space, which is quasi-Banach, and k · kr is subadditive. We assume the statement to be true for any tuple of length n − 1, and prove it for a tuple Rn = (r1, . . . , rn). First, we note that Rn = (r1, R˜n−1), where R˜n−1 = (r2, . . . , rn) is a tuple of length n − 1. We don’t know for sure if R˜n−1 satisfies the hypothesis in the proposition, but if it doesn’t, all the Lebesgue exponents are ≥ 1, and k · kR˜n−1 is r1 subadditive. It also that r1 < 1, and in consequence, k · kRn is subadditive. n−1 rj0 Otherwise, R˜ satisfies the induction hypotheses, and k · k ˜n−1 is subadditive, where LR rj0 := min rj. We note that 2≤rj ≤n

1 j0 j0 j0 r 1/r r rj0 Rn ˜n−1 ˜n−1 k · kL = · LR r1 = · LR r1 . L rj0  1 r rj0 r1 If ≥ 1, then k · k Rn is subadditive; otherwise, k · k Rn is subadditive, which means rj0 L L j the induction statement is true for an index j1 corresponding to the minimal r . 

Remark:. We can reformulate the proposition above in a way that includes the Banach j n 1 n j r 0 j0 case: if R := (r , . . . , r ), with 0 < r ≤ ∞, then k · kLRn is subadditive, where r := min(1, min rj). 1≤j≤n Dualization through Lr spaces. Because the is missing, the duals of Lp quasi-Banach spaces are either too simple (the case of non-atomic spaces, where the dual is {0}), or too complicated (atomic spaces, such as ℓp, the dual of which contains ℓ∞, but doesn’t have an exact characterization). However, for weak Lp spaces, the quasinorm can be “dualized” by using generalized restricted weak-type estimates:

hf, 1E′ i (11) f ∼ sup inf 1 , p,∞ E′⊆E 1− E,0< E <∞ E p major subset

′ ′ ′ where we say E is a major subset of E if E ⊆ E and E ≥ E /2. One can have an equivalent statement by making use of Lr norms. This is very similar to Lemma 2.6 of [MS13]: Proposition 8. The following are equivalent:

(i) kfkp,∞ ≤ A (ii) For any set E of finite measure, there exists a major subset E˜ ⊆ E so that

1 1 r − p kf · 1E˜ kr . A · E .

8 CRISTINABENEAANDCAMILMUSCALU

This can be reformulated as

kf · 1E˜ kr (12) kfkp,∞ ∼ sup inf 1 1 . E˜⊆E r − p 0< E <∞ major subset E

Proof. “(i) ⇒ (ii)” Let E be a set of finite measure and consider the set A Ω := x : f(x) >C · . E 1/p  ˜ ˜ Pick E to be E := E \ Ω, which is a major subset of E for C large enough; this is the only place where we are using inequality (i). Then it’s easy to see that A 1 1 1/r r − p kf · 1E˜kr . C · · E . A · E . E 1/p

“(ii) ⇒ (i)” Let λ> 0 and set E := {x : f(x ) > λ}. From (ii), we know there exists a

˜ major subset E of E for which we have

1 − 1 ˜ 1/r r p λ|E| < kf · 1E˜kr . A · E . ˜ 1/p Since E and |E| are comparable, we get that λ E . A. Since λ> 0 was arbitrary, we deduce (i). 

Remark:. Regarding the notation, from now on we will denote E′ a major subset of E whenever we dualize the k·kLp,∞ quasinorm by identity (11), and E˜ when we use a different Lr space. An Interpolation Theorem. Interpolation for linear or multi-linear Banach-valued operators is almost identical to the scalar case, with the difference that | · |C is replaced by the norm of the Banach space X, k · kX . In the case of bilinear operators, we will prove a quasi-Banach-valued interpolation result, which is the natural modification of the Banach-valued case. First, we recall a new definitions:

Definition 9. A tuple (α1, α2, α3) is called admissible if α1 + α2 + α3 = 1, −1 < α1, α2, α3 < 1 and αj ≤ 0 for at most one index. 1 1 1 In a similar way, a triple of Lebesgue exponents (p,q,s) is called admissible if p , q , s′ is admissible according to the above definition. In most cases, we will have 1 < p,q ≤ ∞ 1 1 1 1  and 2

1 1 1 1 Proposition 10. Let 2

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 9 holds for all admissible tuples (s1,s2, s˜) in a neighborhood of (p,q,s), with the constant

Ks1,s2,s˜ depending continuously on s1,s2, s˜. Then T is of strong type (p,q,s), in the sense that

r 1/r r1 1/r1 r2 1/r2 (14) T (fk, gk) s ≤ Kp,q,s fk p gk q k k k X  X  X  for any sequences of functions for which the RHS is finite. The proof will be postponed to the last Section 8, but it is nothing more than an adaptation of the classical argument. A similar result can be formulated for mixed norm Lp spaces analogous to those ap- pearing in Proposition 7, and for that reason proving restricted weak-type estimates as in (13) will be sufficient for establishing Theorem 4 or Theorem 6. Moreover, the measures involved can be arbitrary (we will need this result with Lebesgue measures replaced by

χ˜I0 dx measures). 1 n 1 n 1 n Proposition 11. Let R1 = r1, . . . , r1 ,R2 = r2, . . . , r2 and R = r , . . . , r be three j j 1 j 1 1 1 tuples satisfying for every 1 ≤ j ≤ n : 1 < r1, r2 ≤ ∞, 2 < r < ∞, and j + j = rj .   r1 r2 Assume T is a bilinear operator satisfying the restricted type estimate: ~ ~ for any sets of finite measure E1 and E2, and any functions f,~g so that f(x) LR1 (W,µ) ≤ 1 (x) and ~g(x) ≤ 1 (x) respectively, the estimate E1 LR2 (W,µ) E2

~ 1/s1 1/s2 (15) k T (f,~g)kLR(W,µ) Ls,˜ ∞(ν) ≤ Ks1,s2,s˜ · ν1 (E1) ν2 (E2) holds for all admissible tuples (s 1,s2, s˜) in a neighborhood of (p,q,s), with the constant Ks1,s2,s˜ depending continuously on s1,s2, s˜. Then T is of strong type (p,q,s), in the sense that ~ ~ kT (f,~g)k R W s ≤ Kp,q,s kfk R1 p k~gk R2 q . L ( ,µ) L (ν) L L (ν1) L L (ν2)

3. Definitions and a layout of the proof In this section, we present the main definitions, as well as a sketch of the method of the proof of our theorems. In addition, we introduce some notation conventions and discuss in detail certain technicalities that are recurrent in the paper.

3.1. Definitions.

Notation:. Given an interval I, we denote by χ˜I the function dist (x,I) (16)χ ˜ (x) := 1+ −10. I |I|  The exponent in the above expression can change all through the presentation, and it can even depend on certain values of p,q,s. This will be only implicit in our estimates, as we attempt to keep the notation simple. In a similar way, the sizes that will be introduced shortly, will appear with exponent 1−ǫ. The ǫ represents a small error, but we will not be tracking its exact value. For example, rj upon an application of H¨older’s inequality, we obtain ǫfinal = ǫinitial · r ; however, we denote both expressions by ǫ, since the final error term can be made arbitrarily small. 10 CRISTINABENEAANDCAMILMUSCALU

Definition 12. A collection {φI }I of smooth bump functions associated to a family I of dyadic intervals is called lacunary if

1 2 α − 1 −|α| M supp φˆ ⊆ , , and |∂ φ (x)| . |I| 2 χ˜ (x). I |I| |I| I I   Similarly, a collection {φI }I of smooth bump functions associated to a family I of dyadic intervals is called non-lacunary if

1 α − 1 −|α| M supp φˆ ⊆ 0, , and |∂ φ (x)| . |I| 2 χ˜ (x). I |I| I I   For lacunary collections we use the notation {ψI }I , and for non-lacunary, {ϕI }I .

Definition 13. The discretized paraproduct ΠI associated to a family I of dyadic intervals is the bilinear expression 1 (17) ΠI(f, g)(x)= cI 1 hf, ϕI ihg, ψI iψI (x), I I 2 XI∈ where {c } I is a bounded of complex numbers. I I∈ In proving the Leibniz rule in Section 7, a special role is played by the paraproducts arising form the classical decomposition into “low” and “high” frequencies. We have (18) f · g(x)= f ∗ ϕk · g ∗ ψk ∗ ψk(x)+ f ∗ ψk · g ∗ ϕk ∗ ψk(x)+ f ∗ ψk · g ∗ ψk ∗ ϕk(x) k k k X  X  X  = Qk(Pkf · Qkg)(x)+ Qk(Qkf · Pkg)+ Pk(Qkf · Qkg). Xk Xk Xk k k k k Here ψk(x) = 2 ψ(2 x), ϕk(x) = 2 ϕ(2 x),ϕ ˆ(ξ) ≡ 1 on [−1/2, 1/2], is supported on [−1, 1] and ψˆ(ξ) =ϕ ˆ(ξ/2) − ϕˆ(ξ). The {Qk}k represent Littlewood-Paley projections k onto the frequency |ξ|∼ 2 , while {Pk}k are convolution operators associated with dyadic dilations of a nice bump function of 1. We refer to any of the expressions on the right hand side of (18) as classical paraproducts. Bilinear Fourier multipliers of the form

(f, g) 7→ fˆ(ξ)ˆg(η)m(ξ, η)e2πi(ξ+η)dξdη ˆR2 given by a multiplier m(·, ·) which is smooth away from the origin, can be expressed as superpositions of classical paraproducts, and hence as a superposition of operators analo- gous to those in (17). The boundedness of the discretized paraproducts will imply that of the classical paraproducts and of the bilinear Fourier multipliers above. For this reason, we only study the discretized paraproducts. Definition 14. Let I be a family of dyadic intervals. For any 1 ≤ j ≤ 3, we define j j |hf, φI i| j size I hf, φ i I = sup , if (φ ) is non-lacunary and I I∈ 1/2 I I I∈I |I|  j 2 j 1 |hf, φI i| 1/2 j size I hf, φI iI∈I = sup k · 1I k1,∞, if (φI )I is lacunary. I0∈I |I0| |I| I⊆I0  XI∈I  QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 11

The energy is defined as j j n energy I hf, φI iI∈I := sup 2 sup( |I|) Z D n∈ I∈D  X where D ranges over all collections of disjoint intervals I0 with the property that |hf, φj i| I0 ≥ 2n, if (φj ) is non-lacunary and 1/2 I I |I0| 1 |hf, φj i|2 k I · 1 1/2k ≥ 2n, if (φj ) is lacunary. 1/2 I 1,∞ I I |I0| |I| I⊆I0 XI∈I  Lemma 15 (Lemma 2.13 of [MS13]). If F is an L1 function and 1 ≤ j ≤ 3, then

j j 1 M size I(hF, φI iI∈I) . sup |F |χ˜I dx I∈I |I| ˆR for any M > 0, with implicit constants depending on M. Lemma 16 (Lemma 2.14 of [MS13]). If F is an L1 function and 1 ≤ j ≤ 3, then j j energy I(hF, φI iI∈I) . kF k1. It is customary to study the trilinear form associated to ΠI rather than the operator itself. The trilinear form is the expression 1 Λ (f,g,h) := c hf, φ1ihg, φ2ihh, φ3i, Π I 1/2 I I I I |I| XI∈ and it can be estimated using the above sizes and energies. Proposition 17 (Proposition 2.12 of [MS13]). Given a paraproduct Π associated to a family I of intervals, 3 (j) j 1−θj (j) j θj ΛΠ(f1,f2,f3) . size I (hfj, φI iI∈I) energy I (hfj, φI iI∈I) , j=1 Y   for any 0 ≤ θ1,θ2,θ3 < 1 such that θ1 + θ2 + θ3 = 1, where the implicit constant depends on θ1,θ2,θ3 only. While the above proposition is the main ingredient, we need localized estimates.

Definition 18. If I0 is some fixed dyadic interval, we define

I (I0) := {I ∈ I : I ⊆ I0} . For a collection I of intervals, we denote I+ := {J dyadic interval : ∃I ∈ I such that I ⊆ J} .

In the particular case of the collection I (I0), we have + I (I0) := {J dyadic interval : J ⊆ 3I0 and ∃I ∈ I such that I ⊆ J} . Definition 19 (Modified Size). We define the following size, which is more suitable for localizations: 1 M (19) size If := sup f · χ˜J dx. J∈I+ J ˆR

g

12 CRISTINABENEAANDCAMILMUSCALU

We note that, thanks to Lemma 15, we have size If . size If. In the particular case of a collection localized to a certain interval I0, we use the notation

+ g (20) size I0 f := size I (I0)f.

g g 3.2. A few Technical Results: Localizations. Throughout this section, we consider I0 to be a fixed interval, and we use the notation

ΠI0 := ΠI(I0). That is, 1 ΠI (f, g)(x)= cI hf, ϕI ihg, ψI iψI (x). 0 |I|1/2 I∈I IX⊆I0 Lemma 20 (Refinement of Lemma 16). If f is a function whose support has the property that dist (supp f,I ) 2k−1 ≤ 1+ 0 ≤ 2k, I0 −kM I . for some k ≥ 1, then energy (I0)(f) 2 f 1 .

Proof. The proof is almost identical to that of Lemma 16, with the only difference that now we have, for any interval I ⊆ I0, 1 2M −kM hf, χ˜I i . 2 inf Mf(y). |I| y∈I

 Lemma 21 (Refinement of Proposition 17). For any functions f, g and h, the trilinear form associated to the paraproduct ΠI0 satisfies

1−θ1 1−θ2 1−θ3 θ1 θ2 θ3 ΛI0 (f,g,h) . size I0 f size I0 g size I0 h f · χ˜I0 1 g · χ˜I0 1 h · χ˜I0 1 for any 0 ≤ θ1 ,θ2,θ3 < 1 such that θ1 + θ2 + θ3 = 1, where the implicit constants depend g g g on θj only.

Proof. We first assume that 0 < θj < 1. Then write f as f = k1≥0 fk1 , where fk1 = f · 1 k . Similarly, {x: dist (x,I0)∼(2 1 −1)|I0|} P

g = gk2 , and h = hk3 . kX2≥0 kX3≥0 From Lemma 16 we have

ΛΠ(I0) (f,g,h) . ΛΠ(I0) (f,g,h) k ,k ,k 1X2 3 1−θ1 1−θ1 1−θ3 . size I0 fk1 size I0 gk2 size I0 hk3 k ,k ,k 1X2 3    θ1 θ2 θ3 · energy I0 fk1 energy I0 gk2 energy I0 hk3 .

For the energy of fk1 we use the estimate in Lemma 20, bounding the expression above by 1−θ1 1−θ2 1−θ3 size I0 fk1 size I0 gk2 size I0 hk3 k ,k ,k 1X2 3    θ1 θ2 θ3 −k1M −k2M −k3M · 2 kfk1 k1 2 kgk2 k1 2 khk3 k1 .       QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 13

For the sizes, we use the trivial estimate size Ifk1 . size If. Now we note that H¨older’s inequality implies

1−gθ1 θ1 2−k1Mθ1 kf kθ1 . 2−k1Mθ1/2(1−θ1) 2−k1M/2kf k k1 1    k1 1 kX1≥0 kX1≥0 kX1≥0  θ1    . kf · χ˜ k . kf · χ˜ kθ1 ,  k1 I0 1 I0 1 kX1≥0   since the fk1 all have almost disjoint support. The proof in the case when some θj = 0 is identical, with the difference that we sum in two indices kι, for ι 6= j.  As immediate consequences, we obtain

Corollary 22. For any functions f, g and h, we can bound the trilinear form ΛΠ(I0) (f,g,h) by

ΛΠ(I0) (f,g,h) . size I0 f · size I0 g · size I0 h · |I0|. Proof. We need only to note that in Lemma 21, we have g g g kf · χ˜I0 k1 . |I0| size I0 f.  g Corollary 23. If E˜ is a fixed set of finite measure, then 1 1 ΠI0 (f, g) · E˜ 1 . size I0 f · size I0 g · size I0 E˜ · I0 . In what follows, we will need a different kind of localization; that is, we fix F, G, E˜ sets g g g of finite measure and define ˜ (21) ΠF,G,E (f, g) (x) := Π (f · 1 , g · 1 ) (x) · 1 (x). I0 I0 F G E˜ This is part of our approach to proving multiple-vector-valued inequalities. We recover also the following result, which first appeared in [BM15]:

Proposition 24. Let I0 be a fixed dyadic interval and F, G, E˜ ⊂ R sets of finite measure. Then 1 1 ˜ ′ −ǫ ′ −ǫ 1 F,G,E r r r −ǫ Λ (f,g,h) . size I 1F 1 · size I 1G 2 · size I 1 ˜ · kf · χ˜I kr kg · χ˜I kr kh · χ˜I k ′ Π(I0) 0  0  0 E  0 1 0 2 0 r g g g 1 1 1 ′ whenever + + ′ = 1, and 1 < r1, r2, r ≤∞. Here ǫ is some small positive number r1 r2 r that will be chosen later. Proof. This result was proved in [BM15] in detail for the bilinear Hilbert transform opera- tor. For paraproducts, we note that it follows from Lemma 21, in the case of restricted-type functions: that is, functions that are bounded above by functions of finite sets. If f(x) ≤ 1E1 (x), g(x) ≤ 1E2 (x) and h(x) ≤ 1E3 (x), then the conclusion is immediate; the general case follows through interpolation. ∞ If there are any L spaces involved (for example, if r2 = ∞), we only need to notice that 1 N size I0 g · 1G := sup |g| · 1G · χ˜I dx ≤ kg · χ˜I0 k∞ · size I0 1G. + ˆ I∈I (I0) |I| R g g 14 CRISTINABENEAANDCAMILMUSCALU

˜ We fix g ∈ L∞ and the trilinear form becomes a bilinear form (f, h) 7→ ΛF,G,E(f,g,h). The Π(I0) desired inequality is proved again for restricted-type functions, which is sufficient, in view of interpolation theory. For characteristic functions of sets, we have the equivalence 1 ′ r1 1F · χ˜I ∼ k1F · χ˜ k , 0 r1 I0 1 ′ ∞ whereχ ˜ is also an L -adapted bump function associated to the interval I0, as defined in I0 (16). The functionχ ˜′ is of the formχ ˜′ =χ ˜α , where α depends on r . For our purposes I0 I0 I0 1 however, the difference betweenχ ˜ andχ ˜′ is not important, and we will denote both of I0 I0 them simply byχ ˜I0 .  An adaptation of Proposition 4.6 from [BM15] is the following: 1 1 Lemma 25. If 1 < r1, r2 < ∞ are so that + = 1, then r1 r2 1 1 F,G,E˜ ′ −ǫ ′ −ǫ 1−ǫ (22) Π (f, g) . size 1 r1 · size 1 r2 · size 1 ·kf ·χ˜ k kg·χ˜ k . I0 1 I0 F I0 G I0 E˜ I0 r1 I0 r2 Proof. Let h be a function so that khk = 1, and define the bilinear form g ∞ g g ˜ Λh(f, g) := ΛI0 (f · 1F , g · 1G, h · 1E˜ ).

Then we have from Lemma 21, with θ3 = 0, that for any sets of finite measure E1 and

E2, and any functions f, g so that f ≤ 1E1 , g ≤ 1E2 , ′ ′ ˜ . 1 1/r˜1 1 1/ r˜2 1 1 1 Λh(f, g) size I0 F size I0 G size I0 E˜ k E1 · χ˜I0 kr˜1 k E2 · χ˜I0 kr˜2 , 1 1 for any tuple (˜r 1, r˜2) in a neighborhood of (r1, r2), with the property that + = 1. r˜1 r˜2 Interpolation theory then implies the inequality 1/r′ 1/r′ ˜ 1 2 Λh(f, g) . size I0 1F size I0 1G size I0 1E˜ kf · χ˜I0 kr˜1 kg · χ˜I0 kr˜2 , for any functions f and g. We note that the implicit constants do not depend on h. The estimate (22) follows, since ˜ ΠF,G,E(f, g) = sup Λ˜ (f, g) . I0 1 h khk∞=1

 In the case of quasi-Banach spaces, when 1 + 1 = 1 > 1, a result resembling Corollary r1 r2 r 23 holds. This cannot be obtained directly from the estimate for ΛΠ(I0)(f,g,h · 1E˜ ), but requires an extra decomposition and handling of the sizes. A similar argument will be used repeatedly throughout the paper, but the details of the decomposition will not be reproduced. Lemma 26. If τ < 1, then for any ǫ> 0 small enough, we have 1 τ τ τ 1 1−ǫ (23) ΠI0 (f, g) · E˜ τ . size I0 f · size I0 g · size I0 E˜ · I0 . 1 1 Proof. Let τ0 > 0 be so that =1+ .    τ gτ0 g g As in Lemma 21, we split

1 ˜ (x) := 1 ˜ , where 1 ˜ = 1 ˜ · 1 k3 . E Ek3 Ek3 E {x: dist (x,I0)∼(2 −1)|I0|} kX3≥0 QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 15

τ We use the of k · kτ , and H¨older’s inequality to get τ τ ΠI0 (f, g) · 1 ˜ . ΠI0 (f, g) · 1 ˜ E τ Ek3 τ k ≥0 X3 1 τ 1 τ . ΠI0 (f, g) · ˜ k ˜ kτ0 . Ek3 1 Ek3 k ≥0 X3 From Corollary 23, we have

ΠI0 (f, g) · 1 ˜ . size I0 f · size I0 g · size I0 1 ˜ · I0 . Ek3 1 Ek3 Using this and the observation that g g g −k3M size I0 1 ˜ . 2 , size I0 1 ˜ . size I0 1 ˜, Ek3 Ek3 E we can estimate the desired expression by g g g

τ τ τ τ−ǫ τ −k3Mǫτ τ ΠI0 (f, g) · 1 ˜ . size I0 f · size I0 g · size I0 1 ˜ · I0 2 1 ˜ . E τ E Ek3 τ0 k ≥0 X3    Similarly to Lemma 21, weg use H¨older’sg inequalityg in order to sum in k3 and bring into play the fast decay when E˜ is away from I0: (24) τ τ τ k Mǫτ τ τ τ 3 τ τ −k3Mǫτ 0 − 2 0 0 τ0 τ0 2 k1 ˜ k1 . 2 1 ˜ · χ˜I0 1 . 1 ˜ · χ˜I0 1 . size I0 1 ˜ · I0 . Ek3 Ek3 E E k ≥0 k ≥0 X3 X3  τ g Since 1 = τ + , we have in fact obtained τ0 1 τ τ τ 1 1−ǫ ΠI0 (f, g) · E˜ τ . size I0 f · size I0 g · size I0 E˜ · |I0|.     g g g

3.3. The method of the proof. 1 2 In [BM15], we proved the vector-valued inequalities T : Lp(R; LR (W,µ))×Lq(R; LR (W,µ)) → s R R W j j j L ( ; L ( ,µ)), whenever 1 < r1, r2 ≤∞ and 1 ≤ r < ∞, where T is either the bilinear Hilbert transform BHT or a paraproduct Π. In the present paper, we are concerned with the case when at least one of the rj is < 1. Whenever 1 ≤ r < ∞, the k · kLr(W,µ) norm can be dualized and the problem reduces to estimating the trilinear form ΛT (f,~g,~ ~h). We recall that in the discrete case, this corre- sponds to k ΛT (fk, gk, hk). ′ There are two coupled statements for the localized trilinear forms ΛF,G,H (f,~g,~ ~h) := P T ;I0 ~ ~ ′ ΛT ;I0 (f · 1F ,~g · 1G, h · 1H ), that are at the very core of our method from [BM15]: 1 1 ′ −ǫ −ǫ 1 F,G,H r′ r′ −ǫ P(n) Λ (f,~g,~ ~h) . size 1 1 size 1 2 size 1 ′ r T ;I0 I0 F I0 G I0 H

~ n n ~ · f R · χ˜I ~g  R ·χ˜I h (Rn)′ · χ˜I ′ g L 1 0gr1 L 2 g0 r2 L 0 r and, for functions f,~g,~ ~h satisfying f~( x) Rn ≤ 1 ( x) , ~g(x) R n ≤ 1 (x) and ~h( x) n ′ ≤ L 1 F L 2 G L(R ) 1 ′ (x) respectively, H ∗ F,G,H′ 1−ǫ 1− ǫ 1−ǫ P (n) Λ (f,~g,~ ~h) . size 1 size 1 size 1 ′ · I . T ;I0 I0 F I0 G I0 H 0    g g g 16 CRISTINABENEAANDCAMILMUSCALU

For paraproducts, the exponents in P∗(n) are 1 − ǫ, while for BHT they are of the form 1+θj 2 − ǫ, where 0 ≤ θ1,θ2,θ3 < 1 and θ1 + θ2 + θ3 = 1. If r< 1, an argument employing the localized trilinear form is not available. Instead, we ˜ ˜ will use only estimates for the localized operators ΠF,G,E and BHT F,G,E. Here we focus I0 I0 on the paraproduct case, for clarity. The localized induction statements are P(n): 1 1 1 ˜ ′ − ′ − − F,G,E p ǫ q ǫ s ǫ Π (f,~g~ ) n . size 1 size 1 size 1 kf~k Rn · χ˜ ~g Rn · χ˜ , I0 LR s I0 F I0 G I0 E˜ L 1 I0 p L 2 I0 q    and, for functions f,~g~ satisfyingg f~(x)g Rn ≤ 1 (xg) and ~g(x) Rn ≤ 1 (x ) respectively, L 1 F L 2 G ˜ 1 1 ∗ F,G,E 1−ǫ 1− ǫ s −ǫ s P (n) : Π (f,~g~ ) n . size 1 size 1 size 1 I . I0 LR s I0 F I0 G I0 E˜ 0 The proof of the induction step P(n − 1) ⇒ P(n) is presented in Theorem 29. The g g g statement P(0) represents the content of Proposition 27, and relies on P∗(0). On the other hand, P∗(0) follows from Lemma 26, where the local estimate for the trilinear form is used. In what follows, we will show how to use P(0) in order to obtain the ℓr-valued estimates, r 1/r for r < 1. We want to estimate k |Π(fk, gk)| s,˜ ∞ under the assumption that ~ r kf(x)kℓr1 ≤ 1F (x) and k~g(x)kℓr2 ≤ 1GP(x). In order to deal with the ℓ quasinorm inside, we dualize through Lr; given E a set of finite measure, we can construct a major subset E˜ ⊆ E so that

r 1/r r 1/r 1 − 1 1 s˜ r (25) Π(fk, gk) s,˜ ∞ ∼ Π(fk, gk) · E˜ r · E . k k X  X 

r 1/r 1 r 1 r The advantage is that k |Π(fk, gk)| · E˜ r = k Π(fk, gk) · E˜ r, and even more, k · kr is subadditive. We use dualization through Lr in order to “linearize” the expression r P  P r 1/r k |Π(fk, gk)| when classical Banach space techniques are not available. Afterwards we employ the helicoidal method as in [BM15]. Through a triple stopping P  r time that will be described shortly, and using the subadditivity of k · kr, (25) is reduced to ˜ 1 r ˜ obtaining “sharp estimates” for ΠI0 (f, g˜) · E˜ r for scalar functions f andg ˜. If we use the trilinear form associated to Π , we get I0

˜ 1 r 1 r−ǫ 1 r−ǫ 1 r−ǫ ΠI0 (f, g˜) · E˜ r . size I0 E1 size I0 E2 size I0 E˜ · |I0|, ˜ 1 1    where we assume f ≤ E1 , g˜ g≤ E2 . However,g we can obtaing a better estimate, which is precisely Lemma 26:

˜ 1 r 1 r−ǫ 1 r−ǫ 1 1−ǫ ΠI0 (f, g˜) · E˜ r . size I0 E1 size I0 E2 size I0 E˜ · |I0|.   1 1−ǫ 1 r−ǫ This improvement (since the sizesg are subunitaryg and r< 1, (sizegI0 E˜ ) ≤ size I0 E˜ ) allows us to prove vector-valued inequalities for paraproducts within the whole Range(Π).  Similarly, the statements P(n) and P∗(n) for Rn = (r1, R˜gn−1) follow fromgP(n − 1) by 1 using Lr -dualization.

3.4. The Triple Stopping Time. All through the paper, we will need estimates along the line

F,G,E˜ F,G,E˜ 1 − 1 (26) Π (f,~g~ ) · 1 . Π k1 · χ˜ k k1 · χ˜ k |E | τ s , I(I0) E˜3 τ I(I0) E1 I0 s1 E2 I0 s2 3

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 17

˜ ~ ~ n 1 where E3 ⊆ E3 is a major subset of E3, and the functions f and ~g satisfy kf(x)kR1 ≤ E1 (x) F,G,E˜ and k~g(x)k n ≤ 1 (x). Here Π represents the operatorial norm, as introduced in R2 E2 I(I0) (6). We note that sometimes we might have F = E1, G = E2 and E˜ = E˜3; also, the estimate doesn’t need to be local, and in this case we regard I0 as being the whole real line, and

χ˜I0 (x) ≡ 1 a. e. The exceptional set represents the set where the values of kf~(x)k n and k~g(x)k n are R1 R2 too large: k1 · χ˜ k k1 · χ˜ k ˜ E1 I0 1 E2 I0 1 Ω := x : M (1E1 · χ˜I0 ) (x) >C ∪ x : M (1E2 · χ˜I0 ) (x) >C , E3 E3 n o n o and E˜3 = E3 \ Ω.˜ Now we partition the collection I of intervals into subcollection Id so c d that for any I ∈ Id, we have dist (I, Ω˜ ) ∼ 2 − 1 |I|. This will allow us to gain some information on the sizes of 1 and 1 : E1 E2  1 k1 · χ˜ k 1 k1 · χ˜ k sup 1 · χ˜ dx . 2d E1 I0 1 , sup 1 · χ˜ dx . 2d E2 I0 1 . ˆ E1 I ˆ E2 I I∈Id |I| R |E3| I∈Id |I| R |E3|

1 M −Md Also, note that sup 1 ˜ · χ˜ dx . 2 . ˆ E3 I I∈Id |I| R

We want to apply our local estimates, but they cannot be directly implemented to ΠI(I0) I nor to ΠId(I0). Instead, we will partition again the collection d(I0) as

(27) I(I0) := Id(I0) := Id(K). n ,n ,n In1,n2,n3 d[≥0 [d 1 [2 3 K∈ [ I I I We will construct collections n1 , n2 , n3 of dyadic intervals, and for every Ij contained in I I I I some nj , we will select d(Ij) ⊆ d(I0) a subcollection of our initial d(I0). Then we say In1,n2,n3 I K ∈ if K = I1 ∩ I2 ∩ I3 for Ij ∈ nj , and

Id(K) := Id(I1) ∩ Id(I2) ∩ Id(I3). In1,n2,n3 In effect, we carry out the local estimates on ΠId(K), where K ∈ . For any n ,n ,n ′ n′ ,n′ ,n′ ′ two such intervals K ∈ I 1 2 3 and K ∈ I 1 2 3 , we could have K ∩ K 6= ∅; but the collections Id(K) and Id′ (K) are going to be disjoint. I The families of dyadic intervals n1 will have the following properties: ′ I (1) the intervals I ∈ n1 are all mutually disjoint ′ n1 (2) moreover, they satisfy |I | . 2 k1E1 · χ˜I0 k1 ′ I I ∈ n1 I X I I (3) whenever I1 ∈ n1 and I ⊆ d(I1) is a subset of the selected d(I1), we have 1 −n1 d k E1 · χ˜I0 k1 + size I(I1) 1E1 . 2 . 2 . |E3|

−n1 (4) as a consequence ofg (3), size + 1 . 2 , where I (I ) := I(I ). In1 (I0) E1 n1 0 1 I ∈I 1[n1 I I Since the construction argumentg is similar for the collections n2 and n3 , we will only I I I I describe it for n1 . We start by setting Stock := d, the collection of intervals in having the property that dist (I, Ω˜ c) ∼ 2d − 1 |I|. I I Assuming the collections n1 were selected for all n1 < n¯1, we construct n¯1 in the following way:  18 CRISTINABENEAANDCAMILMUSCALU

−n¯1 (i) if size IStock 1E1 < 2 , then nothing happens; restart the procedure withn ¯1 :=n ¯1 + 1. −n¯1 I (ii) admitting that we are in the situation where size IStock 1E1 = 2 , look for I ∈ Stock sog that 1 g−n¯1 1E1 · χ˜I dx ∼ 2 . |I| ˆR I (iii) then n¯1 will consist of maximal dyadic intervals I1 ⊆ I0 with the property that they contain at least one interval I ∈ IStock as in (ii), and so that

−n¯1 1 −n¯1 2 ≤ 1E1 · χ˜I1 dx ≤ 2 . |I1| ˆR I I (iv) for every I1 ∈ n¯1 , the collection d(I1) is defined as

Id(I1) := {I ∈ IStock : I ⊆ I1}.

(v) before we restart the procedure from step (i) by increasingn ¯1, we update IStock := IStock \ I(I1). I ∈I 1[n¯1 It’s not difficult to check that conditions (1)-(4) are verified. The last step consist in putting everything together, in order to deduce (26). Here we τ use the subadditivity of k · kτ as follows: F,G,E˜ τ F,G,E˜ τ Π (f,~g~ ) · 1 ˜ . Π (f,~g~ ) · 1 ˜ . I(I0) E3 τ Id(K) E3 τ d≥0 n1,n2,n3 K∈In1,n2,n3 X X X The remaining part follows from the local estimates, as it will be detailed later on. In order to simplify the notation, sometimes we forget about the d parameter.

4. Quasi-Banach Valued Inequalities In the present section, we develop the ideas from Sections 3.3 and 3.4. In fact, we prove p r1 q r2 s r 1 that Π : L (ℓ ) × L (ℓ ) → L (ℓ ), when 2

4.1. A Localization Lemma for quasi-Banach spaces. Fix I0 a dyadic interval, and the sets F, G and E˜, as in (21). Then we have the following result:

1 1 1 1 Proposition 27. For any 1 < r1, r2 ≤ ∞ and 2

f ≤ 1E1 and g ≤ 1E2 . Following standard interpolation theory, it will be enough to prove

˜ ′ ′ (29) ΠF,G,E(f, g) . size 1 1/s1−ǫ · size 1 1/s2−ǫ · size 1 1/s−ǫ I0 s,∞ I0 F I0 G I0 E˜ 1 1  s1  s2  g· k1E1 · χ˜I0 k1 · k1gE2 · χ˜I0 k1 , g QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 19 for all admissible tuples (s1,s2,s) in a neighborhood of (r1, r2, r) that satisfy the usual H¨older scaling condition. To this end, we use Lτ -dualization, as described in Proposition 8. The estimate (28) will be independent on the choice of τ, if we pick τ ≤ r. If we dualize through an Lτ space with τ > r, we obtain on the RHS of (28) the term 1 τ −ǫ size I0 1E˜ . Since the sizes are subunitary, such an estimate is less sharp than (28). This is in part why, for r< 1, we don’t obtain an “optimalg bound” by dualizing through L1, as in Proposition 24. Thus, given a set E3 of finite measure, we define an exceptional set Ω˜ by (30) k1 · χ˜ k k1 · χ˜ k ˜ E1 I0 1 E2 I0 1 Ω := x : M (1E1 · χ˜I0 ) (x) >C ∪ x : M (1E2 · χ˜I0 ) (x) >C , E3 E3 n o n o and set E˜ := E \ Ω.˜ Then E˜ is a major subset of E , and we are left with proving 3 3 3 3 1 1 ˜ 1 − 1 Π (f · 1 , g · 1 ) · 1 · 1 . ΠF,G,E · k1 · χ˜ k s1 · k1 · χ˜ k s2 · E τ s , I0 F G E˜ E˜3 τ I0 E1 I0 1 E2 I0 1 3 ˜ where the operatorial norm ΠF,G, E is given by I0 ˜ 1/r′ −ǫ 1/r′ −ǫ 1/r−ǫ ΠF,G,E := size 1 1 · size 1 2 · size 1 . I0 I0 F I0 G I0 E˜ As in [BM15], we will have a triple stopping time which is dictated by the “concentration” g g g of the sets E1, E2 and E˜3. This is explained in detail in Section 3.4. More exactly, we have three collections of intervals: 1 k1 · χ˜ k In1 −n1−1 −n1 d E1 I0 1 I ∈ : 2 ≤ 1E1 · χ˜I dx ≤ 2 . 2 , I ˆR E3 1 k1 · χ˜ k In2 −n2−1 −n2 d E 2 I0 1 I ∈ : 2 ≤ 1E2 · χ˜I dx ≤ 2 . 2 , I ˆR E3 1 In3 −n3− 1 −n3 − Md I ∈ : 2 ≤ 1E′ · χ˜I dx ≤ 2 . 2 . I ˆR 3 τ Since τ < 1, k · kτ doesn’t satisfy the triangle inequality, but k · kτ does, and we proceed to estimate Π (f · 1 , g · 1 ) · 1 · 1 τ . Π (f · 1 , g · 1 ) · 1 · 1 τ . I0 F G E˜3 E˜ τ I F G E˜3 E˜ τ n1,n2,n3 I∈In1,n2,n3 X X In1 ,n2,n3 We recall that for each I ∈ , ΠI should be understood as ΠId(I), as explained in Section 3.4. F,G,E˜∩E˜3 For each interval I, we can estimate the localized paraproduct ΠI using sizes only. Lemma 26 yields (31) Π f · 1 , g · 1 · 1 · 1 I F G E˜3 E˜ τ 1 1 τ −ǫ . size I f · 1F · size I g · 1G · size I 1 ˜ · 1 ˜ |I| τ E E3 1 −ǫ 1 −ǫ 1 s′ s′ −ǫ 1  2 s  . sizeg I 1F · gsize I 1G g· size I 1E˜ 1 1 1 1 1 s s τ − s +ǫ τ · size I1E 1 · size I 1E  2 · size I1 ˜  · I . g 1 g 2 E3    g g 20 CRISTINABENEAANDCAMILMUSCALU

1 −ǫ 1 −ǫ 1 s′ s′ −ǫ 1 2 s Hence, if we denote K := size I0 1F · size I0 1G · size I0 1E˜ , we actually obtain    g g n τ n τ τ τ − 1 − 2 −n 1− τ +ǫ Π (f · 1 , g · 1 ) · 1 · 1 . K 2 s1 2 s2 2 3( s ) · I . I0 F G E˜3 E˜ τ n1,n2,n3 I∈In1,n2,n3 X X Given the selection process for the collections of intervals In1,n2,n3 , we have that γ γ γ n1 1 1 n2 1 2 3 I . 2 E1 · χ˜I0 1 · 2 E2 · χ˜I0 1 · E3 , I∈In1,n2,n3 X   where γ1, γ2, γ3 are numbers between 0 and 1, with γ1 + γ2 + γ3 = 1. This implies

τ τ τ τ −n1 −γ1 −n2 −γ2 −n 1− τ +ǫ−γ Π (f · 1 , g · 1 ) · 1 · 1 . K 2 s1 · 2 s2 2 3( s 3) I0 F G E˜3 E˜ τ     n ,n ,n 1X2 3 1 γ1 1 γ2 γ3 · E1 · χ˜I0 1 · E2 · χ˜I0 1 · E3 τ τ ˜ 1− τ τ −Md 1 s1 1 s2 s . K 2 E1 · χ˜I0 1 E2 · χ˜I0 1 E3 . Above we used that 2−n1 ≤ 2dk1 · χ˜ k /|E |. The only constraint we have is that E1 I0 1 3 τ τ τ + + 1 − + ǫ> 1= γ1 + γ2 + γ3, s1 s2 s which is definitely true for ǫ> 0. We obtained in this way the desired inequality (29).  4.2. Proof of Theorem 3. Now we are ready to prove a quasi-Banach valued inequality for paraproducts, as well as its localized version, corresponding to the scalar Proposition 27. ˜ Theorem 28. For r , r , r as in Theorem 3, the localized paraproduct ΠF,G,E defined in 1 2 I0 (21) satisfies the estimate ˜ r 1/r 1/s′ −ǫ 1/s′ −ǫ 1/s−ǫ ΠF,G,E (f , g ) . size 1 1 · size 1 2 · size 1 I0 k k s I0 F I0 G I0 E˜ k X     g r1 1/r1 g r2 g1/r2 · fk · χ˜I · gk · χ˜I , 0 s1 0 s2 k k X  X  1 1 1 1 for any 1

r1 1/r1 r2 1/r2 (32) fk ≤ 1F a. e. and gk ≤ 1G a. e. k k X  X  For simplicity, we assume that E = 1, and we construct the set E˜ by removing the parts where M (1 ) and M (1 ) are big: F G

(33) Ω := x : M (1F ) (x) >C F ∩ x : M (1G) (x) >C G ,  

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 21 and E˜ := E \ Ω. We assume that all the intervals I ∈ I have the property that dist (I, Ωc) 1+ ∼ 2d. I

This will allow us to have a better control over the maximal functions of 1F and 1G, and in consequence over their sizes. It will be enough to prove r 1/r 1 −100d 1/s˜1 1/s˜2 (34) Π (fk, gk) · E˜ r . 2 F G , k X  for 1 , 1 arbitrarily close to 1 and 1 respectively. s˜1 s˜2 r1 r2 As usual, we perform a triple stopping time through which we select the collections of I I I intervals n1 , n2 and n3 1 I −n1−1 M −n1 d (35) I ∈ n1 : 2 ≤ 1F (x) · χ˜I dx ≤ 2 . 2 F . I ˆR

Similarly, for 1G and 1 ˜ we have E 1 I −n2−1 M −n1 d (36) I ∈ n2 : 2 ≤ 1G(x) · χ˜I dx ≤ 2 . 2 G , I ˆR

1 I −n3−1 M −n3 −Md (37) I ∈ n3 : 2 ≤ 1E˜ (x) · χ˜I dx ≤ 2 . 2 . I ˆR

In1,n2,n3 I I I r We denote := n1 ∩ n2 ∩ n 3 . Using the sub-additivity of k · k , we get that r r 1/r 1 r 1 r Π (fk, gk) · E˜ r = Π (fk, gk) · E˜ r Xk Xk  r I 1 = Π (I0) (fk, gk) · E˜ r. n1,n2,n3 I ∈In1,n2,n3 X 0 X

In fact, since the functions fk and gk are supported on F and G respectively, we have 1 1 1 1 that ΠI(I0) (fk, gk) · E˜ = ΠI(I0) (fk · F , gk · G) · E˜ , and Proposition 27 implies r 1/r 1 r Π (fk, gk) · E˜ r k X  r r −n1 ′ −ǫ −n2 ′ −ǫ  r   r  −n3(1−ǫ) r r . 2 1 2 2 2 fk · χ˜I · gk · χ˜I 0 r1 0 r2 n1,n2,n3 I ∈In1,n2,n3 k X 0 X X r r −n1 ′ −ǫ −n2 ′ −ǫ  r   r  −n3(1−ǫ) r r . 2 1 2 2 2 1F · χ˜I 1G · χ˜I . 0 r1 0 r2 n1,n2,n3 I ∈In1,n2,n3 X 0 X Above, we used H¨older’s inequality, together with propert y (32). Because of the stopping time properties (1) - (4), we can estimate the norms of 1F · χ˜I0 and 1G · χ˜I0 by: r r r −n1 r r −n2 r 1F · χ˜I . 2 I0 1 , and 1G · χ˜I . 2 I0 2 . 0 r1 0 r2 In this way, the estimate we obtain  for the ℓr-valued paraproduct is 

r 1/r r 1 −n1(r−ǫ) −n2(r−ǫ) −n3(1−ǫ) Π (fk, gk) · E˜ r . 2 2 2 I0 . k n1,n2,n3 I ∈In1,n2,n3 X  X 0 X

22 CRISTINABENEAANDCAMILMUSCALU

The sum I0 can be bounded above by I ∈In1,n2,n3 0 X

2n1 F , 2n2 G , and 2n3 . Hence, if γ , γ , γ are so that 0 ≤ γ , γ , γ ≤ 1 and γ + γ + γ = 1, 1 2 3 1 2 3 1 2 3 n1 γ1 n2 γ2 n3 γ3 I0 . 2 F 2 G (2 ) . I ∈In1,n2,n3 0 X   Since the sizes are all sub-unitary, the equations (35), (36) and (37) imply that (38) r r r 1/r r −n1 −γ1 −n2 −γ2 −n (1−ǫ−γ ) γ1 γ2 1 s˜1 s˜2 3 3 Π (fk, gk) · E˜ r . 2  2  2 F G . k n1,n2,n3 X  X 1 1 1 − ǫ 1 The above converges provided + + > , which is true for (˜s1, s˜2, s˜) in a s˜1 s˜2 r r neighborhood of (r1, r2, r). We obtain that

r r r 1/r r s˜ −γ1 s˜ −γ2 γ γ 1 d  1  d  2  −Md(ǫ−γ3) 1 2 Π (fk, gk) · E˜ r . 2 F · 2 G 2 F G , k X      which is exactly (34). 

Now we proceed with the localized version:

Proof of Theorem 28. The proof consists of two steps: i) First prove Theorem 28 in the case s ≥ r. 1 ii) For 2 < s < r, we make use of the result corresponding to s = r, which was proved in the previous step. i)The case s ≥ r: The result in this case follows from a careful examination of the proof of Theorem 3. We noticed in (38) that we lose some information by changing the exponents of the subunitary sizes from r to r , where 1 < s˜ < ∞. It is this point that we will modify in order to obtain s˜j j the sharper estimate in Theorem 28. This time, the sets F, G, E˜ are fixed, and we have two sequences of functions {fk}k and {gk}k satisfying

r1 1/r1 r2 1/r2 (39) fk ≤ 1E1 , gk ≤ 1E2 , k k X  X  where E1 and E2 are sets of finite measure. Following a variant of Proposition 10 for general measures, it will be enough to prove

F,G,E˜ r 1/r F,G,E˜ (40) Π (fk, gk) . Π · 1E · χ˜I 1E · χ˜I , I0 s,˜ ∞ I0 1 0 s˜1 2 0 s˜2 k X  where the operatorial norm is

1 −ǫ 1 −ǫ 1 F,G,E˜ s˜′ s˜′ −ǫ Π ∼ size 1 1 · size 1 2 · size 1 r I0 I0 F I0 G I0 E˜    g g g QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 23 and (˜s1, s˜2, s˜) is an admissible tuple lying in an arbitrarily small neighborhood of (r1, r2, r). We note that the conditions in (39) imply that ˜ ˜ ΠF,G,E(f , g ) = ΠF ∩E1,G∩E2,E(f , g ). I0 k k I0 k k r We dualize the k · ks,˜ ∞ norm through L , just like in the proof of Theorem 3: given E3 a set of finite measure, we define the exceptional set k1 · χ˜ k k1 · χ˜ k ˜ E1 I0 1 E2 I0 1 Ω := x : M (1E1 · χ˜I0 ) (x) >C ∪ x : M (1E2 · χ˜I0 ) (x) >C , ( E3 ) ( E3 ) ˜ ˜ ˜ and set E3 := E3 \ Ω. Then E3 is a major subset of E3 and ˜ r 1/r ˜ r 1/r 1 − 1 ΠF,G,E(f , g ) ∼ ΠF,G,E(f , g ) · 1 · E s˜ r . I0 k k s,˜ ∞ I0 k k E˜3 r 3 k k X  X  r Since we are in the case r< 1, k · kr is subadditive and we have r −ǫ F,G,E˜ r 1/r r r′ Π (f , g ) · 1 . size 1 · 1 1 I0 k k E˜3 r I F E1 k n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X  X X  r −ǫ r′ 1−ǫ r g r · size 1 · 1 2 size 1 · 1 kf · χ˜ k kg · χ˜ k I G E2 I E˜ E˜3 k I r1 k I r2 k   X . g sizeg1 · 1 r−ǫ · size 1 · 1 r−ǫ size 1 · 1 1−ǫ · I . I F E1 I G E2 I E˜ E˜3 n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X    where we used (28) from Propositiong 27, togetherg with H¨older’sg inequality and the trivial In¯1,n¯2,n¯3 estimate k1F ∩E1 · χ˜I k1 ≤ (size I 1F ∩E1 ) · |I|. Here the collections of intervals are defined in the same way as in Section 3.4. r−ǫ From the expression sizegI 1F · 1E1 , a part will the used to rebuild the norms of 1E1 and 1 , while the rest becomes part of the sharp operatorial norm: E2  r r g r−ǫ ′ −ǫ s˜1 s˜1 size I 1F · 1E1 . size I0 1F · size I1E1 .

Similar estimates are used for 1G· 1E2 and 1 ˜ · 1 ˜ :  g gE E3 g r −ǫ r −ǫ r F,G,E˜ r 1/r r s˜′ s˜′ −ǫ Π (f , g ) · 1 . size 1 1 · size 1 2 · size 1 s˜ I0 k k E˜3 r I0 F I0 G I0 E˜ Xk   n1r n2r r  g −g − −n3 1g− +ǫ · 2 s˜1 · 2 s˜2 2 ( s˜ ) I . n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X r r r s˜1 s˜2 1− s˜ The last part adds up exactly to k1E1 · χ˜I0 k1 k1E2 · χ˜I0 k1 E3 , which proves (40). ii) The case s < r:

This case uses an extra step. The interval I0 and the sets F, G, E,˜ E1, E2, E3 and E˜3 are define as before; the only difference is that now we dualize through Lτ , where τ

24 CRISTINABENEAANDCAMILMUSCALU

I I I where the collection n¯1 , n¯2 , n¯3 are defined by the usual triple stopping time. Now we use 1 1 1 a trick similar to the one in (31): since τ < s < r, we can write = + , where τr > 0. τ r τr We have, due to H¨older, that ˜ ˜ ˜ ˜ ΠF,G,E∩E3 (f , g ) r 1/r . ΠF,G,E∩E3 (f , g ) r 1/r · k1 · 1 · χ˜ k . I k k τ I k k r E˜ E˜3 I τr k k X  X  ˜ ˜ 1/rτ On the right hand side, we initially have E ∩E3 , but we will soon see that size I (1E˜ · 1 ˜ ) appears, accounting for the decay when E˜ ∩ E˜3 is supported away from I. This step E3 was explained in more detail in Proposition 24. g Using the sharp estimate from Theorem 28 in the case s = r, we have − − 1 1 − 1 1 F,G,E˜∩E˜3 r 1/r 1 ǫ 1 ǫ + ǫ + Π (f ,g ) . size 1 ∩ size 1 ∩ size 1 r τr I r τr . I k k τ I F E1 I G E2 I E˜∩E˜3 k X     g g g F ∩E1,G∩E 2, E˜∩E˜3 Here we actually use the result from the previous step for the operator ΠI , F,G,E˜∩E˜3 which coincides with ΠI when applied to functions f supported on E1 and functions g supported on E2. From here on, the proof is identical to the previous case s ≥ r, and we have τ −ǫ τ −ǫ τ F,G,E˜∩E˜3 r 1/r τ s˜′ s˜′ −ǫ Π (f , g ) . size 1 1 size 1 2 size 1 s˜ I0 k k τ I0 F I0 G I0 E˜ k X n τ −n τ    − 1 2 g−n 1− τ +ǫ g g · 2 s˜1 · 2 s˜2 2 3( s˜ ) I n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X τ −ǫ τ −ǫ τ τ τ τ s˜′ s˜′ −ǫ s˜ s˜ 1− 1 2 s˜ 1 2 s˜ . size I0 1F size I0 1G size I0 1E˜ k1E1 · χ˜I0 k1 k1E2 · χ˜I0 k1 E3 , for any admissible tuple (˜s1, s˜2, s˜) in a neighborhood of (s1,s2,s). We don’t include all the details becauseg they areg identical to theg previous case s ≥ r.  4.3. The case p = ∞ or q = ∞. We need to treat separately the cases when p = ∞ or q = ∞. Commonly, these are handled with the help of the two adjoint operators, but here we work with quasi-Banach valued bilinear operators, and cannot consider the associated trilinear form. We will only prove the case q = ∞ for Theorem 28, since Theorem 3 can be seen as a limiting case of the former. Also, the case p = ∞ or or q = ∞ for Theorems 4 or 6 can be treated similarly and we will not elaborate on the details. Proof of Theorem 28 for q = ∞. Here we want to prove F,G,E˜ r 1/r 1 −ǫ 1−ǫ 1 −ǫ Π (f , g ) . size 1 s′ size 1 size 1 s I0 k k s I0 F I0 G I0 E˜ k X     g r1 1/r1 g g r2 1/r2 · fk · χ˜I0 s · gk · χ˜I0 ∞. k k X  X  In order to achieve this, we will use a linear version of Propo sition 10. That is, we r2 1/r2 ∞ consider {gk}k to be a fixed sequence of functions so that gk ∈ L , and it will k X  be enough to prove ˜ ˜ (41) ΠF,G,E(f , g ) r 1/r . kΠF,G,Ekk1 · χ˜ k · g r2 1/r2 · χ˜ , I0 k k s,˜ ∞ I0 E1 I0 s˜ k I0 ∞ k k X  X 

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 25

r1 1/r1 where, this time, k fk ≤ 1E1 , and

PF,G, E˜  1 −ǫ 1−ǫ 1 −ǫ kΠ k := size 1 s˜′ size 1 size 1 s˜ . I0 I0 F I0 G I0 E˜

∞  r2 1/r2  The idea in this case is to isolateg the L normg of gk g ; in other words, we will k X  not use the functions {gk} in the stopping time. The exceptional set is defined as k1 · χ˜ k Ω=˜ x : M (1 · χ˜ ) >C E1 I0 1 , E1 I0 |E |  3  r and E˜3 := E3 \Ω.˜ (As usual, we will dualize through L , and for any given E3, we construct E˜3 a major subset...) The stopping time will differ from the general one described in Section 3.4 in the sense that it will be a double stopping time involving only level sets of 1 and 1 . The F ∩E1 E˜∩E˜3 rest will be analogous to the proof of Theorem 28.

˜ ˜ ΠF,G,E(f , g ) r 1/r · 1 r = ΠF,G,E(f , g ) · 1 r I0 k k E˜3 r I0 k k E˜3 r Xk Xk  ˜ . Π F,G,E(f , g ) · 1 r I k k E˜3 r n1,n3 I∈In1,n3 k X X X r −ǫ r −ǫ r′ r′ 1−ǫ r r . size 1 1 size 1 2 size 1 kf · 1 · χ˜ k · kg · 1 · χ˜ k . I F ∩E1 I G I E˜∩E˜3 k F I r1 k G I r2 n1,n3 I∈In1,n3 k X X    X g g g r Above, we used the subadditivity of k·kr, as well as the local estimate proved in Proposition 27. For the term on the most right, we use H¨older to get

r r r1 1/r1 r r2 1/r2 r kfk · 1F · χ˜I k · kgk · 1G · χ˜I k . fk · 1F · χ˜I · gk · 1G · χ˜I r1 r2 r1 r2 k k k X X  X  . k1 · χ˜ kr · g r2 1/r 2 · χ˜ r · k1 · χ˜ kr . F ∩E1 I r1 k I0 ∞ G I r2 k X 

The estimate (41) reduces to proving

r r r 1− +ǫ 1− r size 1 s˜ size 1 s˜ |I| . k1 · χ˜ k s˜ |E | s˜ . I E1 I E˜3 E1 I0 1 3 n1,n3 I∈In1,n3 X X   g g Here we need to remember the properties achieved through the double stopping time: if I ∈ In1,n3 1 −n1 d k E1 · χ˜I0 k1 −n3 −Md size I 1E1 ≤ 2 . 2 , size I 1E˜ ≤ 2 . 2 |E3| 3 and also g g

n1 γ1 1−γ1 |I| . (2 k1E1 · χ˜I0 k1) |E3| , In ,n I∈X1 3 for any 0 < γ1 < 1. 26 CRISTINABENEAANDCAMILMUSCALU

All of the above imply

r 1− r +ǫ size 1 s˜ size 1 s˜ |I| I E1 I E˜3 n1,n3 I∈In1,n3 X X   r r γ −n1( s˜g−γ1) −n3(1−gs˜ +ǫ−1+γ1) 1 1−γ1 . 2 2 · k1E1 · χ˜I0 k1 · |E3| n1,n3 X r −γ1 k1 · χ˜ k s˜ ˜ . 2d E1 I0 1 2−Mdk1 · χ˜ kγ1 · |E |1−γ1 |E | E1 I0 1 3  3  r r s˜ 1− s˜ . k1E1 · χ˜I0 k1 |E3| , which is precisely what we wanted. This ends the proof for the case when one of p or q is ∞, which is also necessary for the proof of multiple vector-valued estimates when L∞(or ℓ∞) spaces are involved (i.e. some j r1 = ∞). 

5. Multiple Vector-Valued Inequalities We prove the general Theorem 4 by induction. In fact, since the sizes are subunitary, it will be enough to prove its localized version:

n Theorem 29. Let I0 be a fixed dyadic interval, and F, G, E˜ sets of finite measure. R = 1 n j n n r , . . . r is an n-tuple containing at least one index r < 1, and R1 ,R2 are n-tuples j 1 1 1 satisfying component-wise 1 < r ≤ ∞, 1 < rj < ∞, + = . Then the localized i 2 j j rj r1 r2 ˜ paraproduct ΠF,G,E defined in (21), satisfies the following estimates: I0 P (n): 1 1 1 ˜ ′ − ′ − − F,G,E p ǫ q ǫ s ǫ Π (f,~g~ ) n . size 1 size 1 size 1 kf~k Rn ·χ˜ ~g Rn ·χ˜ , I0 LR s I0 F I0 G I0 E˜ L 1 I0 p L 2 I0 q 1 1  1  1  for any p,q,s such thatg+ = , 1g< p,q ≤∞ andg

˜ 1 1 ∗ F,G,E 1−ǫ 1−ǫ s −ǫ s P (n) : Π (f,~g~ ) n . size 1 size 1 size 1 I , I0 LR s I0 F I0 G I0 E˜ 0    whenever f,~g~ are so that f~(x) Rn ≤ 1 (x), ~g(x) Rn ≤ 1 (x). L 1 gF gL 2 G g P P Proof of Theorem 29. Since (0) represents the scalar case, and (1) was proved in The- orem 28, it will be enough to show P(n − 1) ⇒ P(n). Although Theorem 28 deals with discrete ℓr spaces, the result easily extends to general Lr spaces. From Proposition 7, we j r 0 j0 j know that k·k Rn is subadditive, where r = min r . Our iterative argument will depend L 1≤j≤n on the ratio rj0 /r1; more exactly, we will treat the cases r1 = rj0 and r1 > rj0 differently. In both situations, we are re-organizing the (quasi-)norms, in order to obtain subadditivity. We denote the (n − 1)-tuple R˜n−1 := r2, . . . rn , obtaining in this way Rn = r1, R˜n−1 and · LRn = k · kLR˜n−1 r1 .   

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 27

I. The case r1 = min rj < 1. 1≤j≤n r1 In this case, k · kLRn is subadditive, and this will play an important role. We will be proving, for (s1,s2, s˜) in a neighborhood of (p,q,s), that

F,G,E˜ (42) Π (f,~g~ ) n I0 LR s,˜ ∞ 1 1 1 1 1 ′ −ǫ ′ −ǫ −ǫ s s s1 s2 s 1 2 . size I0 1 F size I0 1G size I0 1E˜ 1E1 · χ˜I0 1 · 1E2 · χ˜I0 1 , whenever f,~g~ are so that f~(x) Rn ≤ 1 (x) and ~g(x) Rn ≤ 1 (x), and E , E are g gL 1 E1 g L 2 E 2 1 2 sets of finite measure.

Here again we need to treat two separate cases: (a)s ˜ ≥ r1 (b)s ˜ < r1. 1 In the case (a), it is easier to obtain the exponent s − ǫ for size I0 1E˜; the case (b) relies on case (a) and an intermediate step. g Case I.(a) :s ˜ ≥ r1. r1 We dualize the quasinorm k · ks,˜ ∞ through L , as in Proposition 8. Given E3, a set of finite measure, we define the exceptional set (43) 1 · χ˜ 1 · χ˜ ˜ E1 I0 1 E2 I0 1 Ω := x : M (1E1 · χ˜I0 ) (x) >C· ∪ x : M (1E2 · χ˜I0 ) (x) >C· , E3 E3 n o n o ˜ ˜ and set E3 := E3 \ Ω. Then we have 1 ˜ r ˜ 1 F,G,E r1 1− s˜ F,G,E r Π (f,~g~ ) n · E ∼ Π (f,~g~ ) n · 1 . I0 LR s,˜ ∞ 3 I0 LR E˜3 r1 Now we unfold the Lebesgue norms on the RHS of the above expression:

˜ ˜ 1 ˜ ˜ 1 F,G,E∩E3 ~ r F,G,E∩E3 ~ r ΠI f(x),~g(x) n dx . ΠI f(x),~g(x) n dx ˆR 0 R ˆR R n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3   X X   ˜ ˜ 1 F,G,E∩E3 ~ r = ΠI fw1 (x),~gw1 (x) LR˜n−1 dw1dx ˆR ˆW n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 1 X X   ˜ ˜ 1 F,G,E∩E3 ~ r = ΠI fw1 (x),~gw1 (x) LR˜n−1 dxdw1. ˆW ˆR n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 1 X X   The collection of intervals In¯1,n¯2,n¯3 is defined by a triple time, as in Section 3.4. We used r1 here the subadditivity of k·kRn , and afterwards Fubini. This allows us to use the induction ˜ ˜ F,G,E∩E3 ~ ~ step in order to estimate Π fw1 (x),~gw1 (x) R˜n−1 r1 . First we note that fw1 I L Lx P is supported on E1 and ~g w 1 is supported on E2, and hence (n − 1) implies

˜ ˜ 1 ˜ ˜ 1 F,G,E∩E3 ~ r F ∩E1,G∩E2,E∩E3 ~ r ΠI fw1 (x),~gw1 (x) R˜n−1 dx = ΠI fw1 (x),~gw1 (x) R˜n−1 dx ˆR L ˆR L  r1  r1   ′ −ǫ ′ −ǫ r1 r1 1−ǫ . size 1 ( 1) · size 1 ( 2) · size 1 I F ∩E1 I G∩E2 I E˜∩E˜3 ~ r1 r1 · fw1 ˜n−1 · χ˜I 1 ~gw1 ˜n−1 · χ˜I 1 .  g R1 r1 g R2 r2 g

28 CRISTINABENEAANDCAMILMUSCALU

Upon integrating in w1, we use H¨older’s inequality with conjugate Lebesgue exponents r1 r1 1 and 2 : r1 r1 1 F,G,E˜∩E˜3 r Π f~ (x),~g (x) ˜n−1 dxdw ˆ ˆ I w1 w1 LR 1 W1 R 1   1 r −ǫ r −ǫ 1 ′ 1 ′ 1−ǫ . size 1 (r1) size 1 (r2) · size 1 I F ∩E1 I G∩E2 I E˜∩E˜3 r1 r1 ~ 1 1 · fw1 ˜n−1 · χ˜I r1 r ~gw1 ˜n−1· χ˜I r1 r . g R1 g1 1 R2 g 2 2 Lx Lw1 Lx Lw1

Fubini allows us to switch the order of integration in x and w1 , and thus the expression above is equal to 1 1 r −ǫ r −ǫ 1 ′ 1 ′ 1−ǫ size 1 (r1) size 1 (r2) · size 1 I F ∩E1 I G∩E2 I E˜∩E˜3 r1 r1 ~ 1 1 · fw1 ˜n−1 r1 · χ˜I r  ~gw1 ˜n−1 r1 · χ˜I r . g R1 g1 L 1 R2g 2 L 2 Lw1 x Lw1 x

We recall that the functions f ~ and ~g in fact satisfy

f~(x) Rn ≤ 1 (x) and ~g(x) Rn ≤ 1 (x) L 1 F ∩E1 L 2 G∩E2 because the operator involves the projections onto the sets F and G respectively. So we have proved that (44) 1 1− 1− − F,G,E˜∩E˜3 r r ǫ r ǫ 1 ǫ ~ − 1 1 Π fw1 (x),~gw1 (x) R˜n 1 k k . size I 1F ∩E1 size I 1G∩E2 · size I 1 ˜∩ ˜ · I . I L Lx Lw1 E E3  P∗    We note that this estimate is similar to (n)g from Remark 5.g Turning back tog the initial estimate, we have: ˜ ˜ 1 F,G,E∩E3 ~ r ΠI fw1 (x),~gw1 (x) n dx ˆR 0 R   r1−ǫ r1−ǫ 1−ǫ . size 1 size 1 · size 1 · I . I F ∩E1 I G∩E2 I E˜∩E˜3 n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X    g g g With the purpose of recovering the norms of 1E1 and 1E2 , we separate size 1F · 1E1 as 1 1 1 r r r −ǫ ′ −ǫ (s1) s1 size I 1F · 1E1 . size I0 1F · size I 1E1 ,g and similarly for size 1G · 1E and size 1˜ · 1 ˜ .   g 2 E gE3 g Due to the stopping times, we have g ˜ ˜ g 1 F,G,E∩E3 ~ r ΠI fw1 (x),~gw1 (x) n dx ˆR 0 R 1 1 r  r r1 ′ −ǫ ′ −ǫ −ǫ (s1) (s2) s . size I0 1F · size I0 1G · size I0 1E˜ 1 1 1 n1r n2r r  − s − s −n3 1− s˜ +ǫ  g· g2 1 2 2 2 g  I n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X From Section 3.4, we know that whenever we are performing the stopping times we have γ γ γ n1 1 1 n2 1 2 n3 3 I . 2 E1 · χ˜I0 1 · 2 E2 · χ˜I0 1 · 2 E3 , I∈In¯1,n¯2,n¯3 X   

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 29 where 0 ≤ γj ≤ 1, γ1 + γ2 + γ3 = 1. In the end, we obtain

1 1 1 ˜ 1 r −ǫ r −ǫ r F,G,E r (s )′ (s )′ s −ǫ Π (f,~g~ ) n · 1 . size 1 1 · size 1 2 · size 1 I0 LR E˜3 r1 I0 F I0 G I0 E˜ 1 1 r r r1  s1  s1 1− s˜  g· 1E1 · χ˜I0 1 · 1gE1 · χ˜I0 1 E3 g, which, together with (42) and interpolation, concludes the induction statement P(n−1) ⇒

P(n) for the case (a).

Case I.(b) :s ˜ < r1. τ 1 In this case, we will dualize through an L space, for some τ ≤ s˜ < r . Given E3, we define E˜3 in the same way as before, and hence we have ˜ ˜ 1 1 F,G,E F,G,E s˜ − τ Π (f,~g~ ) n ∼ Π (f,~g~ ) n · 1 · E . I0 LR s,˜ ∞ I0 LR E˜3 τ 3 1 j τ Since τ < r = min r , following Proposition 7, k · kLRn is subadditive. Hence we 1≤j≤n τ have

F,G,E˜ τ F,G,E˜ τ Π (f,~g~ ) n · 1 ≤ Π (f,~g~ ) n · 1 . I0 LR E˜3 τ I LR E˜3 τ n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X 1 1 1 We denote by τr the positive exponent for which 1 + = . This, together with the r τr τ previous case s ≥ r1 will allow us to estimate the term on the RHS of the above expression. Through an intermediate step consisting of a decomposition similar to that appearing in Lemma 26, we eventually obtain ˜ ˜ ˜ ˜ F,G,E∩E3 ~ F,G,E∩E3 ~ Π (f,~g) Rn . Π (f,~g) Rn 1 · 1 ˜ · 1 ˜ · χ˜I I L τ I L r E E3 τr 1−ǫ 1−ǫ 1 + 1 −ǫ 1 + 1 . size 1 · 1 size 1 · 1 size 1 · 1 r1 τr I r1 τ r . I F E1 I G E2 I E˜ E ˜3 ˜ ˜ P∗   F ∩E1,G∩E2,E∩E3 Here in factg we used (n) formg Remark 5, appliedg to the operator Π I . We then obtain a familiar estimate, that will allow us to recover (42):

F,G,E˜ τ τ−ǫ τ−ǫ 1−ǫ Π (f,~g~ ) n · 1 ˜ . size I 1F · 1E size I 1G · 1E size I 1 ˜ · 1 ˜ I . I0 LR E3 τ X X 1  2  E E3  n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 g g g

0 II. The case r1 > rj = min rj < 1. 1≤j≤n 0 Let σ := r1/rj > 1, and R˜n−1 = r2, . . . , rn so that Rn = r1, R˜n−1 . We note that in rj0 rj0 this case k · k n and k · k are both subadditive.   LR LR˜n−1  What we aim for is an inequality similar to (42), for any admissible tuple (s1,s2, s˜) in a neighborhood of (p,q,s). For this, we need to rearrange the quasinorms:

˜ ˜ j0 1 F,G,E F,G,E r j Π (f,~g~ ) n = Π (f~ ,~g ) ˜n−1 1 j r 0 I0 LR s,˜ ∞ I0 w1 w1 LR r /r 0 s,˜ ∞ Lw1 1 ˜ j0  F,G,E ~ r rj0 = Π (fw1 ,~gw1 ) R˜n−1 σ s˜ . I0 L Lw ,∞ 1 rj0

30 CRISTINABENEAANDCAMILMUSCALU

s˜ ,∞ j rj0 Case II.(a) : s ≥ r 0 . Now we will want to “dualize” the Lx norm; we regard it as an σ σ Lw1 -valued function, and we will use the fact that σ > 1 and hence Lw1 is a Banach space. ′ ˜ ˜ For a set E3 of finite measure, we set E3 = E3 \ Ω, where Ω is defined in (43), and we note that

j ˜ rj0 ˜ rj0 r 0 −1 F,G,E ~ F,G,E ~ s˜ Π (f ,~g ) ˜n−1 s˜ ∼ Π (f ,~g ) ˜n−1 h(w , x)dw dx · E , I0 w1 w1 LR Lσ ,∞ I0 w1 w1 LR 1 1 3 w1 j0 r ˆR ˆW1

where h(w ,x) is so that h(·,x) ′ . 1 ′ (x). That is, we are using a Banach-valued 1 Lσ E3 w1 version of Proposition 8, which can be found in [BM15]. j Now we can finally make use of the subadditivity of k · kr 0 . With the collections of LR˜n−1 intervals In¯1,n¯2,n¯3 as in Section 3.4, we have

˜ ′ j0 F,G,E∩E3 ~ r Π (fw ,~gw ) R˜n−1 h(w1,x)dw1dx ˆ ˆ I0 1 1 L R W1 F,G,E˜∩E′ j0 3 ~ r . ΠI (fw1 ,~gw1 ) LR˜n−1 h(w1,x)dw1dx ˆR ˆW n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 1 X X ˜ ′ j0 F,G,E∩E3 ~ r = ΠI (fw1 ,~gw1 ) LR˜n−1 h(w1,x)dxdw1 ˆW ˆR n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 1 X X ˜ ′ j0 F,G, E∩E3 r ~ ˜n−1 1 ′ ′ . ΠI (fw1 ,~gw1 ) LR Lσ Lσ · h(w1, ·) · E˜∩E′ · χ˜I Lσ Lσ . x w1 3 x w1 n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X

Above, we used Fubini and H¨older’s inequality several times. The decaying factorχ ˜I appearing in the last line is motivated by the same line of ideas as those appearing in Lemma 26. Following the induction step P(n − 1), we have

˜ ′ j0 ˜ ′ F,G,E∩E3 r F ∩E1,G∩E2,E∩E3 rj0 ~ ˜n−1 ~ ˜n−1 ΠI (fw1 ,~gw1 ) R σ = ΠI (fw1 ,~gw1 ) R r1 L Lx L Lx j0 j0 r −ǫ r −ǫ j 1 ′ 1 ′ r 0 (r ) (r ) 1 −ǫ . size 1 · 1 1 · size 1 · 1 2 · size 1 · 1 ′ r I F E1 I G E2 I E˜ E3 ~ rj0 rj0 · f ˜n−1 · χ˜ 1 · ~g ˜n−1 · χ˜ 1.  w1 R I r1 w1 R I r2 g L 1 Lx g L 2 Lx g

We integrate the last line in w 1 and we use H¨older’s inequality for the Lebesgue exponents j0 j0 1 1 1 1 r r r1 r2 r ~ j , j , j , since = + . One of the terms we obtain in this way is fw1 R˜n−1 · r 0 r 0 r 0 1 1 1 σ r1 r2 L  rj0  χ˜ r1 r1 , which can be rewritten, using Fubini, as I 1 1 Lx j L r 0 w1

0 rj j0 1 j0 r r1 r1 r ~ 1 ~ 1 ~ n 1 fw1 ˜n−1 · χ˜I r r1 = fw1 ˜n−1 · χ˜I dxdw1 = f R · χ˜I r . R1 1 1 R1 L 1 1 L Lx j ˆW ˆR L Lx L r 0 1 w1 

The function f~ is supported on F ∩ E1 in the x coordinate, and the above expression is bounded by

j0 r rj0 rj0 r1 1 1 1 r1 r1 1F ∩E1 · χ˜I 1 . size I 1F · 1E1 · I .  g QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 31

Similarly, the term corresponding to ~g will be bounded above by

j0 r rj0 rj0 r1 1 1 2 r2 r2 1G∩E2 · χ˜I 1 . size I 1G · 1E2 · I .  For the function h(w1,x ), we have the estimateg

1 1 σ′ σ′ h(w , ·) · 1 ′ · χ˜ ′ ′ . 1 ′ · 1 · χ˜ ′ . size 1 · 1 ′ · I . 1 E˜∩E I Lσ Lσ E3 E˜ I Lσ I E˜ E3 3 x w1 x  Returning to the stopping time, our initial estimate become s g

F,G,E˜ ~ rj0 Π (fw ,~gw ) R˜n−1 h(w1,x)dw1dx ˆ ˆ I0 1 1 L R W1 rj0 −ǫ rj0 −ǫ 1−ǫ . size 1 · 1 · size 1 · 1 · size 1 · 1 ′ · I I F E1 I G E2 I E˜ E3 n¯ ,n¯ ,n¯ In¯1,n¯2,n¯3 1X2 3 I∈ X j j    r 0 −ǫ g r 0 −ǫ g rj0 g s′ s′ −ǫ 1 2 s˜ . size I0 1F · size I0 1G · size I0 1E˜ j j j n1r 0 n2r 0 r 0  − − −n3 1− +ǫ n  γ1 n γ2 γ3 s1 s2 s˜ 1 1 2 1 g· 2 2g 2  g 2 E1 · χ˜I 1 2 E2 · χ˜I 1 E3 , n¯1,n¯2,n¯3 X  

j0 where, as usual, γ1 + γ2 + γ3 = 1. The series will converge if s ≥ r , and we obtain (42), with both terms in the inequality raised to the rj0 power.

Case II.(b) : s < rj0 . Here we will make use of the previous result from II.(b) corre- sponding to s ≥ rj0 , after rewriting the localized paraproduct. First, we want to use the rj0 subadditivity of k · kLRn , and hence we write

1 ˜ ˜ j0 F,G,E F,G,E r rj0 Π (f,~g~ ) n = Π (f,~g~ ) n . I0 R I0 R s˜ L s,˜ ∞ L 0 ,∞ rj

s˜ τ 0 ,∞ The L rj quasinorm will be dualized as in Proposition 8, through L rj0 , for some τ τ rj0 τ ≤ s. In particular, < 1, hence k · k τ is subadditive. rj0 rj0 We have E3 a set of finite measure, and E˜3 a major subset constructed as before, for which

j0 j0 ˜ rj0 ˜ rj0 r − r F,G,E ~ F,G,E ~ s˜ τ Π (f,~g) Rn s˜ ∼ Π (f,~g) Rn · 1 ˜ τ E3 . I0 L 0 ,∞ I0 L E3 0 rj rj

τ j r 0 rj0 Employing the subadditivity of k · kLRn and k · k τ , we have, for the collections of rj0 intervals In¯1,n¯2,n¯3 as in Section 3.4 (45) τ ˜ j0 0 F,G,E ~ r rj Π (f,~g) Rn · 1 ˜ τ I0 L E3 0 rj τ ˜ j0 0 ˜ ˜ F,G, E ~ r rj F,G,E∩E3 ~ τ . Π (f,~g) Rn · 1 ˜ τ = Π (f,~g) Rn τ . I L E3 0 I L L rj n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X X X

32 CRISTINABENEAANDCAMILMUSCALU 1 1 1 j0 Since τ ≤ s < r , there exists τr > 0 so that j + = . An analysis similar to the r 0 τr τ one in Lemma 26, together with the case II.(a) of the present theorem, imply that ˜ ˜ ˜ ˜ F,G,E∩E3 ~ F,G,E∩E3 ~ Π (f,~g) Rn τ . Π (f,~g) Rn j0 · 1 ˜ ˜ · χ˜I I L L I L Lr E∩E3 τr 1−ǫ 1−ǫ 1 −ǫ 1 1 1 . size 1 · 1 size 1 · 1 size 1 · 1 rj0 · I rj0 · size 1 · 1 τr · I τr I F E1 I G E2 I E˜ E˜ 3 I E˜ E˜3 1 1 1−ǫ 1−ǫ τ −ǫ τ = size I 1F · 1E  size I 1G · 1E  size I 1 ˜ · 1 ˜  · I .  g 1 g 2 g E E3 Returning to inequality (45), we have  g g τ g ˜ j0 0 F,G,E ~ r rj Π (f,~g) Rn · 1 ˜ τ I0 L E3 0 rj . size 1 · 1 τ−ǫ size 1 · 1 τ−ǫ size 1 · 1 1−ǫ · I I F E1 I G E2 I E˜ E˜3 n¯1,n¯2,n¯3 I∈In¯1,n¯2,n¯3 X X    τ −ǫ τ −ǫ τ −ǫ p′ g q′ g s g . size I0 1F size I0 1G size I0 1E˜ τ τ −n1 −n2 −n 1− τ +ǫ n γ1 n γ2 γ3 · 2 s1 2 s2 2 3( s˜ ) 2 1 1 · χ˜ 2 2 1 · χ˜ E . g g g E1 I0 1 E2 I0 1 3 n¯1,n¯2,n¯3 X   0 The series above converge because we are under the assumptio n that τ ≤ s < rj , and eventually we obtain inequality (42). This ends the proof of Theorem 29. 

6. Similar results for BHT The bilinear Hilbert transform, BHT in short, is a bilinear operator whose Fourier multiplier is singular along a line. Its study reduces to that of the model operator

1 1 2 3 BHTP(f, g)(x) := hf, φ ihg, φ iφ (x). 1/2 P P P P IP PX∈ Instead of families φI indexed after a collection of intervals (the paraproduct case), we have as index set P, a collection of tritiles. A tile is a product I × ω of an interval I in space and an interval ω in frequency. A tritile P is a set of three tiles sharing the spacial interval:

P = (P1, P2, P3) , Pj = IP × ωPj , |IP | · |ωPj |∼ 1. j j The functions φP are called “wave packets” associated to the tritiles: φP is supported 2 inside the frequency interval .9ωPj , and is L adapted to IP , in the sense that α j −1/2−|α| M b |∂ φP (x)|≤ Cα,M |IP | χ˜IP (x). The collection P of tritiles associated to the model operator BHTP is of one, reflecting the of the singularity. That is, the tiles can be located anywhere in frequency, but there is only one degree of freedom. For more properties of BHT , reduction to the model operator, as well as a self-contained proof, we refer the interested reader to [MS13]. We recall a few results and definitions, that can be found in [BM15]. Definition 30. Traditionally, the size is defined to be a supremum over suitable trees of discretized square functions. Instead, we will use this term for expressions that bound the ‘classical sizes’: 1 M P size f := sup |f(x)|χ˜IP dx. P ∈P |IP | ˆR QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 33

If I0 is a fixed interval, then P(I0) denotes the collection of tritiles in P whose spatial interval is contained inside I0:

P(I0) := {I ∈ P : IP ⊆ I0}. In this case, we define a new size:

1 M 1 M size P(I0)f := max sup |f(x)|χ˜I dx, |f(x)|χ˜I dx . P |I | ˆR P |I | ˆR 0 P ∈ P 0  In our approachg from [BM15], a very important role is played by localized results. Here ΛBHT ;P is the trilinear form associated to the model operator BHTP, and in general, the P collection will be understood from the context (sometimes we write BHTI0 for BHTP(I0)).

Proposition 31 (Lemma 5 from [BM15]). If I0 is a fixed dyadic interval, and P is a rank 1 collection of intervals, then

θ1 θ2 θ3 1−θ1 1−θ2 1−θ3 P P P P ΛBHT ; (I0)(f,g,h) . size (I0)f size (I0)g size (I0)h f·χ˜I0 2 g·χ˜I0 2 h·χ˜I0 2 , for any 0 ≤ θ1,θ2 ,θ3< 1 withθ1+ θ2 + θ3= 1. 

As an immediate consequence, we have: Corollary 32. If F, G and E˜ are sets of finite measure, and f,g,h are so that f ≤ 1 (x), g ≤ 1 (x) and h ≤ 1 (x), then F G E˜ 1+θ1 1+θ2 1+θ3 1 2 1 2 1 2 Λ P(I0)(f,g,h) . size P(I0) F size P(I0) G size P(I0) E˜ · I0 . Corollary 33. If F, G and E˜ are sets of finite measure, and f, g are so that f ≤ g g g 1 (x), g ≤ 1 (x), then F G 1+θ1 1+θ2 1+θ3 1 . P 1 2 P 1 2 P 1 2 BHT I0 (f, g) · E˜ 1 size (I0) F size (I0) G size (I0) E˜ · I0 . Our approach to proving vector-valued estimates for the bilinear operators involves lo- calizations. Just like in the paraproductg case, weg define g ˜ BHT F,G,E(f, g)(x) := BHT (f · 1 , g · 1 )(x) · 1 (x). I0 I0 F G E˜ For the trilinear form associated to this localized operator, we have proved in [BM15] the following inequality:

Proposition 34 (Proposition 8 of [BM15]). If 1 < r1, r2 ≤∞, and 1 ≤ r< ∞, then

1+θ1 1 1+θ2 1 1+θ3 1 − −ǫ − −ǫ − ′ −ǫ 2 r1 2 r2 2 r Λ F,G,E˜ (f,g,h) . size I 1F size I 1G size I 1 ˜ BHT 0 0 0 E I0    · f · χ˜I g · χ˜I h · χ˜I ′ , g 0 r1 0 r2 g 0 r g provided the exponents appearing above are all strictly positive.

s 2 The BHT operator is bounded on L , for 3

34 CRISTINABENEAANDCAMILMUSCALU

This actually follows from the following: Proposition 36. For any 2

1+θ1 1 1+θ1 1 1+θ3 1 F,G,E˜ − −ǫ − −ǫ − ′ −ǫ (47) BHT (f, g) . size 1 2 r1 size 1 2 r2 size 1 2 r I0 r I0 F I0 G I0 E˜

· f · χ˜I0  g · χ˜I0 ,   g r1 r2 g g provided the exponents above are all positive. That is, there exist 0 ≤ θ ,θ ,θ < 1 so that 1 2 3 θ1 + θ2 + θ3 = 1 and

′ 1 1+ θ1 1 1+ θ2 1 1+ θ3 C(r1, r2, r ) < , < , ′ < . r1 2 r2 2 r 2 Proof. It will be enough to prove (48) 1+θ1 1 1+θ2 1 1+θ3 1 F,G,E˜ − −ǫ − −ǫ − ′ −ǫ BHT (f, g) . size 1 2 s1 size 1 2 s2 size 1 2 s˜ P(I0) s,˜ ∞ I0 F I0 G I0 E˜

(49) · 1E · χ˜I 1E · χ˜I   g1 0 s1 2 g0 s2 g whenever f(x) ≤ 1 (x), g (x) ≤ 1 (x ), and for (s ,s , s˜) admissible tuple in a neigh- E1 E 2 1 2 borhood of (r , r , r). We will dualize the weak-Ls˜ norm through an Lτ space, with τ < s˜. 1 2 Given E3 a set of finite measure, we set E˜3 := E3 \ Ω.˜ The exceptional set Ω˜ is defined by the same formula (30). We write P := Pd, where all the tiles in Pd have the property d≥0 that [ dist (I , Ω˜ c) 1+ P ∼ 2d. IP k1 ·χ˜ k −n1 d E1 I0 For every n1 with 2 ≤ 2 , we perform a stopping time similar to the one E3 described in Section 3.4. The stopping time will yield a collection In1 of mutually disjoint n1 intervals, and for every I ∈ I , also a collection P(I) ⊆ Pd of tri-tiles. For each interval I ∈ In1 , we have

−n1−1 1 −n1 2 ≤ 1E1 · χ˜Idx ≤ 2 ∼ size P(I)1E1 . I ˆR

n1 g As a consequence, I∈In1 I . 2 k1E1 · χ˜I0 k1. ′ −n n n Moreover, whenever P ⊆ P (I), size P′ 1 . 2 1 . The collections of intervals I 2 , I 3 P E1 associated to 1 and 1 will have similar properties. E2 E˜3 We choose a τ < s<˜ 1, and it will be enough to estimate ˜ ˜ ˜ BHT F,G,E(f, g) · 1 = BHT F ∩E1,G∩E2,E∩E3 (f, g) . P(I0) E˜3 τ P(I0) τ τ The subadditivity and monotonicity of k · kτ implies that ˜ ˜ ˜ ˜ BHT F ∩E1,G∩E2,E∩E3 (f, g) τ . BHT F ∩E1,G∩E2,E∩E3 (f, g) τ . P(I0) τ P(I) τ n1,n2,n3 I=I1∩I2∩I3 X XInj Ij ∈ An estimate similar to Lemma 26 is needed; informally, this reduces to

F ∩E1,G∩E2,E˜∩E˜3 F ∩E1,G∩E2,E˜∩E˜3 (50) BHT (f, g) . BHT (f, g)k1 · 1 ˜ ˜ · χ˜I , P(I) τ P(I) E∩E3 τ0

QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 35 where τ is so that 1 +1= 1 . Even though we cannot expect to prove such an estimate, 0 τ0 τ we will show that 1+θ1 1+θ2 1+θ3 − F ∩E1,G∩E2,E˜∩E˜3 2 2 2 ǫ BHT (f,g) . size P 1 ∩ size P 1 ∩ size P 1 P(I) τ (I) F E1 (I) G E2 (I0) E˜∩E˜3

· 1 ˜∩ ˜ · χ˜I  I .   gE E3 τ0 g g

Compared to the estimate in Corollary 33, we loose an ǫ in the exponent of size P 1 . (I) E˜∩E˜3 The proof is very similar to the estimate in Lemma 26: first we write 1 as E˜∩E˜3 g

1 ˜ ˜ = 1 ˜ , where 1 ˜ (x)= 1 ˜ ˜ · 1 k3 . E∩E3 Ek3 Ek3 E∩E3 {x: dist (x,I)∼(2 −1)|I|} kX3≥0 τ Using again the subadditivity of k · kτ , together with the estimate in Corollary 33, we have that ˜ ˜ F ∩E ,G∩E ,E˜ F ∩E1,G∩E2,E∩E3 τ 1 2 k3 τ ˜ τ/τ0 BHTP(I) (f, g) τ . BHTP(I) (f, g) 1 · Ek3 k ≥0 X3 τ· 1+θ1 τ· 1+θ2 τ· 1+ θ3 τ τ/τ 2 2 2 ˜ 0 . size P(I)1F ∩E1 size P(I)1G∩E2 size P(I)1 ˜ · I · Ek3 . Ek3    Noticingg that g g 1−ǫ −k3Mǫ size P(I)1 ˜ . size P(I)1 ˜ ˜ · 2 Ek3 E∩E3 and that  g g τ ˜ τ/τ τ −k3M ˜ 0 0 2 · Ek3 . k1 ˜ · χ˜I k1 , Ek3 H¨older’s inequality eventually implies

F ∩E1,G∩E2,E˜∩E˜3 τ BHTP(I) (f, g) τ τ· 1+θ1 τ· 1+θ2 τ· 1+θ3 + τ −ǫ . size P 1 2 size P 1 2 size P 1 2 τ0 · I . (I) F ∩E1 (I) G∩E2 (I) E˜∩E˜3    Now weg are ready to prove inequalityg (47); indeed,g we have ˜ ˜ ˜ ˜ BHT F ∩E1,G∩E2,E∩E3 (f, g) τ . BHT F ∩E1,G∩E2,E∩E3 (f, g) τ P(I0) τ P(I) τ n1,n2,n3 I=I1∩I2∩I3 X XInj Ij ∈ τ· 1+θ1 τ· 1+θ2 τ· 1+θ3 + τ −ǫ . size P 1 2 size P 1 2 size P 1 2 τ0 · I (I) F ∩E1 (I) G∩E2 (I) E˜∩E˜3 n1,n2,n3 I=I1∩I2∩I3 X XInj    Ij ∈ g g g 1+θ1 1 1+θ2 1 1+θ3 1 τ· − τ· − τ· − ′ −ǫ . size P 1 2 s1 size P 1 2 s2 size P 1 2 s˜ (I0) F   (I0) G∩E2   (I0) E˜∩E˜3   n τ n τ − 1 − 2 −n 1− τ +ǫ  s1 s2 3( s˜ )  · g 2 2 g2 · I . g n1,n2,n3 I=I1∩I2∩I3 X XInj Ij ∈ The last line can eventually be bounded above by

τ τ τ −Md˜ s1 s2 1− s˜ 2 k1E1 · χ˜I0 k1 k1E2 · χ˜I0 k1 E3 , proving, upon summation in d ≥ 0, the estimate in (48). 

36 CRISTINABENEAANDCAMILMUSCALU

For the general case of Theorem 6, we would need to prove inductively the following statements: P(n) : 1+θ1 1 1+θ2 1 1+θ3 1 ˜ − −ǫ − −ǫ − ′ −ǫ F,G,E ~ 2 p 2 q 2 s BHTP f,~g n . size P 1F size P 1G size P 1 ˜ (I0)   LR s (I0)  (I0)  (I0) E  g g g · f~ Rn · χ˜ ~g Rn · χ˜ . L 1 I0 p L 2 I0 q

Also, whenever f~(x) Rn ≤ 1 (x) and ~g(x) Rn ≤ 1 (x), we have L 1 F L 2 G P∗(n) : 1+θ1 1+θ2 1+θ3 1 ˜ −ǫ −ǫ − ′ −ǫ 1/s F,G,E ~ 2 2 2 s BHTP f,~g n . size P 1F size P 1G size P 1 ˜ · I0 . (I0)   LR s (I0)  (I0)  (I0) E g g g We note that P(0) is precisely Lemma 35, and as a consequence we also obtain P∗(0). P(1) follows through Lr dualization, and the P∗(0) statement is needed. More generally, 1 P(n) follows through Lr dualization, as a consequence of P∗(n − 1). The proof separates in two cases: s ≥ rj0 and s < rj0 , just like in the paraproduct case. In fact, the proof 1+θj follows the same principle, with the difference that now the exponents of the sizes are 2 and not 1. The details are left to the interested reader.

7. Mixed norm estimates for Π ⊗ Π and the Leibniz Rule

We present the proof of Theorem 1, in the case when s2 < 1 (and as a consequence, 1

Proof of Theorem 1. Since the other cases are very similar, we can assume that Πy, the paraproduct acting on the variable y is of the form

Πy (·, ·)= Qk (Pk (·) ,Qk (·)) . Xk Then we can write Π ⊗ Π as 2 y y Π ⊗ Π(f, g)(x,y)= QkΠ Pk ,Qk (x). k X  2 1/2 Using the inequality QkΦ p ≤ |QkΦ| p, which is true for any 0

p1 p2 ∞ q1 q2 2 s1 s2 2 Πx : Lx Ly ℓ × Lx Ly ℓ → Lx Ly ℓ , which is a consequence of Theorem  29.   Together with the result in [BM15], we obtain the boundedness of Π ⊗ Π in the whole possible range of Lebesgue exponents.  Remark:. In a similar way, mixed-norm Lp estimates for BHT ⊗Π can be deduced, using this time the multiple vector-valued estimates for BHT from Theorem 6. QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 37

Now we provide a proof for Theorem 2, which will also clarify the necessity of the conditions imposed on the Lebesgue coefficients s1 and s2. α β Proof of Theorem 2. As usual, the derivatives D1 and D2 will not act directly on the product f · g, but on the paraproducts. In the bi-parameter case, the product can be written as a sum of nine paraproducts: f · g(x,y)= (f ∗ ϕ ⊗ ϕ · g ∗ ψ ⊗ ψ ) ∗ ψ ⊗ ψ (x,y)+ . . . +(f ∗ ψ ⊗ ψ · g ∗ ψ ⊗ ψ ) ∗ ϕ ⊗ ϕ (x,y). X k l k l k l k l k l k l k,l

| 9 terms{z } We now claim that the derivative of a paraproduct becomes a paraproduct of certain α β ˜ ˜ α β derivatives of f and of g : D1 D2 Π ⊗ Π(f, g) = Π ⊗ Π(D1 f, D2 g) or a like term. The derivatives initially are placed on the outer-most terms of the paraproduct, giving α β α β α β rise to expressions of the form D1 ψk ⊗D2 ψl, D1 ψk ⊗D2 ϕl or D1 ϕk ⊗D2 ϕl. On the dyadic k α kα frequency shell |ξ|∼ 2 , the D1 derivative acts as multiplication by 2 : |ξ|α Dαψ (x) = 2kαψ˜ (x), where ψ˜ = ψ (ξ). 1 k k k 2kα k For a paraproduct of the type k Qk (Pkf · Qkg), web have b α kα ˜ D1 (f ∗ ϕk · g ∗ ψkP) ∗ ψk(x) = 2 (f ∗ ϕk · g ∗ ψk) ∗ ψk(x), k k X  X and now the idea is to transform the multiplication by 2kα again into a derivative. Note that kα α ˜ α ˜ α ˜ 2 g ∗ ψk(x)= g ∗ D1 ψk(x)= D1 (g ∗ ψk)(x) = (D1 g) ∗ ψk(x), ˜ ˜ 2kα where ψ˜ is defined by ψ˜ (ξ) := ψ (ξ). In addition, it becomes evident that we couldn’t k k |ξ|α k have placed the derivativeb on f ∗ ϕk because 0 is contained in its Fourier support. Consequently, in this case, b α α ˜ ˜ ˜ α D1 (Π(f, g))(x)= f ∗ ϕk · D1 g ∗ ψk ∗ ψk(x) := Π(f, D1 g)(x). k X  Similarly, inside the ball |ξ|≤ 2k we have |ξ|α Dαϕ (x) = 2kαϕ˜ (x), where ϕ˜ = ϕ (ξ). 1 k k k 2kα k ˜ b The difference now is that ϕ˜k is not smooth at the origin (unlikeb ψk, the support of ϕk k k contains the origin), and ϕ˜k has only finite decay: every ϕ˜k(x) = 2 ϕ˜(2 x), where b 1 b c (51) ϕ˜(x) . . 1+ x 1+α α A paraproduct associated to a function of fixed decay as in (51) will be denoted Π :

α (52) Π (f, g)(x) := (f ∗ ψk · g ∗ ψk) ∗ ϕ˜k(x). Xk ˜ To deal with the finite decay in (51), we split each ϕ˜k into onto the set |ξ|≤ 2k: 2πinξ 2πinξ b ˜ k k ϕ˜k(ξ) := cn,ke 2 = cn,ke 2 ϕ˜k(ξ) := cn,kϕ˜k,n(ξ), n n n X X X b b b 38 CRISTINABENEAANDCAMILMUSCALU where ϕ˜k is similar to ϕk, but it is going to be constantly equal to 1 on supp ψk +supp ψk. k k Moreover, for any function Φ ∈ S, we use the notation Φk,n(x) := 2 Φ 2 x + n = b n b b b Φ (x + ). As a consequence of (51), the Fourier coefficients satisfy uniformly in k  k 2k 1 |cn,k| . . (1+ |n|)1+α

Now we can see how the derivative in the first variable acts on the paraproduct k Pk(Qkf· Qkg): P α kα D1 (f ∗ ψk · g ∗ ψk) ∗ ϕk(x) = cn,k2 (f ∗ ψk,n · g ∗ ψk,n) ∗ ϕ˜k(x) k n k X  X X α ˜ = cn,k D1 f ∗ ψk,n · g ∗ ψk,n ∗ ϕ˜k(x). n k X X   We denote Pk,nf(x) := f ∗ ϕk,n(x),Qk,nf(x) := f ∗ ψk,n(x). In frequency, these corre- 2πinξ 2πinξ \ k \ k spond to Pk,nf(ξ) := fˆ(ξ)ϕk(ξ)e 2 and Qk,nf(ξ) := fˆ(ξ)ψk(ξ)e 2 , respectively. We also used that b cn,k (f ∗ ψk · gb∗ ψk) ∗ ϕk,n(x)= cn,k (f ∗ ψk,n · g ∗ ψk,n) ∗ ϕk(x), Xk Xk which becomes obvious when written on the frequency side:

2πin(ξ1+ξ2) k 2πix(ξ1+ξ2) cn,kfˆ(ξ1)ψk(ξ1)ˆg(ξ2)ψk(ξ2)ϕk(ξ1 + ξ2)e 2 e dξ1dξ2 ˆR2 Xk 2πinξ1 2πinξ2 b k b b k 2πix(ξ1+ξ2) = cn,kfˆ(ξ1)ψk(ξ1)e 2 gˆ(ξ2)ψk(ξ2)e 2 ϕk(ξ1 + ξ2)e dξ1dξ2. ˆR2 Xk Hence, the Leibniz ruleb reduces to the boundednessb ofb the shifted paraproduct

(53) Πn(f, g)(x) := Pk (Qk,nf · Qk,ng) (x)= Pk,n (Qkf · Qkg) (x). Xk Xk The bilinear operator Πn is very similar to the classical paraproduct Π from (18), except that we need in this case shifted maximal operators and square functions 2 n 1 n |hf, ψIn i| 1/2 M (f)(x) := sup |f(y)| · χ˜In (y)dy and S (f)(x) := · 1I (x) . |I| ˆR |I| I∋x I X  Above, for a fixed interval I, we denote by In := I +n|I|, the translation of I n units to the right (or to the left, if n < 0). It is well known (a complete proof is provided in [MS13]), that these operators are bounded on every Lp space for 1

α 1 |Π (f, g)(x)|≤ 1+α |Πn(f, g)(x)| and n (1+ |n|) X 1 (log (1 + |n|))2 α s . s . kΠ (f, g)ks (1+α)s kΠn(f, g)ks (1+α)s kfkpkgkq. n (1+ |n|) n (1 + |n|) X X QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 39

Provided (1 + α) s> 1, we obtain the Lp × Lq → Ls boundedness of Πα. A similar analysis will yield the general Leibniz of Theorem 2; in the end, we will α β need to study the boundedness of Π ⊗ Π s1 s2 , Π ⊗ Π s1 s2 , Π ⊗ Π s1 s2 or Lx Ly Lx Ly Lx Ly α β p Π ⊗ Π s1 s2 . Ultimately, the range of L estimates in Theorem 2 will be determined Lx Ly α β by the “worst” term, which is Π ⊗ Π . If s0 denotes the minimum between s1 and s2, s0 Proposition 7 implies that k · k s1 s2 is subadditive. Following the arguments presented Lx Ly earlier, we have

α β s0 1 β s0 (54) Π ⊗ Π s1 s2 . Πn ⊗ Π s1 s2 . Lx Ly (1+α)s0 Lx Ly n (1+ |n|) X

Provided 1 (55) min(s ,s ) > , 1 2 1+ α

α β β the boundedness of Π ⊗ Π reduces to that of Πn ⊗ Π , with an operatorial norm that depends at most logarithmically on n. β p1 p2 q1 q2 s1 s2 If n = 0, we need to prove that Π ⊗ Π : Lx Ly × Lx Ly → Lx Ly , whenever 1 (56) s > 2 1+ β

The conditions (55) and (56) are equivalent to the constraints of s1 and s2 from the hypotheses of Theorem 2. β p1 p2 q1 q2 s1 s2 In order to establish that Π ⊗ Π : Lx Ly × Lx Ly → Lx Ly , we use restricted type interpolation, as in Proposition 10: it will be enough to prove

1 1 1 1 β p q s − s (57) kkΠ ⊗ Π (f, g)k s2 · 1 k s2 . |F | 1 · |G| 1 · |E| 1 2 , Ly E˜ Lx where F, G and E are sets of finite measure, E˜ is a major subset of E to be constructed (it is defined by (43)), while f and g are functions satisfying kf(x, ·)k p2 ≤ 1 (x) and Ly F kg(x, ·)k q2 ≤ 1 (x), respectively. Ly G The cases s2 < 1 and s2 ≥ 1 need to be treated separately. We first deal with the case F,G,E˜ β F,G,E˜ s2 < 1. In fact, we will be proving sharp estimates for Π ⊗ Π s2 , where Π is I0 Lxy I0 the same discretized, localized paraproduct introduced in (21). Before, we were using the localized paraproducts in order to deduce multiple vector- valued inequalities for Π, and p β from there, mixed norm L estimates for Π ⊗ Π. Now we work directly with ΠI(I0) ⊗ Π , and we want to prove that

s s2−ǫ s2−ǫ 1−ǫ β s 2 kkΠI(I0) ⊗ Π (f, g)k 2 · 1 ˜ k s2 . size I0 1F size I0 1G size I0 1 ˜ · |I0|, Ly E Lx E    for any f and g as above. g g g Formerly, we decomposed Πα using Fourier series in frequency, and now we are going to do the same for Πβ. In this way, we can write it as

β (58) Π (f, g)(y) := cm,l (f ∗ ψl,m · g ∗ ψl,m) ∗ ϕ˜l(y), m X Xl 40 CRISTINABENEAANDCAMILMUSCALU

1 where the Fourier coefficients satisfy |cm,l| . , and ϕ˜ ≡ 1 on supp ψl,m + (1+|m|)(1+β) l supp ψ . Since s < 1, we have l,m 2 b b

bF,G,E˜ β s2 1 2 F,G,E˜ y y s2 ΠI ⊗ Π (f, g)k s2 . Pl ΠI (Q f,Q g)(x)k s2 (I0) Lxy s2(1+β) (I0) l,m l,m Lxy m (1+ |m|) l X X ˜ 1 F,G,E y y s2 . |ΠI (Q f,Q g)(x)|k s2 . s2(1+β) (I0) l,m l,m Lxy m (1+ |m|) l X X

To deduce the last inequality, we used that ϕ˜l ≡ 1 on supp ψl,m + supp ψl,m, which further indicates that Pl(f ∗ ψl,m · g ∗ ψl,m)(x)= f ∗ ψl,m · g ∗ ψl,m(x). ˜ The vector-valued estimates for ΠF,G,E from Theoremb 29 implyb that b I0

1 1 1 ˜ ′ −ǫ ′ −ǫ F,G,E y y p q s −ǫ |Π (Q f,Q g)(x)|k s2 . size 1 2 · size 1 2 · size 1 2 I(I0) l,m l,m Lxy I0 F I0 G I0 E˜ l X    y 2 1/2 g y 2 g 1/2 g · |Q f| · χ˜I (x) p2 · |Q g| · χ˜I (x) q2 . l,m 0 Lxy l,m 0 Lxy l l X  X 

Then, because 1

˜ F,G,E y y s2 2 1 s2−ǫ 1 s2−ǫ 1 1−ǫ |ΠI(I ) (Ql,mf,Ql,mg)(x)|k s2 . (1 + log |m|) size I0 F · size I0 G · size I0 E˜ · |I0|. X 0 Lxy    l g g g

With the above estimate and the usual stopping times from Section 3.4, for each d ≥ 0, In1,n2,n3 we have collections d of dyadic intervals for which 1 |F | n1 −n1 d (1) if I ∈ I , then 2 ∼ 1F · χ˜I dx . 2 |I| ˆR |E| 1 |G| n2 −n2 d (2) if I ∈ I , then 2 ∼ 1G · χ˜I dx . 2 |I| ˆR |E| 1 In3 −n3 −Md (3) if I ∈ , then 2 ∼ 1E˜ · χ˜I dx . 2 . |I| ˆR In1,n2,n3 I Moreover, for every I0 ∈ d , there exists a certain collection (I0) associated to I0, which is selected through the stopping time. This yields a partition of I as I := I(I0), which we use in order to estimate n ,n ,n d≥0 n1,n2,n3 I ∈I 1 2 3 [ [ 0 d[

β s2 F,G,E˜ β s2 Π ⊗ Π s2 . ΠI ⊗ Π s2 Lxy (I0) Lxy n ,n ,n d≥0 n1,n2,n3 I ∈I 1 2 3 X X 0 Xd s2−ǫ s2−ǫ 1−ǫ . size I0 1F · size I0 1G · size I0 1E˜ · |I0| n ,n ,n d≥0 n1,n2,n3 I ∈I 1 2 3 X X 0 Xd    gs2 s2 g g −n1 p −n2 p −n3(1−ǫ) . 2 1 2 2 2 |I0|. n ,n ,n d≥0 n1,n2,n3 I ∈I 1 2 3 X X 0 Xd QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 41

n1 γ1 n2 γ2 n3 γ3 The sum |I0| is bounded above by (2 |F |) · (2 |G|) (2 |E|) , where n ,n ,n I ∈I 1 2 3 0 Xd 0 ≤ γ1, γ2, γ3 ≤ 1 and γ1 + γ2 + γ3 = 1. Hence we obtain

s2 s2 s2 −n1 −γ1 −n2 −γ2 β p1 q1 −n3(1−ǫ−γ3) γ1 γ2 γ3 Π ⊗ Π s2 . 2   · 2   · 2 |F | · |G| · |E| , Lxy d≥0 n1,n2,n3 X X and the series above converge (we have the freedom to choose γ1, γ2 and γ3) provided s2 + s2 + 1 − ǫ > γ + γ + γ = 1. The condition is satisfied thanks to the contribution of p1 q1 1 2 3 size I0 1E˜ which comes with an exponent arbitrarily close to 1. Finally, it is not difficult to see that all of the above imply exactly (57). β gWe still have to treat the case s2 ≥ 1: that is, we want to prove that Π ⊗ Π is bounded in the space k·k s1 s2 . Since s2 ≥ 1, we can dualize the inner norm, and using generalized Lx (Ly ) restricted type interpolation, it is enough to prove

1 1 1 β 1− (59) Π ⊗ Π (f, g)(x,y) · h(x,y)dxdy . |F | p1 · |G| q1 · |E| s1 , ˆR2

′ whenever F, G and E are sets of finite measure, E is a major subset of E to be con- structed (it is also defined by (33)), while f, g and h are functions satisfying kf(x, ·)k p2 ≤ Ly 1 (x), kg(x, ·)k q2 ≤ 1 (x), and kh(x, ·)k s′ ≤ 1 ′ (x) respectively. F Ly G 2 E Ly From here on, everything follows the same pattern:

β 1 Π ⊗ Π (f, g)(x,y) · h(x,y)dxdy . 1+β Π ⊗ Πm(f, g)(x,y) · h(x,y)dxdy , ˆR2 ˆR2 m (1 + |m|) X ′ and in fact we will need to estimate ΠF,G,E ⊗ Π . We don’t repeat the argument because I0 m it’s identical to the situation s2 < 1. The case p2 = ∞ or q2 = ∞ (which is acceptable since now s2 ≥ 1) needs an additional p1 ∞ q1 q2 justification, but the proof reduces to the boundedness of Π ⊗ Π : Lx Ly × Lx Ly → s1 q2 Lx Ly . The latter was proved in [BM15], using a similar strategy: due to restricted-type ′ interpolation, it is enough to prove a sharp estimate for the adjoint ΠF,G,H ⊗ Π 1,∗ which I0 is defined by the relation  ′ ′ ΠF,G,H ⊗ Π(f, g)(x,y) · h(x,y)dxdy = ΠF,G,H ⊗ Π 1,∗(h, g)(x,y) · f(x,y)dxdy. I0 I0 ˆR2 ˆR2 The sharp estimate concerns the operatorial norm: 

′ 1 1 F,G,H ∗,1 −ǫ ′ −ǫ 1−ǫ ′ ′ ′ q q ΠI ⊗ Π q q q q 1 1 . size I0 1H size I0 1G size I0 1F . 0 Lx Ly ×LxLy→LxLy     This ends the proof in the case n = 0, wheng the paraproductg Πn is a classicalg paraprod- uct. We are left with proving, for any |n|≥ 1, that β 2 (60) Πn ⊗ Π (f, g) s1 s2 . log(1 + |n|) kfk p1 p2 · kgk q1 q2 . Lx Ly Lx Ly Lx Ly α β Together with (54), the above inequality implies the boundedness of Π ⊗ Π . Similarly to the case n = 0, we use vector-valued restricted type interpolation, and the equivalent of (57) in this case is

1 1 1 1 β 2 p q s − s (61) Πn ⊗ Π (f, g) s2 · 1 ˜ s2 . log(1 + |n|) |F | 1 · |G| 1 · |E| 1 2 . Ly E Lx

42 CRISTINABENEAANDCAMILMUSCALU

To achieve this, we need to prove local estimates for the discretized version of Πn, which is the operator 1 (62) Πn(f, g)(x) := hf, ψIn ihg, ψIn iϕI (x). I |I| XI∈ If we look at the intervals I ∈ I so that I ⊆ I0, their translates In need not be contained inside I0; in fact, there are approximately log |n| translates of I0 that could possibly contain such a In. This is also the key observation in proving the boundedness of the shifted maximal operator Mn or of the shifted square function Sn with an operatorial norm not larger that 1 + log |n|. In order to make sure that log |n|= 6 0, we replace it by the equivalent expression loghni := 2 1/2 ♯ log 1+ n . Then given a fixed dyadic interval I0, we denote by I0, with 0 ≤ ♯ ≤ loghni the translates of I0 that contain some intervals In with I ∈ I(I0). These are actually the l  l 2 -translates of I0, for 2 ≤ n. The local estimate for Πn, corresponding to Proposition 27, reads as r0 r0 ˜ ′ −ǫ ′ −ǫ r0 F,G,E r0 r r r −ǫ r0 r0 f,g . ♯ 1 1 ♯ 1 2 1 f χ ♯ g χ ♯ , Πn,I0 ( ) size 1 F size 2 G size I0 E˜ · · ˜ 1 · · ˜ 2 r X I0  I0   I0 r1 I0 r1 0≤♯1,♯2≤loghni g g g

where r0 := min (r, 1). The localized vector-valued paraproduct will satisfy similar esti- mates. The stopping time is similar, but it is defined by more parameters; the size for the functions f and g are given by n 1 g size I f := sup |f(x)|χ˜In (x)dx, I∈I |I| ˆR I so the collections n1 from theg stopping time described in Section 3.4 will be replaced by I I I collections n1,♯1 , with 0 ≤ ♯1 ≤ loghni. An interval I1 ∈ n1,♯1 if there exists I ∈ Stock,I ⊆ ♯1 I1 so that In ⊆ I1 and

−n1−1 1 −n1 (63) 2 ≤ 1F · χ˜In dx ≤ 2 . |I| ˆR

♯1 In fact, I1 is selected prior to I1, from the set of the intervals containing such a In satisfying ♯1 the above condition. Moreover, we require I1 to satisfy a condition similar to (63), and I to be maximal among the intervals meeting these properties. Then the collection n1,♯1 (I1) will consist of I I ♯1 n1,♯1 (I1) := {I ∈ Stock : I ⊂ I1,In ⊆ I1 }. I We note that an interval I1 can be selected in several collections n1,♯1 , but however in no more than loghni of them. We also note that

n −n1 d |F | size I 1F . 2 . loghni2 , n1,♯1(I1) |E| since in this case the exceptionalg set is defined by: |F | |G| Ω := x : Mn1 >C loghni ∪ x : Mn1 >C loghni . F |E| G |E| n o n n o n The parameter d ≥ 0 is introduced as before, in order to control M 1F and M 1G: we dist (I, Ωc) have a partition I := I , where for all I ∈ I we require that 1 + ∼ 2d. d d |I| d[≥0 QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 43

The stopping time for Πn is more elaborate because we need to find a way of grouping n −n1 the intervals in I so that the shifted size size I 1F ∼ 2 , and at the same time we need to assure some disjointness that will allow us to estimate |I|. Using the disjointness of g I ♯1 I X the intervals I1 as I1 varies in n1,♯1 , we have

♯1 n1 |I1| = |I1 | . 2 |F |, I ∈I I ∈I 1 Xn1,♯1 1 Xn1,♯1

♯1 −n1−1 since every I1 ⊆ {M1F > 2 } (it satisfies the condition (63)). β We are now ready to prove the desired estimate for kΠ ⊗ Π (f, g)(x,y) · 1 (x)k s2 . For n E˜ Lxy simplicity, we illustrate the main ideas in the case s2 < 1; in the case s2 ≥ 1, we only need to rewrite the argument by employing the trilinear form. Using the above stopping time, we have

β s2 F,G,E˜ β s2 Πn ⊗ Π (f, g) · 1 ˜ s2 ≤ Π ⊗ Π (f, g) s2 . E Lxy n,♯1,♯2,I0 Lxy n1,n2,n3 d≥0 n1,n2,n3 0≤♯1,♯2≤loghni I0∈I X X X X♯1,♯2

β Now we decompose Π as in (58), and provided (1+β)s2 > 1, it will be enough to prove

F,G,E˜ y y F,G,E˜ Π (Q f,Q g) s2 . Π n,♯1,♯2,I0 l,m l,m Lxy n,♯1,♯2,I0 l X y 2 1/2 y 2 1/2 · |Q f| · χ˜ ♯1 p2 · |Q g| · χ˜ ♯2 p2 , l,m I0 Lxy l,m I0 Lxy l l X  X  with an operatorial norm

1 1 1 F,G,E˜ ′ −ǫ ′ −ǫ −ǫ p2 p2 s2 Πn,♯ ,♯ ,I . size ♯1 1F · size ♯2 1G · size I0 1E˜ . 1 2 0 I0 I0    This follows from the boundedness g of the shiftedg square functiong (a certain power of log 1+ |m| will appear, but it doesn’t affect the summation in m) and the usual vector-valued ˜ estimates for the paraproduct ΠF,G,E , which localizes well. n,♯1,♯2,I0 In the end, we will have

β s2 s2−ǫ Πn ⊗ Π (f, g) · 1E˜ s2 . size ♯1 1F Lxy I0 n1,n2,n3 d≥0 n1,n2,n3 0≤♯1,♯2≤loghni I0∈I X X X X♯1,♯2 

s2−ǫ 1−ǫ g · size ♯2 1G · size I0 1E˜ |I0|. I0 Since we have control over all the sizes and over |I |, we can easily obtain the g g I0 0 inequality sP2 s2 s2 β s2 2 p p s −1 Πn ⊗ Π (f, g) · 1 ˜ s2 . loghni |F | 1 · |G| 2 · |E| 1 , E Lxy which ends the proof because this is exactly the estimate (61). We want to emphasize however that without using the vector-valued point of view (and vector-valued restricted-type interpolation), it is difficult to remove the constraint that 1 α β s1 > 1+β , which is implied by splitting both Π and Π from the beginning, as in (54). s2 Also, in the case s2 ≥ 1, dualizing through L is not enough, and we have to bring forth the trilinear form.  44 CRISTINABENEAANDCAMILMUSCALU

8. Proof of Interpolation Proposition 10

Proof. The tuples (p,q,s) and (r1, r2, r) are fixed, but (s1,s2, s˜) are allowed to vary in a neighborhood of (p,q,s). We will decompose both f~ = {fk}k and ~g = {gk}k into pieces that we can control: ~ ~ ~ f = fm1 = f · 1 m ~ m +1 , x:2 1 ≤ f r <2 1 m m ℓ 1 X1 X1 n o and similarly

~g = ~gm2 = ~g · 1 m m +1 . 2 2 ≤ ~g r <2 2 m m ℓ 2 X2 X2 n o Z ~ r m1 We note that for every m1 ∈ , kfm1 kℓ 1 ∼ 2 and is supported on a set of finite measure. r1 1/r1 r2 1/r2 For simplicity, we will assume that k fk p = k gk q = 1. p Given a function ϕ, we will use the distribution function dϕ for estimating the L norm P +  P  of ϕ. We recall that dϕ is a function from R to [0, ∞ ], defined by

dϕ(λ) := x : ϕ(x) > λ . Then we have, for any 0

p p−1 (64) kϕkp = p λ dϕ(λ)dλ. ˆR+ We will be using a discrete variant of the formula above: p np n kϕkp ∼ 2 dϕ(2 ). Z nX∈ The assumptions on the Lp ℓr1 and Lq ℓr2 norms of f~ and ~g respectively translate into   r 1/r p (65) f 1 1 ∼ 2m1pd (2m1 ) ∼ 1 k p f~ ℓr1 k m1 X  X and r 1/r q (66) g 2 2 ∼ 2m2qd (2m2 ) ∼ 1. k q ~g ℓr2 k m2 X  X

For T (f, g), we have the estimate r 1/r s ns n T (fk, gk) ∼ 2 d r 1/r (2 ). s T (f ,g ) k n k k k X  X P r  However, since r < 1, the k · kℓ is not subadditive, and this is a property that plays an important role in the classical proofs of interpolation theorems. We will use instead the r subadditivity of k · kℓr : r 1/r s r s/r ns/r r n (67) T (fk, gk) s = T (fk, gk) s/r ∼ 2 d (2 ). k T (fk,gk) k k n X  X X P r We need to estimate the distribution function of k T (fk, gk) . First, we note that n n r r d (2 ) ≤ d P (cn,m1,m2 2 ), T (fk,gk) T (fk,m ,gk,m ) k m ,m k 1 2 P X1 2 P where c > 0 will be chosen later, with the property that c ∼ 1. n,m1,m2 m1,m2 n,m1,m2 P QUASI-BANACH VALUED INEQUALITIES VIA THE HELICOIDAL METHOD 45

Condition (13) generalizes to a weak-type condition:

1 r 1/r r r T F , G T F , G K F r1 G r2 ( k k) s,˜ ∞ = ( k k) s˜ ≤ s1,s2,s˜ k kℓ s k kℓ s X  X r ,∞ 1 2 k k

r r whenever kF kℓ 1 ∼ A11E1 and kGkℓ 2 ∼ A21E2 . This further implies that

s˜ − s˜ s˜ s˜ d r (λ) ≤ K λ r kF k r1 kGk r2 . s1,s2,s˜ ℓ s ℓ s k T (Fk,Gk) 1 2 P ~ ~ We will apply this to the functions fm1 and ~gm2 . We also note that, due to the way fm1 and ~gm2 were defined, we have that

1 1 m m m m ~ r 1 1 s1 r 2 2 s2 kfm1 kℓ 1 . 2 d ~ (2 ) , k~gm2 kℓ 2 . 2 dk~gk r (2 ) . s1 kfkℓr1 s2 ℓ 2

Hence,

s˜ 1 s˜ 1 s˜ n s˜ n − m1 m1 m2 m2 r r s1 s2 d (cn,m1 ,m2 2 ) . Ks ,s ,s˜ (cn,m1,m2 2 ) · 2 d ~ (2 ) · 2 dk~gk r (2 ) , T (f ,g ) 1 2  kfkℓr1   ℓ 2  Pk k,m1 k,m2

where in fact the tuple (s1,s2, s˜) depends on n,m1,m2 and is to be chosen later. For simplicity, we don’t write down this dependency, but it is an important step in our proof. ˜ ˜ s s˜ All of the above imply that, for K so that K := sup(s1,s2,s˜) Ks1,s2,s˜, we have

s n(s−s˜) − s˜ s˜ s˜ r r s r m1s˜ m1 s m2s˜ m2 s s ˜ r 1 2 T (fk, gk) . K 2 cn,m1,m2 2 d ~ r (2 ) 2 dk~gk r2 (2 ) . r kfkℓ 1 ℓ k n m1,m2 X X X

We clearly need to make use of conditions (65) and (66), and of the H¨older condition 1 = 1 + 1 . We note that the above expression can be eventually written as s˜ s1 s2

s˜ 1 1 ns 1 1 ns − s˜ − m1p− s˜ − m2q− r p s1 ( r ) q s2 ( r ) cn,m1,m2 2   2   n m1,m2 X X s˜ s˜ s˜ s˜ m1p m2q s1 m1 s1 s2 m2 s2 · 2 d ~ (2 ) 2 dk~gk r (2 ) kfkℓr1 ℓ 2 s˜ 1 1 ns 1 1 ns − s˜ − m1p− s˜ − m2q− r p s1 ( r ) q s2 ( r ) = cn,m1,m2 2   2   n m p− ns ,m q− ns X 1 rX 2 r s˜ ns ns 1 ns ns s˜ s˜ ns ns 1 ns ns s˜ m1p− + m1p− + m2q− + m2q− + s1 ( r r ) p ( r r ) s1 s2 ( r r ) q ( r r ) s2 · 2 d ~ (2 ) · 2 dk~gk r (2 ) . kfkℓr1 ℓ 2

Now we will turn to our advantage the freedom to choose the triples (s1,s2, s˜) and the numbers cn,m1,m2 . First we fix ǫ small and we choose (s1,s2, s˜) sufficiently close to (p,q,s) so that 1 1 ns 1 1 ns np np s˜ − m p − , s˜ − m q − ≤−ǫ max m p − , m q − . p s 1 r q s 2 r 1 r 2 r  1   2       

We then choose cn,m1,m2 of the form

ǫr np ǫr np − 3˜s m1p− r − 3˜s m2q− r cn,m1,m2 := γ2 · 2 , where γ is so that c = 1, and it depends only on (p,q,s) and ǫ. m1,m2 n,m1,m2 P 46 CRISTINABENEAANDCAMILMUSCALU

np np Making the change of variablesm ¯ 1 := m1p − r andm ¯ 2 := m2q − r , we obtain r 1/r s − ǫ max m¯ , m¯ ˜ 3 1 2 T (fk, gk) s . Kγ 2   m¯ ,m¯ Xk X1 2 s˜ ns 1 ns s˜ s˜ ns 1 ns s˜ m¯ 1 + m¯ 1+ m¯ 2+ m¯ 2+ s1 ( r ) p ( r ) s1 s2 ( r ) q ( r ) s2 · 2 d ~ (2 ) 2 dk~gk r (2 ) . kfkℓr1 ℓ 2 n X Applying H¨older’s inequality in the last line, and thanks to identities (65) and (66), we obtain the Lp × Lq → Ls strong-type estimate. Concerning the constant K, we can see that

Kp,q,s ∼ sup Ks1,s2,s˜. (s1,s2,s˜)∈V(p,q,s)  The proof of Proposition 11 is similar, and the fact that we allow for arbitrary measures ~ rj0 is of no consequence. In this situation, we use the subadditivity of kT (f,~g)kLR , for some index 1 ≤ j0 ≤ N as in Proposition 7. References [BM15] Cristina Benea and Camil Muscalu. Multiple vector valued inequalities via the helicoidal method. https://arxiv.org/abs/1511.04948, 2015. [CM97] R. Coifman and Y. Meyer. , Calder´on-Zygmund Operators and Multilinear Operators. Cambridge University Press, 1997. [CUMP04] D. Cruz-Uribe, J. M. Martell, and C. P´erez. Extrapolation from A∞ weights and applications. J. Funct. Anal., 213(2):412–439, 2004. [DPO15] Francesco Di Plinio and Yumeng Ou. Banach-valued multilinear singular . http://arxiv.org/abs/1506.05827, 2015. Online; accessed June 2016. [GM04] Loukas Grafakos and Jos Mar´ıa Martell. Extrapolation of weighted norm inequalities for mul- tivariable operators and applications. J. Geom. Anal., pages 19–46, 2004. [Ken04] Carlos Kenig. On the local and global theory for the KP 1 equation. Ann. I. H. Poincar´e, pages 827–838, 2004. [LT99] Michael Lacey and Christoph Thiele. On Calder´on’s conjecture. Ann. of Math. (2), 149(2):475– 496, 1999. [MPTT04] Camil Muscalu, Jill Pipher, Terence Tao, and Christoph Thiele. Bi-parameter paraproducts. Acta Mathematica, pages 269–296, 2004. [MPTT06] Camil Muscalu, Jill Pipher, Terence Tao, and Christoph Thiele. Multi-parameter paraproducts. Rev. Mat. Iberoamericana, pages 963–976, 2006. [MS13] Camil Muscalu and Wilhem Schlag. Classical and Multilinear . Cambridge University Press, 2013. [Sil14] Prabath Silva. Vector valued inequalities for families of bilinear Hilbert transforms and appli- cations to bi-parameter problems. J. Lond. Math. Soc., pages 695–724, 2014.

Cristina Benea, Universite´ de Nantes, Laboratoire Jean Leray, Nantes 44322, France E-mail address: [email protected]

Camil Muscalu, Department of Mathematics, Cornell University, Ithaca, NY 14853, USA E-mail address: [email protected]