1 Historical origins of the Hardy spaces Hp

The Hardy spaces in one variable have their original setting in . They first appeared as spaces of holomorphic functions and were introduced with the aim of characterizing boundary values of holomorphic functions on the D := {|z| < 1}. Namely, let us look at the following problem: what are the possible functions S1 → C arising as boundary values of some F : D → C? This question, as just stated, is too vague: due to the lack of compactness of D, holomorphic functions defined on D could exhibit a wild behaviour as we approach the boundary (for instance, we can prescribe arbitrary values of F on any discrete subset of D). In order to obtain a meaningful notion of boundary value, it is natural to impose integrability conditions on our functions F . As a motivation of the forthcoming definitions, let us make a heuristic remark: if the trace of F on ∂D = S1 is some complex function f ∈ Lp(S1), for some 1 ≤ p ≤ ∞, then F (which is holomorphic and thus harmonic) is given by the Poisson integral of f. In polar coordinates we have the formula Z iθ i(θ−η) iη F (re ) = Pr(e )f(e ) dη. S1

R iη The is everywhere positive and satisfies S1 Pr(e ) dη = 1 for any r, so (by Young’s inequality on the group S1)

kF (r·)kLp(S1) ≤ kfkLp(S1) and in particular all the norms in the left-hand side remain bounded as r ↑ 1.

In 1915 Hardy observed that, for any holomorphic function F : D → C, the map r 7→ kF (r·)kLp is nondecreasing (for an arbitrary 0 < p ≤ ∞). These observations lead us to define the   p H (D) := F : D → C holomorphic with lim kF (r·)kLp(S1) < +∞ . r↑1

When p > 1, using the weak* compactness of Lp(S1), it is not difficult to show that any F ∈ Hp(D) is given by the Poisson integral of a complex-valued function f ∈ Lp(S1) satisfying fb(k) = 0 for all k < 0. Conversely, given any such f, its Poisson integral lies in Hp(D). Moreover, one can show that F (r·) → f in Lp(S1) and lim F (reiθ) = f(eiθ) for a.e. θ, r↑1 so that f deserves to be regarded as the set of boundary values of F (we mention that for a.e. θ one has even a nontangential convergence of F to f(eiθ)). This settles the problem for 1 < p ≤ ∞. Let us also remark that the condition fb(k) = 0 (for all k < 0) amounts to saying that =(f), up to constants, equals the Hilbert-Riesz transform of −<(f). When 1 < p < ∞ the Hilbert-Riesz transform maps Lp(S1)

1 into itself, so any function in Lp(S1) arises as the real part of the trace of some element of Hp(D). The case of 0 < p ≤ 1 is more difficult. F. Riesz, in a paper published in 1923, introduced the notation Hp(D) for these spaces of holomorphic function (the letter H stands of course for Hardy) and proved many interesting properties, such as the following factorization theorem.

Theorem 1. Any F ∈ Hp(D) can be written as F = BG, for suitable holomorphic functions B,G : D → C such that |B| ≤ 1, G 6= 0 everywhere and G ∈ Hp(D)(B is the so-called Blaschke product associated to the zeros of F ).

This theorem enabled him to prove the existence of a trace f ∈ Lp(S1) such that we have again all the convergence results mentioned before for the case p > 1: the trick 1 is that, as G 6= 0 everywhere, one can take a k-th root of G (for an arbitrary k > p ) and, observing that G1/k ∈ Hkp(D) and B ∈ H∞(D), we get back to the previous case. Again, when p = 1, it can be shown that the possible traces are precisely the complex-valued functions f ∈ L1(S1) satisfying fb(k) = 0 for all k < 0. As before, the possible real values of traces of functions form the set

H1(S1) = f ∈ L1(S1): Rf ∈ L1(S1)

(here R denotes the Hilbert-Riesz transform). But R does not map L1(S1) into itself any longer, so this set is a proper subspace of L1(S1).

2 These results still hold in the upper half-plane R+ := {z = x + iy : y > 0}, replacing S1 with R (see [2]), leading to the definition of H1(R). Later, in 1960, Stein and Weiss introduced the systems of conjugate harmonic func- tions in several variables, inspiring the correct definition of Hardy spaces in higher dimension. The first characterization avoiding conjugate functions was provided by Burkholder, Gundy and Silverstein in 1971: they proved that a holomorphic function belongs to Hp if and only if the nontangential maximal function of its real part lies in Lp. The importance of this result lies in the fact that it allows to decide the membership of a function f to Hp by looking just at f itself. In 1972 Fefferman and Stein, in a single pioneering paper, provided new real char- acterizations of the Hardy spaces, introducing different useful maximal functions and showing that the Poisson kernel can be replaced essentially by any other ker- nel. In this paper Fefferman and Stein also proved that singular integrals map Hardy spaces to themselves (and in particular H1 to L1), as well as the duality (H1)∗ = BMO. The Littlewood–Paley characterization of these spaces was first given by Peetre, while the atomic decomposition was obtained by Coifman (in one dimension) and Latter (in arbitrary dimension).

2 2 Equivalent definitions and basic properties of H1(Rn)

We now introduce the H1(Rn), with a strong emphasis on the modern real-variable point of view outlined in the last part of [.]

This important space can be characterized in many useful ways: indeed, H1(Rn) is the space of all functions in L1(Rn) satisfying one of the equivalent definitions provided by Theorem 2 ((8) being the closest to the historical one).

Before stating the theorem, let us recall that the S(Rn) is a Fr´echet space, with the following increasing of (semi)norms:

|α| 2 k/2 X ∂ ψ kψkk := sup (1 + |x| ) α (x), k ≥ 0. x∈ n ∂x R |α|≤k

n R Theorem 2. Fix any ϕ ∈ S(R ) with ϕ(x) dx 6= 0. There exists an N ≥ 0 (independent of ϕ) such that the following are equivalent, for a function f ∈ L1(Rn): v 1 n (1) the vertical maximal function Mϕf(x) := supt>0 |ϕt ∗ f| (x) lies in L (R ); (2) the conical maximal function

c Mϕf(x) := sup |ϕt ∗ f| (y) t>0, y∈Bt(x)

lies in L1(Rn); (3) the tangential maximal function  −n−1 t |y − x| Mϕf(x) := sup |ϕt ∗ f| (y) 1 + t>0, t n y∈R lies in L1(Rn); (4) the grand maximal function

n GMf(x) := sup {|ψt ∗ f| (x) | t > 0, ψ ∈ S(R ), kψkN ≤ 1} lies in L1(Rn); (5) the similar grand maximal function

0 ∞ GM f(x) := sup {|ϕt ∗ f| (x) | t > 0, ψ ∈ Cc (B1(0)), k∇ψkL∞ ≤ 1} lies in L1(Rn);

(6) there exists an atomic decomposition, namely there exist λk ≥ 0 and ∞-atoms ak (see Definition 4) such that ∞ ∞ X 1 n X f = λkak in L (R ), λk < +∞; k=0 k=0

3 v 1 n (7) the vertical maximal function with the Poisson kernel, i.e. MP f, lies in L (R ) (notice that P (x) = F −1(e−t|ξ|) = cn 6∈ S( n); (1+|x|2)(n+1)/2 R

−1  ξj  1 n (8) the tempered distributions Rjf := F −i |ξ| fb(ξ) belong to L (R );

˙ 0 n (9) f belongs to the homogeneous Triebel-Lizorkin space F1,2(R ).

Each of the preceding statements defines also a on H1(Rn): (1) defines the v norm Mϕf 1 (and similarly for (2), (3), (4), (5), (7)), (6) induces the norm P L inf k λk (the infimum ranging among all the possible decompositions), (8) provides Pn 1 n the norm kfk 1 + kRjfk 1 and (9) defines kfk ˙ 0 (which is a norm on L ( )∩ L j=1 L F1,2 R ˙ 0 n F1,2(R )). The proof of this theorem is scattered across the next sections. We will show the following diagram of implications (with the corresponding bounds on the induced norms):

(7) (8)

ϕ

(4) (6) for a suitable

(1) (2) (3) for a suitable

ϕ (5)

We left out (9) in the diagram, since its equivalence with the other definitions is slightly involved and invokes the vector-valued space H1(`2), which will be intro- duced in Section 8. Similarly, as we will see in Section 4, the proof of (1) ⇒ (2) is quite circuitous and uses a particular refinement of the proofs of (2) ⇒ (3) ⇒ (4).

Let us rely on definition (1) as for now, i.e. let H1(Rn) denote the space of functions f ∈ L1( n) satisfying (1) and let kfk := Mv f . A first basic question is R H1 ϕ L1 ∞ n 1 n whether Cc (R ) is included in H (R ). Surprisingly, this property (which holds for most of the common functional spaces) fails for H1(Rn), as the next proposition shows.

1 n R Proposition 3. If f ∈ H (R ), then f(x) dx = 0. R Proof. Assume by contradiction that m := f(x) dx 6= 0. Choose any x0 6= 0 such

4 that c := |ϕ| (x0) 6= 0. Then we can find R > 0 such that Z Z m cm f(x) dx ≥ , kϕkL∞ |f| (x) dx ≤ . 2 n 4 BR(0) R \BR(0)

For any z ∈ Rn close to 0 and any large r > 0 we have Z n −1 cm r |ϕr ∗ f| (r(x0 + z)) ≥ ϕ(x0 + z − r y)f(y) dy − BR(0) 4 Z Z −1 cm ≥ c f(y) dy − ϕ(x0 + z − r y) − ϕ(x0) |f| (y) dy − BR(0) BR(0) 4 cm −1 cm cm ≥ − kfkL1 max ϕ(x0 + z − r y) − ϕ(x0) − ≥ 4 y∈BR(0) 8 8

−1 cm provided that kfk 1 max |ϕ(x + z − r y) − ϕ(x )| ≤ , which holds if L y∈BR(0) 0 0 8 −1 |x0| |z| <  and r >  for some small . We can assume that  < 2 . For such z, r it holds cm Mv f(r(x + z)) ≥ r−n |r(x + z)|−n . ϕ 0 8 & 0 −1 But E := {r(x0 + z) | |z| < , r >  } is an open cone minus a bounded set, so Z Z Z v v −n Mϕf(x) dx ≥ Mϕf(x) dx & |x| dx = +∞, E E contradicting the fact that f ∈ H1(Rn).

As shown by the next proposition, the mean-zero property is the only requirement ∞ n 1 n which a function in Cc (R ) needs to satisfy in order to be in H (R ).

Definition 4. For any 1 < p ≤ ∞, a p-atom is a function a ∈ Lp supported in some ball B, with zero mean and

1/p0 kakLp |B| ≤ 1.

We remark that (for 1 < p < ∞) the last condition can be rewritten as

 Z 1/p |B| − |a|p ≤ 1. B By H¨older’sinequality, any q-atom a is also a p-atom for every 1 < p ≤ q and 1 kakL1 ≤ 1. We now show that the H -norm is bounded, as well.

v 0 Remark 5. Mϕf . Mf pointwise, since letting h(x) := max|x0|≥|x| |ϕ| (x ) we have (noticing that h is radial and that the superlevel sets {h > s} are either open balls

5 or empty, for all s > 0) Z −n −1 |ϕt ∗ f| (x) ≤ t h(t y) |f| (x − y) dy Z = h(y) |f| (x − ty) dy Z ∞ Z = |f| (x − ty) dy ds 0 {h>s} Z ∞ ≤Mf(x) |{h > s}| 0 Z =Mf(x) h(y) dy . Mf(x),

R v as h(y) dy is finite. The same proof with P in place of f shows that MP f ≤ Mf.

Proposition 6. If a is a p-atom supported in B, then a ∈ H1(Rn) and

0 kak kak |B|1/p ≤ 1. H1(Rn) . p The implied constant depends on n, p and ϕ.

Proof. Let B = Br(x0). For x ∈ B2r(x0) we use the last remark to estimate

v Mϕa(x) . Ma(x), which gives (by H¨older’sinequality and Hardy-Littlewood maximal inequality) Z Z v 1/p0 1/p0 Mϕa(x) dx . Ma(x) dx ≤ |B2r(x0)| kMakLp . |B| kakLp . B2r(x0) B2r(x0)

For x 6∈ B2r(x0) we use instead the mean-zero property: Z ϕt ∗ a(x) = (ϕt(x − y) − ϕt(x − x0))a(y) dy.

By the mean value theorem, |ϕt(x − y) − ϕt(x − x0)| ≤ r |∇ϕt| (x − z) for some z on the segment joining x0 to y. So |z − x0| ≤ r, thus

x − z  x − z −n−1 |∇ϕ | (x − z) = t−n−1 |∇ϕ| t−n−1 |x − x |−n−1 t t . t . 0 since ϕ ∈ S(Rn). Hence, Z Z v −n−1 1/p0 Mϕa(x) dx . r kakL1 |x − x0| dx . |B| kakLp . n n R \B2r(x0) R \B2r(x0)

1 n Proposition 7. For any f ∈ H (R ) we have kfkL1 . kfkH1 .

6 Proof. We assume (without loss of generality) that R ϕ(x) dx = 1. Recall that 1 n limh→0 kf(· + h) − fkL1 = 0 for all functions f ∈ L (R ). Thus, Z Z

kϕt ∗ f − fk 1 = ϕ(y)(f(x − ty) − f(x)) dy dx L ZZ ≤ |ϕ| (y) |f(x − ty) − f(x)| dx dy → 0 as t → 0, by the dominated convergence theorem: indeed, the inner integral is bounded by 2 kfkL1 |ϕ| (y) and tends to 0 for all y, by the aforementioned property of functions in L1( n). So kfk = lim kϕ ∗ fk ≤ Mv f . R L1 t→0 t L1 ϕ H1

Proposition 8. The space H1(Rn) is a .

1 n Proof. If (fj) is a Cauchy sequence in H (R ), then we have

kfj − fkkL1 . kfj − fkkH1 → 0 as j, k → ∞

1 n by Proposition 7, so (fj) is a Cauchy sequence in L (R ). Hence, fj → f for some f ∈ L1(Rn). For any x ∈ Rn we have

v |ϕt ∗ f| (x) = lim |ϕt ∗ fj| (x) ≤ lim inf Mϕfj(x), j→∞ j→∞

v v so Mϕf(x) ≤ lim infj→∞ Mϕfj(x) and, by Fatou’s lemma, we deduce

v v kfk 1 = Mϕf 1 ≤ lim inf Mϕfj 1 < +∞. H L j→∞ L

1 n 1 n So f ∈ H (R ). Moreover, since f − fj = limk→∞(fk − fj) in L (R ), the same argument shows that

kf − fjk 1 ≤ lim inf kfk − fjk 1 . H k→∞ H But the right-hand side can be made small at will, by taking j large enough (since 1 n 1 n (fj) is a Cauchy sequence in H (R )). This proves that fj → f in H (R ).

1 n Proposition 9. If (fj)j∈N is a bounded sequence in H (R ) then, up to extracting 1 n ∗ 0 n a subsequence, there exists f ∈ H (R ) such that fj * f in S (R ).

This result is related to the fact that H1(Rn) is a . Notice that the same 1 n ∗ R statement is false in L (R ): for instance, it is easy to see that ϕt * ( ϕ(x) dx)δ. In general, a distributional limit of functions in L1(Rn) is a finite which can possess a singular part.

Proof. We assume (without loss of generality) that R ϕ(x) dx = 1. Recall that the n dual space of C0(R ), the space of continuous functions which are infinitesimal at n ∗ n 1 n infinity, is C0(R ) = M(R ), the space of finite (signed) measures. L (R ) is isometrically embedded into M(Rn): a function g ∈ L1(Rn) can be regarded as the finite measure g dx.

7 1 n n By Proposition 7, (fj) is bounded in L (R ) as well. Since C0(R ) is separable, by Banach-Alaoglu any closed ball in its dual is weakly* sequentially compact, so there n exists a subsequence, which we still denote (fj), and a measure µ ∈ M(R ) such that ∗ n ∗ fj dx * µ in C0(R ) . We claim that µ is absolutely continuous with respect to the Lebesgue measure. Indeed, for any x ∈ Rn and any t > 0 we have Z Z ϕt ∗ fj(x) = ϕt(x − y) fj(y) dy → ϕt(x − y) dµ(y) =: ϕt ∗ µ(x).

Arguing as in the previous proof, we deduce

v sup |ϕt ∗ µ| (x) ≤ lim inf Mϕfj =: g. t>0 j→∞ n n It is easy to check that ϕt(−·) ∗ ρ → ρ in C0(R ), for any ρ ∈ C0(R ). This ∗ implies, using Fubini’s theorem, that ϕt ∗ µ dx * µ. Let now E be a Borel set with |E| = 0. We can find a decreasing sequence of open sets (Vk) such that E ⊆ ∩kVk and |Vk| = 0. By weak* convergence we have Z Z |µ| (E) = |µ| (Vk) ≤ lim inf |ϕt ∗ µ| (x) dx ≤ g(x) dx. t→0 Vk Vk But g ∈ L1( n) (by Fatou’s lemma, since lim inf Mv f < +∞), so taking R j→∞ ϕ j L1 the limit as k → ∞ we deduce Z |µ| (E) ≤ lim g(x) dx = 0. k→∞ Vk Hence the claim is proved, i.e. µ = f dx for some f ∈ L1(Rn). We deduce that 1 n ∗ 0 n f ∈ H (R ) as in the previous proof. The convergence fj * f in S (R ) follows n n from the fact that S(R ) injects continously into C0(R ).

3 H1 → H1 boundedness of Calder´on-Zygmund operators

In this section we will take for granted Theorem 2 (with an abuse of notation, we will denote by k·kH1 any of the equivalent norms introduced above) and we will show why H1(Rn) is the good replacement of L1(Rn) from the point of view of . Namely, its norm has the same behaviour as the L1-norm: for any λ > 0,

kfλkH1 = kfkH1

(to be precise, this identity becomes kfkH1 kfλkH1 . kfkH1 if we use the norm given by (9), with an implied constant independent of f and λ). Furthermore, Calder´on- Zygmund operators map H1(Rn) into itself: this property holds also for Lp(Rn) with 1 < p < ∞, but it dramatically fails for L1(Rn).

8 Let us mention that, as another confirmation of the appropriateness of Hardy spaces, if one carries over the theory into the general case of Hp(Rn) spaces then, for 1 < p < ∞, they collapse to Lp(Rn) (for which many important results in harmonic analysis already hold).

Theorem 10. Let K : Rn \{0} → R be a Calder´on-Zygmund kernel, i.e. a measur- able function satisfying (for some finite constants A, B > 0) −n •| K| (x) ≤ A |x| for all x ∈ Rn \{0} R n n • |x|>2|y| |K(x − y) − K(x)| dx ≤ B for all y ∈ R R n • r<|x|

K ∗ f := lim K ∗ f →0 exists in L1(Rn). We have the estimate

kK ∗ fkL1 ≤ C(n, A, B) kfkH1 .

2 Proof. Recall that K ∈ L still satisfies the above conditions (with B possibly 1 n replaced by C(n)B) and that Kb ≤ C(n, A, B). Fix any f ∈ H (R ): by L∞ the characterization involving the atomic decomposition, we can find λk ≥ 0 and P P ∞-atoms ak with f = k λkak and k λk . kfkH1 . It suffices to prove the thesis for ∞-atoms: once this is done, for any  > 0 X X kK ∗ fkL1 ≤ λk kK ∗ akkL1 . λk . kfkH1 . k k

1 n Moreover, (K ∗ f) is Cauchy in H (R ) as  → 0: indeed, for an arbitrary N,

lim sup kK ∗ f − K0 ∗ fkL1 ,0→0 X X ≤ lim sup λk kK ∗ ak − K0 ∗ akkL1 + lim sup λk kK ∗ ak − K0 ∗ akkL1 ,0→0 ,0→0 k≤N k>N X X ≤ 0 + lim sup λk (kK ∗ akkL1 + kK0 ∗ akkL1 ) . λk, ,0→0 k>N k>N which can be made arbitrarily small by letting N → +∞. Thus K ∗ f converges in L1(Rn) and the limit satisfies the same estimate.

Let now a be an ∞-atom supported in BR(x0). Recall that

kK ∗ akL2 ≤ C(n, A, B) kakL2 ≤ C(n, A, B)

2 and that lim→0 K ∗ a exists in L . Using H¨older’sinequality we infer that (K ∗ a)1B (x ) satisfies 2R 0

(K ∗ a)1B (x ) ≤ C(n, A, B) 2R 0 L1

9 and converges in L1 as  → 0. Moreover, using the mean-zero property of a, for any n x ∈ R \ B2R(x0) we can write Z n K ∗ a(x) = (K(x − y) − K(x − x0)) a(y) dy , Rn so that Z Z Z n n n |K ∗ a| (x) dx ≤ |K(x − y) − K(x − x0)| |a(y)| dy dx n n n R \B2R(x0) R \B2R(x0) R Z Z n n = |K(x − y) − K(x − x0)| |a(y)| dx dy n n R R \B2R(x0)

≤B kakL1 ≤ B.

Adding this to the preceding inequality we deduce that kK ∗ akL1 ≤ C(n, A, B). 1 Finally, (K ∗ a)1 n is Cauchy in L as well, since  R \B2R(x0)

(K ∗ a)1 n = (K 0 ∗ a)1 n  R \B2R(x0)  R \B2R(x0) whenever , 0 ≤ R.

The multiplier version of Calder´on-Zygmund theorem holds as well, with the follow- ing statement.

Theorem 11. Assume m ∈ C∞(Rn \{0}) satisfies |α| |α| ∂ m sup |ξ| α (ξ) < +∞ ξ∈Rn\{0} ∂ξ for any α ∈ Nn. Then, for any f ∈ H1(Rn), the distribution mfb ∈ L∞(Rn) lies in F(L1(Rn)) and

−1 F (mfb) . kfkH1 . L1 P Proof. Take an atomic decomposition f = λkak as in the preceding proof and n fix a dyadic partition of unity (ψ`)`∈Z in R \{0}. Recall that the kernels KN := −1 PN  n F `=−N ψ`m ∈ S(R ) satisfy the H¨ormandercondition for some constant B independent of N and have equibounded Fourier transforms. Thus we can argue as in the previous proof (without the need of truncating the kernel KN , since it is a Schwartz function) and we get

kKN ∗ akkL1 . 1. −1 2 n But, by Plancherel’s theorem, KN ∗ ak → F (mbak) in L (R ) as N → ∞, thus F −1(ma ) ∈ L1( n) and F −1(ma ) 1 bk R bk L1 . (by Fatou’s lemma, since a subsequence K ∗a converges a.e. to F −1(ma )). Thus Nj k bk the limit X −1 g := λkF (mbak) k 1 n P exists in L (R ) and satisfies kgkL1 . kfkH1 , as well as gb = k λk(mbak) = mfb.

10 Actually, in the preceding theorems we can easily upgrade the H1 → L1 boundedness to H1 → H1.

Corollary 12. Under the hypotheses of Theorem 11, for any f ∈ H1(Rn) we have

−1 1 n −1 F mfb ∈ H (R ), F mfb . kfkH1 . H1

Proof. By the characterization of H1(Rn) using Riesz transforms, it suffices to show −1 1 n 1 that RjF (mfb) ∈ L (R ) with an estimate on its L -norm (for any 1 ≤ j ≤ n). But   −1 −1 ξj RjF (mfb) = F −i m(ξ)fb(ξ) |ξ| and the multiplier still satisfies the hypotheses of Theorem 11.

Corollary 13. Under the hypotheses of Theorem 10, for any f ∈ H1(Rn) we have 1 n K ∗ f ∈ H (R ) and the limit

K ∗ f := lim K ∗ f →0 exists in H1(Rn), with the estimate

kK ∗ fkH1 ≤ C(n, A, B) kfkH1 .

1 n Proof. From Corollary 12 we know that, for any 1 ≤ j ≤ n, Rjf ∈ H (R ) with kRjfkH1 . kfkH1 . Moreover,   −1 ξj Rj(K ∗ f) = F −i Kb(ξ)fb(ξ) = K ∗ (Rjf), |ξ| so, by the conclusion of Theorem 10, (Rj(K ∗ f)) is Cauchy as  → 0. As a 1 n consequence, (K ∗ f) is Cauchy in H (R ).

4 Equivalence of some maximal functions

The goal of this section is to prove the equivalence among the norms defined by (1), (2), (3) and (4). Trivially, we have v Mϕf . GMf pointwise (with the implied constant depending only on ϕ), so Mv f GMf ϕ . L1 . kGMfkL1 and (4) ⇒ (1) hold as well. We also remark the following pointwise inequalities:

v c n+1 t Mϕf ≤ Mϕf ≤ 2 Mϕf

11 −n−1 n+1  |y−x|  pointwise (the second inequality follows from the fact that 2 1 + t ≥ 1 whenever y ∈ Bt(x)). Let us now turn to the first nontrivial inequality, namely the fact that Mt f ϕ L1 . Mc f , which will give (2) ⇒ (3). ϕ L1

Lemma 14. For any x ∈ Rn we have

(n+1)/n t  c n/(n+1) Mϕf(x) ≤ M Mϕf (x).

c Proof. The key observation is the fact that |ϕt ∗ f| (y) ≤ Mϕf(z) whenever z ∈ Bt(y) (since z ∈ Bt(y) is equivalent to y ∈ Bt(z)). From this it follows that Z n/(n+1) 1 c n/(n+1) |ϕt ∗ f| (y) ≤ (Mϕf) (z) dz |Bt(y)| Bt(y) Z Bt+|y−x|(x) ≤ − (Mc f)n/(n+1)(z) dz, |B (y)| ϕ t Bt+|y−x|(x) which gives  −n n/(n+1) |y − x| c n/(n+1) |ϕt ∗ f(y)| 1 + ≤ M M f (x). t ϕ

n+1 Raising both sides to the power n we obtain the thesis.

Corollary 15. Using the L(n+1)/n-boundedness of the Hardy-Littlewood maximal function, we deduce

(n+1)/n t c n/(n+1) c Mϕf 1 ≤ M Mϕf . Mϕf 1 . L L(n+1)/n L

Now we prove that the grand maximal function GMf is controlled pointwise by Mt f, which will trivially give (3) ⇒ (4) and kGMfk Mt f . The choice of ϕ L1 . ϕ L1 the k·kN will be specified by the proof of the next lemma, which roughly says that every ψ ∈ S(Rn) is a superposition of dilations of ϕ.

Lemma 16. Any ψ ∈ S(Rn) can be written as a series ∞ X (k) ψ = η ∗ ϕ2−k k=0 converging in S(Rn), where the functions η(k) ∈ S(Rn) satisfy Z 2(n+1) (k) −k(n+2) (1 + |y|) η (y) dy . 2 kψkN for a suitable seminorm k·kN depending only on n (while the implied constant de- pends also on ϕ).

12 n Proof. Let (ρk)k∈N be a (inhomogeneous) dyadic partition of unity in R , which can ∞ be obtained by taking ρ0 ∈ Cc (B2), ρ0 ≡ 1 in a neighbourhood of B1 and letting −k −(k−1) ρk := ρ0(2 ·) − ρ0(2 ·) for k > 0 (so that, for k > 0, ρk is supported in the open annulus B2k+1 \ B2k−1 ). R Since ϕ(x) dx 6= 0, we have ϕb(0) 6= 0. By continuity we can find k0 ≥ 0 such that (k) n ϕb(ξ) 6= 0 for all ξ ∈ B21−k0 . For k ≥ k0 let η ∈ S(R ) be defined by

(k) ρk−k0 ψb ηd := −k ϕb(2 ·)

(notice that the right-hand side makes sense on B2k−k0+1 and vanishes near the boundary of this ball, so it can be smoothly extended by 0 on the complement). Let (k) η := 0 for k < k0. The series ∞ ∞ X X η(k)ϕ(2−k) = ρ ψ d b k−k0 b k=0 k=k0 converges to ψb in S(Rn), so (by the continuity of F −1) we also have ∞ X (k) η ∗ ϕ2−k = ψ k=0

n R 2(n+1) in S(R ). We can find a seminorm k·kN 00 such that (1+|y|) |η| (y) dy . kηbkN 00 , so that for k ≥ k0

Z 2(n+1) (k) ρk−k0 ψb (1 + |y|) η (y) dy . . ϕ(2−k·) b N 00

Using the Leibniz rule it is easy to see that, for a suitable bigger seminorm k·kN 0 independent of ϕ,

ρk−k0 ψb −k(n+2) . 2 ψb ϕ(2−k·) N 0 b N 00 (the implied constant, however, will depend on ϕ and k , i.e. on ϕ). We can finally 0

find k·kN such that ψb . kψkN . N 0

n t Corollary 17. For any x ∈ R we have GMf(x) . Mϕf(x).

n P∞ (k) Proof. Fix ψ ∈ S(R ) such that kψkN ≤ 1. Since ψt = k=0 ϕ2−kt ∗ (η )t, ∞ X (k) |ψt ∗ f| (x) ≤ ϕ2−kt ∗ (η )t ∗ f (x) k=0 ∞ Z X −n (k) −1 ≤ |ϕ2−kt ∗ f| (x − y) t η (t y) dy k=0 ∞ Z  n+1 t X |y| −n (k) −1 ≤M f(x) 1 + t η (t y) dy. ϕ 2−kt k=0

13 But the last integral is bounded by Z  n+1 Z k(n+1) |y| −n (k) −1 k(n+1) n+1 (k) −k 2 1 + t η (t y) dy = 2 (1 + |y|) η (y) dy 2 t .

t for all k ≥ 0. So we obtain |ψt ∗ f| (x) . Mϕf(x) and the thesis follows by taking n the supremum over t > 0 and over ψ ∈ S(R ), kψkN ≤ 1.

In order to prove the implication (1) ⇒ (2), we need some technical preliminaries. Fix 0 <  < 1 and define the following modified maximal functions:

 n+1 c t Mfϕf(x) := sup |ϕt ∗ f| (y) , −1 0

c c Clearly Mfϕf converges to Mϕf pointwise from below, as  → 0, and most impor- tantly it always lies in L1(Rn): from t +  |y| ≥ t +  |y| ≥  |x| we infer  t n+1 tn+1 |ϕ ∗ f| (y) ≤ kϕ k kfk t t +  +  |y| t L∞ L1 ( +  |x|)n+1 −n n+1 −n−1 −n−1 .t kfkL1 t  (1 + |x|) −n−2 −n−1 1 n ≤ (1 + |x|) ∈ L (R ).

Lemma 18. We still have

t c GMg f . Mfϕf . Mfϕf , L1 L1 L1 the implied constants being independent of  and f.

Proof. The second inequality is proved exactly as we did for the original maximal −1 functions: again we have, whenever 0 < t <  and z ∈ Bt(y),  n+1 t c |ϕt ∗ f| (y) ≤ Mf f(z) t +  +  |y| ϕ

n and, raising this inequality to the power n+1 , averaging as z varies in Bt(y) and then n+1 raising to the power n , we get again

 n/(n+1)(n+1)/n t c Mfϕf(x) ≤ M Mfϕf (x) for any x ∈ Rn, from which the second inequality follows (using the L(n+1)/n- boundedness of the Hardy-Littlewood maximal function).

14 ∞ (k) P −k Let us turn to the first inequality. Using the decomposition ψt = k=0 ϕ2 t ∗ η t we obtain again (for any x ∈ Rn and any 0 < t < −1) ∞ Z X −n (k) −1 |ψt ∗ f| (x) ≤ |ϕ2−kt ∗ f| (x − y) t η (t y) dy, k=0 but now we estimate (using 0 < 2−kt < −1)

 n+1  −k n+1 t |y| 2 t +  +  |x − y| |ϕ2−kt ∗ f| (x − y) ≤ Mf f(x) 1 + . ϕ 2−kt t

n+1  t  Inserting this into the preceding inequality and multiplying both sides by t++|x| we arrive at  n+1 ∞ t t X |ψt ∗ f| (x) ≤ Mf f(x) Ik, t +  +  |x| ϕ k=0 where Ik denotes the following integral:

Z  n+1  −k n+1 |y| 2 t +  +  |x − y| −n (k) −1 Ik := 1 + t η (t y) dy. 2−kt t +  +  |x|

The second factor in the definition of Ik is bounded by

t +  +  |x| +  |y|n+1   |y| n+1  |y|n+1 = 1 + ≤ 1 + , t +  +  |x| t +  +  |x| t where we used the assumption  < 1, while the first factor is again bounded by n+1 k(n+1)  |y|  2 1 + t . Thus,

Z  2(n+1) k(n+1) |y| −n (k) −1 Ik ≤2 1 + t η (t y) dy t Z k(n+1) 2(n+1) (k) −k =2 (1 + |y|) η (y) dy . 2 , in view of the statement of Lemma 16. So we get

 n+1 t t |ψt ∗ f| (x) Mf f(x) t +  +  |x| . ϕ

−1 and taking the supremum over 0 < t <  we obtain the pointwise bound GMg f(x) . t Mfϕf(x), from which we infer the first inequality of the thesis.

c v Theorem 19. For any 0 <  < 1 we have Mfϕf . Mϕf 1 (the implied L1 L constant is independent of ).

15 R c Proof. We claim that it suffices to bound the integral E Mfϕf(x) dx on the ‘bad’ set n c o E := GMg f ≤ λMfϕf , for some large enough λ. Indeed, Z Z Z c −1 −1 c Mfϕf(x) dx ≤ λ GMg f(x) dx ≤ Cλ Mfϕf(x) dx, Rn\E Rn\E

c since by the preceding lemma GMg f ≤ C Mfϕf (for some C depending only L1 L1 on n and ϕ). Choosing λ := 2C we get Z Z c 1 c Mfϕf(x) dx ≤ Mfϕf(x) dx. Rn\E 2

We can now subtract the finite quantity 1 R Mc f(x) dx from both sides (this 2 Rn\E fϕ step is the reason why we needed to introduce these modified maximal functions: c the same integral with Mϕf could a priori be infinite) and obtain Z Z 1 c 1 c Mfϕf(x) dx ≤ Mfϕf(x) dx, 2 Rn\E 2 E so that Z Z Z Z c c c c Mfϕf(x) dx = Mfϕf(x) dx + Mfϕf(x) dx ≤ 2 Mfϕf(x) dx. Rn\E E E

−1 Fix now x ∈ E and let (y, t) such that 0 < t <  , y ∈ Bt(x) and

 n+1 t 1 c |ϕt ∗ f| (y) ≥ Mf f(x). t +  +  |y| 2 ϕ

1 1 0 We aim at showing that the same inequality holds, with 4 in place of 2 , for all y in a small ball Bηt(y) (0 < η < 1 will be specified later). Once this is achieved, we will have

Z !2 1 c v 1/2 0 0 Mfϕf(x) ≤ − (Mϕf) (y ) dy 4 Bηt  n Z !2 t + ηt v 1/2 0 0 ≤ − (Mϕf) (y ) dy ηt Bt+ηt(x) 2n 1 + η  2 ≤ M(Mv f)1/2 (x), η ϕ from which the thesis follows as usual (integrating over E and using the L2-boundedness of the Hardy-Littlewood maximal function).

16 n+1 0 0  t  Let g(y ) := ϕt ∗f(y ) t++|y| . The function g is locally Lipschitz and is smooth n 0 on R \{0}, so for y ∈ Bηt(y) |g(y0) − g(y)| ≤ ηt sup |∇g| (z). z∈Bηt(y)\{0} We compute

 t n+1 (n + 1)tn+1 z ∇g(z) = t−1(∇ϕ) ∗ f(z) − ϕ ∗ f(z) ·  . t t +  +  |y| t (t +  +  |y|)n+2 |z| But, writing z = x + th (with |h| < 1 + η < 2), Z x + th − u Z x − u  ϕ ∗f(z) = t−nϕ f(u) du = t−nϕ + h f(u) du = (ϕ(·+h)) ∗f(x) t t t t and similarly (∇ϕ)∗f(z) = (∇ϕ(·+h))t ∗f(x). Assuming without loss of generality 1  < 4 , we also have 1 t +  +  |z| ≥ t +  + (|x| − (1 + η)t) ≥ (t +  +  |x|) 2 t (as t − (1 + η)t ≥ 2 ). Putting everything together,  t   t n+1 |∇g| (z) t−1 |∇ϕ ∗ f| (x) + |ϕ ∗ f| (z) . t +  +  |z| t +  +  |z|  t n+1 t−1 |(∇ϕ(· + h)) ∗ f| (x) + |(ϕ(· + h)) ∗ f(x)|  . t t t +  +  |x| −1 .t GMg f(x),

∂ϕ thanks to the fact that the quantities sup|h|<2 kϕ(· + h)kN and sup|h|<2 ∂x (· + h) i N are finite (for i = 1, . . . , n). Hence,

0 0 −1 0 c |g(y ) − g(y)| ≤ ηt · C t GMg f(x) ≤ ηC λMfϕf(x)

0 1 1  (for some C depending only on n and ϕ), as x ∈ E. Choosing η := min 2 , 4C0λ we arrive at 1 1 1 g(y0) ≥ g(y) − |g(y) − g(y0)| ≥ Mfc f(x) − Mfc f(x) = Mfc f(x), 2 ϕ 4 ϕ 4 ϕ which is what we wanted to obtain.

5 Further remarks

We collect in this section the proofs of some easier parts of Theorem 2. By what 0 n n R 0 we proved in the previous section, given ϕ ∈ S(R )(R ) with ϕ (x) dx 6= 0, we have v M 0 f kGMfk kM fk ϕ L1 . L1 . ϕ L1

17 v v v v 1 and similarly M f M 0 f . So M f and M 0 f have comparable L - ϕ L1 . ϕ L1 ϕ ϕ norms. This shows that, in order to prove (5) ⇒ (1) and (7) ⇒ (1), we are free to choose ϕ at will (provided it satisfies R ϕ(x) dx 6= 0).

∞ Proof of (5) ⇒ (1). As just remarked, we can assume ϕ ∈ Cc (B1) and k∇ϕkL∞ ≤ v 0 1. The thesis follows from the trivial pointwise inequality Mϕf ≤ GM f.

Proof of (7) ⇒ (1). First of all, we claim that there exists a continuous function k ρ : [1, +∞) → R such that ρ is rapidly decreasing at infinity (i.e. supt t |ρ| (t) < +∞ for every k ≥ 0) and

Z ∞ Z +∞ ρ(t) dt = 1, tkρ(t) dt = 0 for k = 1, 2,... 1 1 (these integrals make sense by the rapid decrease assumption on ρ). An explicit example is the following: e ρ(t) := = exp e3πi/4(t − 1)1/4 . πt The rapid decrease at infinity is clear since |ρ| (t) ≤ e exp < e3πi/4(t − 1)1/4 =   πt e exp − √1 (t − 1)1/4 . Let πt 2 e g : Ω := \{t ∈ , t ≥ 1} → , g(z) := exp e3πi/4(t − 1)1/4 , C R C π where (z − 1)1/4 means the unique holomorphic function h :Ω → C such that 4 1/4 h (z) = z − 1 and lim→0+ h(t + i) = (t − 1) for every real t > 1. We remark that 3πi/4 1/4  iθ 3 5  z 7→ e (t − 1) maps Ω into re | r > 0, θ ∈ 4 π, 4 π and so   e 3πi/4 1/4 e 1 1/4 −k |g| (z) ≤ exp < e (z − 1) ≤ exp −√ |z − 1| . |z| . π π 2 Let γ be the loop (in Ω) obtained by concatenating the parametrized paths √ 2 2 iα t + i (t ∈ [1,R]), R +  e (α ∈ [α0, 2π − α0]), π 3π  R − t − i (t ∈ [0,R − 1]), 1 + e−iα (α ∈ , ), 2 2 R k−1 for arbitrary , R > 0. By Cauchy’s theorem we have γ z g(z) dz = 0 for k = 1, 2,... and Z z−1g(z) dz = 2πig(0) = 2ei exp e3πi/4eπi/4 = 2i. γ Taking the imaginary part of both identities, sending  → 0 and then R → ∞ (and noticing that the contributions of the two circular arcs are infinitesimal), we get precisely Z ∞ Z ∞ 2 tkρ(t) dt = 0 for k = 1, 2,..., 2 ρ(t) dt = 2, 1 1

18 which is the claim. We now build a Schwartz function out of the Poisson kernel: let Z ∞ ϕ(x) := ρ(t)Pt(x) dt. 1

−n −1 −n This integral converges (as |Pt| (x) = t P (t x) ≤ t P (0) ≤ P (0)) and defines a function in L1(Rn), since Z Z ∞ Z Z ∞ |ϕ| (x) dx ≤ |ρ| (t)Pt(x) dx dt = |ρ| (t) dt < +∞. Rn 1 Rn 1 Moreover, using Fubini’s theorem, ϕ(ξ) = R ∞ ρ(t)e−t|ξ| dt. It is easy to show induc- b 1 tively that, for ξ 6= 0 and any multiindex α 6= 0,

|α| Z ∞ ∂ ϕb(ξ) −1 −t|ξ| α = ρ(t) · t Qα(t, ξ, |ξ| )e dt ∂ξ 1

n for a suitable polynomial Qα. In particular, ϕb is smooth on R \{0} and all its are rapidly decreasing at infinity. Moreover, ϕb is clearly continuous. Given any α 6= 0, we write

X (−s)k e−s = + R (s) k! K k

−K and notice that s |RK | (s) is bounded for s ∈ R \{0} close to the origin, while it K is infinitesimal as |s| → ∞; thus |RK | (s) . |s| . This implies

|α| Z ∞ k ! ∂ ϕ(ξ) −1 X (−t |ξ|) b = ρ(t) · t Q (t, ξ, |ξ| ) + R (t |ξ|) dt. ∂ξα α k! K 1 k

0 Calling d and d the degrees of Qα with respect to its first and last argument, respectively, we obtain that for every K > d0 Z ∞ Z ∞ −1 1+d+K ρ(t) · t Qα(t, ξ, |ξ| ) |RK | (t |ξ|) dt . t |ξ| dt . |ξ| 1 1 (whenever 0 < |ξ| ≤ 1), while

Z ∞ k −1 X (−t |ξ|) ρ(t) · t Q (t, ξ, |ξ| ) dt = 0 α k! 1 k

19 ∞ Proof of (6) ⇒ (5). Let ψ ∈ Cc (B1(0)) with k∇ψkL∞ ≤ 1. Given an ∞-atom supported in Br(x0), we have

|ψt ∗ a| (x) ≤ kψtkL1 kakL∞ = kψkL1 kakL∞ . kakL∞ n for any x ∈ B2r(x0). Fix now x ∈ R \ B2r(x0) and notice that ψt ∗ a(x) = 0 if t < |x − x0| − r (since in this case ψt(x − ·) and a are supported in the disjoint balls Bt(x) and Br(x0)). Assume instead that t ≥ |x − x0| − r: in this case we get |x−x0| t ≥ 2 , so Z |ψt ∗ a| (x) ≤ |ψt(x − y) − ψt(x − x0)| |a| (y) dy

≤r k∇ψtkL∞ kakL1 −n−1 r .rt . n+1 |x − x0| −n−1 −1 (as ∇ψt(x) = t ∇ψ(t x)). Thus, Z Z 0 r kGM fkL1 . kakL∞ dx + n+1 dx n B2r(x0) R \B2r(x0) |x − x0| Z ∞ ρn−1 dρ . kakL∞ |Br(x0)| + r n+1 . 1. 2r ρ P Hence, if f = k λkak is an atomic decomposition,

X X kGM0fk ≤ λ GM0a λ . L1 k k . k k L1 k

Proof of (6) ⇒ (7). The proof of Proposition 6 can be repeated verbatim, with ϕ replaced by P , to show that v kMP akL1 . 1 P for any ∞-atom a. Hence, if f = k λkak is an atomic decomposition,

X X kMv fk ≤ λ Mv a λ . P L1 k P k . k k L1 k

Proof of (6) ⇒ (8). It suffices to notice that the proof of Theorem 11 used only the ξj atomic decomposition of f. So, choosing m(ξ) := −i |ξ| , we deduce X kRjfkL1 . inf λk (the infimum ranging over all the possible atomic decompositions). Moreover, for any decomposition X X kfkL1 ≤ λk kakkL1 ≤ λk. k k Pn P Thus, kfkL1 + j=1 kRjfkL1 . inf k λk.

20 6 Characterization with the Riesz transforms

We now show the implication (8) ⇒ (7) among the equivalent definitions of H1(Rn). The proof will implicitly show the corresponding inequality on the norms, namely

v kMP fkL1 . kfkL1 + kR1fkL1 + ··· + kRnfkL1 .

Assume that f and all its Riesz transforms are in L1(Rn). So far we have tacitly allowed any function to be either real or complex valued, but now it is convenient to assume f real valued (without loss of generality, as Rj maps real functions to real distributions). The functions

uj(x, t) := Pt ∗ Rjf(x) for 1 ≤ j ≤ n, un+1(x, t) := Pt ∗ f(x) form a system of conjugate harmonic functions on Hn+1 := {(x, t) ∈ Rn+1 : t > 0}, i.e. they satisfy the following system of generalized Cauchy-Riemann equations:

n+1 X ∂uj ∂uj ∂uk = 0, = for any 1 ≤ j, k ≤ n + 1, ∂x ∂x ∂x j=1 j k j where xn+1 is an alias for the auxiliary variable t. This can be checked using the formulas

−n/2 −t|ξ| −n/2 ξj −t|ξ| F(Pt ∗ f)(ξ) = (2π) e fb(ξ), F(Pt ∗ Rjf)(ξ) = −(2π) i e fb(ξ). |ξ|

Clearly it suffices to prove that

1 n sup |u| (·, t) ∈ L (R ), t>0 where u := (u1, . . . , un+1). We could bound |u| (x, t) by the Hardy-Littlewood max- imal function of (f, R1f, . . . , Rnf) at x, but this would be useless (as the Hardy- Littlewood maximal function satisfies only a weak (1,1) bound). Rather, we aim at q 1/q n showing that |u| (x, t) . Mg(x) for some q < 1 and some g ∈ L (R ) with

1/q kgkL1/q . kfkL1 + kR1fkL1 + ··· + kRnfkL1 , from which the thesis will follow since

1/q 1/q sup |u| (·, t) . kMgkL1/q . kgkL1/q . t>0 L1

n−1 Pick now n < q < 1 (so that in particular q > 0). From Lemma 20 below, we know that (|u|2 + 2)q/2 is subharmonic (for any  > 0). Thus it satisfies the following n+1 0 version of the maximum principle: for any Ω b H and any h ∈ C (Ω) harmonic in Ω, the implication

(|u|2 + 2)q/2 ≤ h on ∂Ω ⇒ (|u|2 + 2)q/2 ≤ h on Ω

21 holds. Sending  → 0, it is easy to check that |u|q satisfies the same property. Lemma 21 below tells us that this property applies also with the q n+1 h(x, t) := Pt−δ ∗ |u| (x, δ) on the unbounded domain {(x, t): t > δ} ⊆ H , for any δ > 0. Notice that

q 1/q sup k|u| (·, δ)kL1/q = sup ku(·, δ)kL1 ≤ kfkL1 + kR1fkL1 + ··· + kRnfkL1 . δ>0 δ>0

Since any closed ball in L1/q is weakly sequentially compact, we can find a sequence 1/q n 1/q δk → 0 and a function g ∈ L (R ) (whose L -norm satisfies the same bound) q (1/q)0 such that |u| (·, δk) * g. Since Pt−δk → Pt in L , we deduce that

q q |u| (x, t) ≤ lim (Pt−δ ∗ |u| (·, δk))(x) = Pt ∗ g(x). k→∞ k

Finally, by Remark 5, we have Pt ∗ g ≤ Mg, which was our goal. It remains to prove the two lemmas.

n−1 2 2 q/2 Lemma 20. For any q ≥ n the function (|u| +  ) is subharmonic, i.e.

∆ (|u|2 + 2)q/2 ≥ 0.

∂ Proof. We will use the shorthand notation ∂j := . We compute ∂xj

2 2 q/2 2 2 (q/2)−1 ∂j(|u| +  ) = q(|u| +  ) u · ∂ju,

X 2 2 2 q/2 X 2 2 (q/2)−2 2 X 2 2 (q/2)−1 2 ∂jj(|u| + ) = q(q−2)(|u| + ) (u·∂ju) + q(|u| + ) |∂ju| j j j n−1 (using ∆u = 0). The thesis follows immediately if q ≥ 2, so we can assume n ≤ n+1 q < 2, i.e. 0 < 2 − q ≤ n . It suffices to prove that

n + 1 X 2 X 2 (u · ∂ u)2 ≤ |u| |∂ u| . n j j j j

This inequality is a consequence of the generalized Cauchy-Riemann equations: in- deed, the matrix A := (∂juk(x))jk is symmetric, so (by the ) we can find P ∈ O(n + 1) and a diagonal matrix D such that A = P tDP . The coefficients on the diagonal of D are the eigenvalues λ1, . . . , λn+1 of A. We remark that X λj = tr(D) = tr(A) = 0. j

We pick j0 such that |λj0 | = maxj |λj|. By Cauchy-Schwarz we have

2 2 X X 2 (n + 1) |λ | = n |λ | + λ ≤ n |λ | , j0 j0 j j j6=j0 j

22 ! u1(x) so, letting v := P . , we can estimate . ! 2 n + 1 X 2 n + 1 X u1(x) n + 1 2 n + 1 2 2 |u · ∂ju| = A . = |Dv| ≤ |λj | |v| n n . n n 0 j j . X 2 2 ≤ |λj| |u| (x). j We finally observe that

X 2 t t t t t X 2 |λj| = tr(D D) = tr(PA P P AP ) = tr(A A) = |∂ju| . j j

q q Lemma 21. For t > δ we have |u| (x, t) ≤ (Pt−δ ∗ |u| (·, δ))(x).

Proof. Let us first prove that, for every η > 0, there exists an arbitrarily large radius R > 0 such that |u| ≤ η on the set {(x, t): t ≥ δ, |(x, t)| ≥ R}. From the mean-value property of harmonic functions, for any (x, t)) in this set we have 1 Z |u| (x, t) ≤ |u| (y, s) dy ds. Bt/2(x, t) Bt/2(x,t) If |x| ≤ t then t ≥ √R and we can estimate 2 Z 1 −n −n |u| (x, t) ≤ |u| (y, s) dy ds . At . AR B (x, t) n t 3 t/2 R ×( 2 , 2 t)

(where A := sups>0 ku(·, s)kL1 < +∞), which becomes small at will taking R large enough. Otherwise, if |x| > t, then |x| ≥ √R and any point (y, s) ∈ B (x, t) satisfies 2 t/2 |x| |y| > 2 , so Z 3t/2 Z Z 3t/2 Z −n−1 −n−1 |u| (x, t) . t |u| (y, s) dy ds . s √ |u| (y, s) dy ds. t/2 |y|>|x|/2 t/2 |y|>R/ 8 But the latter quantity can be uniformly estimated by Z ∞ Z −n−1 s √ |u| (y, s) dy ds, δ/2 |y|>R/ 8 which can be made arbitrarily small taking R large enough, thanks to the dominated convergence theorem (since the inner integral is bounded by A and tends to 0 as R → +∞). q Now h(x, t) := (Pt−δ ∗ |u| (·, δ))(x) is harmonic on {(x, t): t > δ} and extends con- q tinuously to the boundary Rn ×{δ}, where it coincides with |u| . So we have proved that q q q |u| (x, t) ≤ (Pt−δ ∗ |u| (·, δ))(x) + η on the boundary of SR := {(x, t): t > δ, |(x, t)| < R} for any R large enough. We deduce that this inequality is also true on SR itself. Thus, letting R → +∞, we infer that it holds on {(x, t): t > δ}. The thesis follows as we let η → 0.

23 7 Existence of the atomic decomposition

In this section we show that any function f ∈ L1(Rn) with GMf ∈ L1(Rn) admits an atomic decomposition X f = λkak k=0 P with λk ≥ 0, (ak) a collection of ∞-atoms and k λk . kGMfkL1 , thereby proving the implication (4) ⇒ (6) and the bound on the corresponding norms. [··· work in progress ··· ]

8 Littlewood-Paley characterization

1 n ˙ 0 n In this section we are going to prove that H (R ) = F1,2(R ), in the sense speci- fied by Theorem 23, denoting by H1(Rn) the space of functions satisfying (any of) the definitions (1) − (8), whose equivalence has been established in the previous sections. We fix a function ψ ∈ C∞(B \B ) such that P ψ(2−jξ) = 1 for any ξ ∈ n\{0}. c 2 1/2 j∈Z R ∞ Recall that such a ψ can be produced by taking any φ ∈ Cc (B2) such that φ ≡ 1 in a neighbourhood of B1 and letting ψ := φ − φ(2·). n R n We let ϕ ∈ S( ) be any function such that n ϕ dx 6= 0 and supp (ϕ) ⊆ B2. R R For instance we can take ϕ := F −1(φ) for any φ as above (using R ϕ dxn = Rn (2π)−n/2φ(0) 6= 0).

Lemma 22. For any f ∈ S0(Rn) and any r ∈ (0, ∞) we have r 1/r sup |ϕt ∗ Pjf| (x) ≤ C(n, r)M |Pjf| (x) . t>0

0 n Proof. Recall that, whenever v ∈ S (R ) has its supported in B1, we have the inequality

|v(x − z)| r 1/r sup n/r . (M |v| ) (x). z (1 + |z|)

−1 More generally, if ub is supported in Bs, letting v := u(s ·) we obtain

|u(x − z)| |v(sx − z)| r 1/r r 1/r sup n/r = sup n/r . (M |v| ) (sx) = (M |u| ) (x). z (1 + s |z|) z (1 + |z|) If 2 ≤ 2j−1 (i.e. if t ≥ 22−j) we have ϕ ∗ P f ≡ 0, since the supports of ϕ and t t j bt ψ(2−j·) are disjoint in this case. Assume now that t ≤ 22−j: in this case ψ(2−j·) is n −N supported in B8/t, so choosing any N > r + n and estimating |ϕ(z)| . (1 + |z|) we get Z Z −n −n |Pjf| (x − z) |Pjf| (x − z) t |ϕt ∗ Pjf| (x) . t −1 N dz ≤ sup −1 n/r −1 N−n/r dz. (1 + t |z|) z (1 + t |z|) (1 + t |z|)

24 The last integral is a finite constant independent of t, while

|P f| (x − z) |P f| (x − z) sup j sup j (M |P f|r)1/r(x). −1 n/r . 8 n/r . j z (1 + t |z|) z (1 + t |z|)

Before stating and proving the next theorem, we introduce the vector-valued Hardy space H1(Rn, `2): it is the subspace of   Z !1/2 1 n 2  1 n X 2  L (R , ` ) := (fj)j∈Z ⊆ L (R ): |fj| < +∞  j  made of elements (fj) satisfying one of the equivalent definitions (1) − (7) in vec- torized form. For instance, (1) amounts to ask that

1 n sup kϕt ∗ (fj)k`2 = sup k(ϕt ∗ fj)k`2 ∈ L (R ). t>0 t>0 Their equivalence comes from the fact that the proofs for the scalar case can be repeated verbatim for the vectorial case (we exclude definition (8) since its equiv- alence with the other definitions uses real numbers in an essential way, due to the appeal to the spectral theorem).

Theorem 23. For any f ∈ H1(Rn) we have

k(Pjf)j∈ZkL1(`2) . kfkH1 .

0 n Conversely, if for some f ∈ S (R ) we have k(Pjf)j∈ZkL1(`2) < ∞, then there exists a unique polynomial Q such that f − Q ∈ H1(Rn); moreover kf − Qk k(P f) k . H1(Rn) . j j∈Z L1(`2)

Proof. The first statement follows immediately from the H1 → L1(`2) version of −1 −j Theorem 10, applied with Kj := F (η(2 ·)), with assumptions (1) and (3) re- placed by the validity of the L2 → L2(`2) bound (which holds as a consequence of Plancherel’s theorem). This variant of Theorem 10 is simply obtained by vectorizing the same proof (and in this case there is no need of truncating the kernel). Recall that this `2-valued kernel satisfies the H¨ormandercondition Z n k(Kj(x − y) − Kj(x))k`2 dx . 1. |x|>2|y|

·  We now turn to the converse. Pick η := ψ(2·) + ψ + ψ 2 and notice that η ≡ 1 near the support of ψ. Let

˜ 0 n 0 n ˜ −1 −j Pj : S (R ) → S (R ), Pj(g) := F (η(2 ·)Fg)

25 ˜ 1 2 1 and remark that PjPj = Pj. Applying the H (` ) → L version of Theorem 10 with −1 −j f := (Pjf)j∈Z and Kj := F (η(2 ·)), we can estimate

N N N X X ˜ X Pjf = PjPjf . Pjf . j=−N L1 j=−N L1 j=−N H1(`2)

1 PN 2 We can similarly estimate the L -norm of Rk j=−N Pjf, using the ` -valued kernel   −1 ξk −j Kj := F −i |ξ| η(2 ·) . Thus,

N N N !1/2 X X X 2 Pjf . Pjf = sup |ϕt ∗ Pjf| . t>0 j=−N j=−N j=−N H1 H1(`2) L1 Using Lemma 22 with any 0 < r < 1, as well as the Hardy-Littlewood maximal estimate for L1/r(`2/r), the last quantity is bounded up to constants by

N !1/2 X r r 1/r r 1/r (M |P f| )2/r = k(M |P f| )k k(|P f| )k j j 1/r 2/r . j 1/r 2/r L `N L `N j=−N L1

= k(Pjf)kL1(`2) ,

p where `N denotes the truncated space of a = (a−N , . . . , aN ) with the 1/p PN p norm kak p := |aj| . The same argument shows that the partial sums `N j=−N PN 1 n j=−N Pjf form a Cauchy sequence in H (R ) and thus, by Proposition 8, converge to some g ∈ H1(Rn).

PN  0 n But F j=−N Pjf → fb in D (R \{0}), so the tempered distribution fb− gb is supported in {0}. This means that

−1   Q := f − g = F fb− gb is a polynomial. So f −Q = g ∈ H1(Rn) and, letting N → ∞ in the above estimate, we also have

kf − QkH1 = kgkH1 . k(Pjf)kL1(`2) .

References

[1] D.L. Burkholder, R.F. Gundy, and M. L. Silverstein. “A maximal func- tion characterization of the class Hp”. In: Transactions of the American Math- ematical Society 157 (1971), pp. 137–153. [2] P.L. Duren. Theory of Hp spaces. Courier Corporation, 2000.

26 [3]C. Fefferman and E.M. Stein.“Hp spaces of several variables”. In: Acta mathematica 129.1 (1972), pp. 137–193. [4]L. Grafakos. Modern Fourier Analysis. Vol. 250. Springer-Verlag New York, 2009. doi: 10.1007/978-0-387-09434-2. [5] G.H. Hardy. “The mean value of the modulus of an analytic function”. In: Proceedings of the London Mathematical Society 2.1 (1915), pp. 269–277. [6]F. Riesz.“Uber¨ die Randwerte einer analytischen Funktion”. In: Mathema- tische Zeitschrift 18.1 (1923), pp. 87–95. [7]W. Rudin. Real and complex analysis. McGraw-Hill, 1987. [8] E.M. Stein. Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals. Vol. 43. Princeton University Press, 1993. [9] E.M. Stein and G. Weiss. “On the theory of harmonic functions of several variables”. In: Acta Mathematica 103.1-2 (1960), pp. 25–62. [10]H. Triebel. Theory of function spaces. Birkh¨auserBasel, 1983. doi: 10 . 1007/978-3-0346-0416-1.

27