arXiv:2011.10651v1 [math.CA] 20 Nov 2020 a eue odefine to used be can operator nwrti usin ie naiorpcHryspace Hardy anisotropic an stan Given operator differential question: a this from answer discre characterization is obvious characterizations these no in employed has dilation the characteriza of molecular scale and the atomic, maximal, in Gauss grounded contributions. formulation 17], other many [16, among spaces [22] Hardy Hardy weighted exponent of variable and settings estimat the off-diagonal e in form reinforced divergence generalized and to [14], associated estimates [1], Davies-Gaffney estimates off-diagonal an fying iainmatrix dilation etn nwihtedlto cl scniuu.Antrlstigtu th setting where natural 7], A [6, Torchinsky continuous. dilations Calder´on and is of of scale group dilation Hardy the no. which parabolic is in question setting second a the to answer the PDE, this of nature omltoso lsia ad pc.T aejs e refere v few the a operator just in an name limitations To with as associated space. well Hardy spaces as grand classical challenges ope a the differential its of of various formulations introducing employment with setting versatile associated each spaces the with Hardy to respective of led equations, study the [12] Dirichlet and Stein heat and Fefferman the to solutions damental oso qain hs soitdmxmloeaosaein are operators maximal associated whose equations Poisson tt nvriy n ol iet hn irePra o discussio for Portal Pierre thank to like would and supported University, was author State second The discussions. helpful F¨uhr for mns l hsatvt,teaiorpcHrysae[]apast appears [2] space Hardy anisotropic the activity, this all Amongst faPEde xs ocaatrz h nstoi ad pc,t space, Hardy anisotropic the characterize to exist does PDE a If lsial,Hryspaces Hardy Classically, e od n phrases. and words Key 2010 Date h rtato a atal upre yNFgatDMS-195639 grant NSF by supported partially was author first The D HRCEIAINO NSTOI AD SPACES HARDY ANISOTROPIC OF CHARACTERIZATION PDE A oebr2,2020. 24, November : ahmtc ujc Classification. Subject Mathematics dniyn twt aaoi ad pc soitdwt eea c general a with associated space Hardy allows parabolic a with it identifying eas rvd lsicto fdltoscrepnigt equivalen transformations. to linear corresponding to dilations respect of with classification spaces a provide also We Abstract. L sw ilse h nwrt h rtqeto sys u ie the given but yes, is question first the to answer the see, will we As ? H A p A ob enduigaprblcdffrnileuto fCald´eron and of equation differential parabolic a using defined be to steeaprildffrnileuto PE hs fundamental whose (PDE) equation differential partial a there is , eoti ieeta hrceiainfrteaiorpcHardy anisotropic the for characterization differential a obtain We { A t H } nstoi ad pcs aaoi ad pcs otnosgr continuous spaces, Hardy parabolic spaces, Hardy anisotropic t> ACNBWI N IA AILWANG DANIEL LI-AN AND BOWNIK MARCIN A p 1. lentvl,cnw hrceiei as it characterize we can Alternatively, ? 0 satisfying , nrdcinadMi Results Main and Introduction H p a ecaatrzda onayvle oteha or heat the to values boundary as characterized be can L 23 3K0 22,42B35). 42B25, (35K10, 42B30 aifigapitieha enlbud 1] satis- [11], bounds kernel heat pointwise a satisfying A s A t = 1 A st rvdsti eddsrcue swell as structure, needed this provides , ytedprmna ud rmSmHouston Sam from funds departmental the by H nti oi tIMTi 2013. in ICMAT at topic this on n A p soitdwt nanisotropic an with associated .H ol iet hn Harmut thank to like would He 5. L p n hs enl r fun- are kernels these and , pit u betv sto is objective Our dpoint. c ok,w aeHardy have we works, nce niuu ru.This group. ontinuous aoscniu unabated, continue rators e h nstoi setting anisotropic the te, s[] hs r further are These [5]. es lpi prtr 1] the [15], operators lliptic y vna h okof work the as Even ly. H nstoi Hardy anisotropic t ea ule.Wt its With outlier. an be o in,adtefc that fact the and tions, a ad pcs[18], spaces Hardy ian e ems rtfind first must we hen L s facontinuous a of use e p o oedifferential some for aia function, maximal n u ob the be to out rns space rosequivalent arious Torchinsky. uso dilations. of oups H A p parabolic by solution as a PDE we can immediately use. Strictly speaking, the original form of this group is not flexible enough to accommodate the geometry of an anisotropic dilation, where the natural geometric object are ellipsoids of changing eccentricities. Our first task, in light of this, is to consider a wider class of continuous groups that can capture the essential information from a dilation matrix in the anisotropic setting. This is the content of Theorem 2.4. Next, we will show that every anisotropic dilation matrix A is associated with a unique continuous group that represents the same geometric information. This is the content of Theorem 3.1. Lastly, we see that the Hardy spaces associated with both types of dilations are precisely the same space. This is the content of Theorem 3.8. p Theorem 3.8 also answers the question we posed. Once the anisotropic Hardy space HA is identified with the parabolic Hardy space Hp , we have a PDE characterization of Hp . Let At A { } Dtf(x) = f(Atx) be the dilation operator induced by continuous group. Then we consider the initial value problem that seeks to find u : Rn+1 C so that + → ∂u 1 1 n+1 n = (D− ∆D )u (x, t) R = R R , (1.1) ∂t t t t ∈ + × + u(x, 0) = f(x) x Rn.  ∈ p ˜ A tempered distribution f is in HA exactly when the solution u(x, t) = f Φt(x) satisfies the regularity condition that ∗ p sup (Φ˜ t f)(x) L . ρ˜(x y)

j b if x Bj+1 Bj ρA(x)= ∈ \ (0 if x =0.

ω By setting ω to be the smallest integer so that 2B0 A B0 = Bω, ρA is a quasi-norm with the triangle inequality constant c = bω. While any two⊂ quasi-norms associated with A will give the same anisotropic structure [2, Lemma 2.4], the step norm will be our ‘canonical’ norm, denoted by ρA. A general quasi-norm (without being associated to a dilation) is a mapping ρ : Rn [0, ) satisfying ρ(x) = 0 exactly when x = 0 Rn, and for some c > 0, satisfies the inequality→ ∞ n∈ n ρ(x + y) c(ρ(x)+ ρ(y)) for all x, y R . Then if ρ1 and ρ2 be two quasi-norms on R , we say they≤ are equivalent if there exists∈C > 0 such that for all x Rd, ∈ 1 ρ (x) ρ (x) Cρ(x). C 1 ≤ 2 ≤

Definition 2.1. We say two dilations (A1, ρ1) and (A2, ρ2) are equivalent if their associated quasi-norms ρ1 and ρ2 are equivalent.

We are now in a position to define Hardy spaces adapted to the geometry of dilations. We denote as the Schwartz class, and ′ the space of tempered distributions. For a dilation S S k k A, ϕ , and k Z, we denote the anisotropic dilation by ϕk(x) = b− ϕ(A− x). We have four∈ S maximal∈ functions, corresponding to their classical counterparts, any one of which p can be used to define HA. These are radial and non-tangential maximal functions and the 3 corresponding grand maximal functions, defined for f ′, respectively, ∈ S 0 Mϕf(x) = sup f ϕk(x) , k Z | ∗ | ∈ Mϕf(x) = sup sup f ϕk(y) , k Z ρ(x y)

M N f(x) = sup Mϕf(x). S ϕ ∈SN Here, , N N, denotes the set of all ϕ such that SN ∈ ∈ S α N ϕ N = sup sup ∂ ϕ(x) ρA(x) 1. k kS x Rn α N | | ≤ ∈ | |≤ Theorem 2.2. [2, Proposition 3.10 and Theorem 7.1] Let A be a dilation, p (0, ), ϕ ∈ ∞ ∈ S be such that ϕ = 0, and N N be sufficiently large. If f ′, then the following are equivalent: 6 ∈ ∈ S R 0 p p 0 p p Mϕf L , Mϕf L , M f L , M N f L . ∈ ∈ SN ∈ S ∈ In this case, we say f Hp . ∈ A We will also use the following fact.

Theorem 2.3. [2, Theorem 10.5] Let A1 and A2 be two dilations. Then, A1 and A2 are p p equivalent if and only the corresponding anisotropic Hardy spaces coincide HA1 = HA2 for some 0

0 that would characterize HA. That is, ∂u { } T u u + t Ttf solves the Cauchy problem ∂t = Lu and u(x, 0) = f(x), with Lu = limt 0 t− . To overcome this obstacle, we will instead relate the anisotropic dilation matrix with→ continuous groups At t>0. While this setting has been studied by a number of authors, our approach is informed{ } by Calder´on and Torchinsky [6, 7] and Stein and Wainger [20].

2.2. Parabolic Setting. A continuous group At t>0 is a collection of linear operators n n { } At : R R such that for all s,t > 0, it satisfies the algebraic identity AtAs = Ast and the → n mapping t Atx is continuous for all x R . This guarantees the existence of a unique n n matrix7→P , which we call the generator∈ of A , such that × { t} At = exp(P ln t), t> 0. Conversely, given any real-valued n n matrix P , we can construct a continuous group given by the above exponential formula. × In [6], the continuous group considered carries the requirement that there are c1,c2 1 such that such that for all t> 1 and x Rn, ≥ ∈ (2.1) tc1 x A x tc2 x , | |≤| t | ≤ | | which forces the Euclidean ball to be invariant under At for all t > 1. However, this is not sufficient to capture the setting of an anisotropic dilation, where such invariance might fail for t > 1 close to 1. Instead, the natural geometric objects are ellipsoids of changing 4 eccentricities whose major axes do not stay fixed. Because of this, we will work with a more general collection of semigroups A , characterized by the following theorem. { t} Theorem 2.4. Given a continuous group At t>0 with generator P , the following are equiv- alent. { } n (a) For all x R , lim Atx =0. ∈ t 0 | | (b) All eigenvalues λ→of P satisfy Re(λ) > 0. (c) There exists t > 1 such that for all x Rn 0 , A x > x . 0 ∈ \{ } | t0 | | | Definition 2.5. A continuous group At t>0 that satisfies the conditions from Theorem 2.4 is called an expansive continuous group.{ } Remark 2.6. This characterization has appeared numerous times in literature, under var- ious names. In Stein and Wainger [20, Part II], these groups were simply called dilations, and it was stated that condition (a) implied (b), while in [21, page 126], it was stated that (b) implied (a). While the proof of either direction only requires elementary arguments in linear algebra, we provide them here for reference. While we will not need condition (c) in the rest of the paper, we include it for completeness. Proof of Theorem 2.4. We first establish the implication (a) (b). Let λ be an eigenvalue ⇒ of P . We write λ = λr + iλi, where λr and λi R are the real and complex parts of n ∈ λ, respectively. Let xλ C be the associated eigenvector, also potentially with complex entries, so we write ∈ xλ = xr + ixi, n where xr and xi are both vectors in R . Using the relationship At = exp(P ln t) for t > 0, we have A x = exp(P ln t)x = exp(λ ln t)x = tλx = tλr tiλi x = tλr x . | t λ| | λ| | λ| | λ| | λ| | λ| iλi We used the fact that for any t> 0, t = 1. Since Atxλ 0 as t 0, we have λr > 0. Next, we establish (b) (c). Suppose| | that all eigenvalues→ of P have→ positive real parts. ⇒ λ Define the matrix B = A1/2 = exp( P ln 2). If λ is an eigenvalue of P , then 2− is an eigenvalue of B. Hence, all eigenvalues−λ of B satisfy λ < 1. By the spectral radius formula k 1/k | | k k limk B < 1. Hence, there exists k N such that B < 1. Let t0 = 2 . Since →∞ || k || ∈ || || A1/t0 = B , we have x A A x = Bk A x for x Rn. | |≤|| 1/t0 ||| t0 | || ||| t0 | ∈ Hence, (c) follows. Finally, we establish (c) = (a). Suppose that (c) holds for some t0. We will first show that there exists c > 1 so that⇒ for all x Rn, 0 ∈ (2.2) A x c x . | t0 | ≥ 0| | n 1 n Indeed, the function x A x defined on the unit sphere S − R achieves a minimum 7→ | t0 | ⊂ value c0. By(c) we deduce that c0 > 1. Hence, by the homogeneity we have (2.2). Applying (2.2) recursively we have for any k N, ∈ 1 k A −k x (c )− A −k+1 x ... (c )− x . | (t0) | ≤ 0 | (t0) | ≤ ≤ 0 | | Letting k yields (a).  →∞ 5 2.3. Homogeneous quasi-norms. In analogy to discrete anisotropic setting we adapt the following definition of a homogeneous quasi-norm. n Definition 2.7. Let At t>0 be an expansive continuous group. We sayρ ˜ : R [0, ) is a homogeneous quasi-norm{ } with respect to A if there exists c> 0 such that → ∞ { t} ρ˜(x)=0 x = 0, ⇐⇒ n ρ˜(Atx)= tρ˜(x) for all x R , t> 0, ρ˜(x + y) c(˜ρ(x)+˜ρ(y)) for all x,∈ y Rn. ≤ ∈ The following construction yields a particularly useful example of a homogeneous quasi- norm. Let P be a generator of an expansive group At t>0. Then, there exists a positive definite symmetric matrix B that can be defined algebraically{ } [6, Lemma 1.2] by

(2.3) BP + P ∗B = I, or by the integral [20, Proposition 1-7]

∞ B = exp( tP ∗) exp( tP )dt. − − Z0 The matrix B satisfies the identity d 1 1 BA x, A x = (BP + P ∗B)A x, A x = A x, A x > 0. dth t t i t h t t i t h t t i n Hence, for any x R , the quantity BAtx, Atx is strictly increasing with respect to t. This allows the following∈ construction ofh a homogeneousi norm [20, Definition 1-8]. Because of this property, we call B the norm-inducing matrix.

Proposition 2.8 ([20, Proposition 1-9]). Let At t>0 be an expansive continuous group with generator P . Let B be a positive definite symmetric{ } matrix satisfying (2.3). For x = 0 6 we define ρ˜(x) = t to be the unique t > 0 such that BA −1 x, A −1 x = 1. For x = 0 we h t t i set ρ˜(x)=0. Then, ρ˜ is a homogeneous quasi-norm with respect to At , which is C∞ on Rn 0 . { } \{ } With respect to the Fourier transform, let A∗ be the adjoint of the continuous group, { t }t>0 which is itself another continuous group with P ∗ as its generator and with B∗ as its norm- inducing matrix. We denoteρ ˜ to be the associated quasi-norm. If At are diagonal matrices, thenρ ˜ =ρ ˜. ∗ With∗ this class of continuous groups, we extend the definition of a parabolic Hardy space from [6, 7] verbatim, and denote such a Hardy space Hp , emphasizing we are using the At { } whole group At to define the Hardy space. Fix such a group, and with it, generator P , trace γ = tr({P ),} and quasi-normρ ˜. Then the parabolic non-tangential maximal operator, associated with ϕ , is given by ∈ S M˜ ϕf(x) = sup sup f ϕ˜t(y) t>0 ρ˜(x y) 0, but for brevity we will only use the aperture a = 1. 6 Definition 2.9. [7, Definition 1.1] Let At t>0 be an expansive continuous group. If f ′ and ϕ with ϕ = 0, then we say{f } Hp if the non-tangential maximal operator∈ S At ∈ S 6 ∈ { } ˜ p p ˜ Mϕf L . In this case, we set the norm f H = Mϕf Lp . ∈ R k k {At} k k The following theorem was shown in [7, Theorem 1.2] under the assumption that a con- tinuous group of dilations At t>0 satisfies (2.1). However, as we will see later it also holds for all expansive groups. { } Theorem 2.10. Every choice of ϕ , with ϕ =0 will result in the same space Hp . At ∈ S 6 { } 3. Connection Between AnisotropicR and Parabolic Setting Having established the anisotropic and parabolic settings, the following result establishes their close relationship. Theorem 3.1 is inspired by the work of Cheshmavar and F¨uhr [8] on classification of anisotropic Besov spaces. Theorem 3.1. Suppose A is an anisotropic dilation matrix. That is, all eigenvalues λ of A satisfy λ > 1. Then there exists a unique continuous group of dilations At = exp(P ln t) | ,| such that: { }t>0 (i) Its generator P has all positive eigenvalues and tr(P )=1, and (ii) A is equivalent to At for all t> 1.

More precisely, if A′ = exp(P ′ ln t) is another one-parameter group of dilations with a { t }t>0 generator P ′ satisfying (i) and A is equivalent to At′ for some t > 1, then P = P ′. In this case, we say A is the continuous group associated with the anisotropic dilation A. { t}t>0 Remark 3.2. Due to our choice of the generator P , we have tr(P ) = 1, so det At = t, and the dilation of a function f, with respect to the continuous group, takes the form ˜ 1 1 ft(x)= t− f(At− x). For the rest of the paper, we will always take our generator P to have trace 1. In the proof of Theorem 3.1 we need to use the following three lemmas from [2] and [8] about equivalence of expansive dilations, see Definition 2.1. Lemma 3.3 ([2, Lemma 10.2]). Let A and B be two expansive dilations. Then, A and B are equivalent if and only if

k ǫk ln det A sup A B−⌊ ⌋ < , where ǫ = | | . k Z || || ∞ ln det B ∈ | | Lemma 3.4 ([8, Lemma 7.6]). Let A and B be expansive matrices of the form A = exp(tX) and B = exp(sX) for some matrix X and s,t > 0. Then, A and B are equivalent. Lemma 3.5 ([8, Theorem 7.9(a)]). Let A and B be expansive matrices having only positive eigenvalues and satisfying det A = det B. Then, A and B are equivalent if and only if A = B. In addition, we will need the following strengthening of a lemma due to Cheshmavar and F¨uhr [8, Lemma 7.7]. Lemma 3.6. Let A be an expansive matrix. Then, there exists an expansive matrix B such that 7 (i) A is equivalent to B, (ii) B has all positive eigenvalues, (iii) det B = det A , and (iv) for all r>| 1 and| m =1, 2,... we have (3.1) dim ker(A λI)m = dim ker(B rI)m. − − λ =r |X| More precisely, there is a one-to-one correspondence between blocks in a complex Jordan normal form of A and blocks in a Jordan normal form of B such that each Jordan block of A for an eigenvalue λ C corresponds to a Jordan block of B of the same size for an eigenvalue λ . ∈ | | 1 Proof. There exists an invertible matrix S GL(n, R) such that S− AS has real Jordan 1 ∈ normal form. That is, S− AS is a block diagonal matrix consisting of Jordan blocks corre- sponding to either real or complex conjugate eigenvalues of A. We shall define the matrix B 1 1 such that S− BS is a block diagonal matrix where each Jordan block of S− AS is replaced by a certain matrix as follows. If λ C R is a complex eigenvalue of A, then the corresponding real Jordan block is 2k 2k∈matrix\ of the form × Mλ I2 Mλ I2 Re λ Im λ 1 0 (3.2) J =  . .  where Mλ = , I2 = M1 = . .. .. Im λ Re λ 0 1 −     M   λ  That is, J is obtained from two complex Jordan blocks of size k for conjugate eigenvalues λ and λ. Write λ = λ ω, ω = 1. Then, J can be written as a product of two commuting factors | | | | Mω M λ Mω | | Mω M λ Mω J =  .   | | . .  ......  Mω   M λ     | |  Let D1,D2 denote these factors. Then, D1 is an  , whereas D2 has only one eigenvalue λ . We claim that the Jordan normal form of D consists of two blocks each of size k. | | 2 Indeed, let T = D2 λ I2k. Then, an easy calculation shows that T is a product of two commuting factors −| |

Mω 02 I2 Mω 02 I2 0 0 T =  .   . .  where 02 = ...... 0 0    M   0   ω   2   k 1    Hence, T is nilpotent, T − has all zero entries except 2 2 upper right block Mωk−1 . Thus, k 1 k × T − has rank 2 and T = 0. This shows the claim. Moreover, the fact that D1 and D2 commute implies that for any m Z, ∈ m m m (3.3) J − (D ) = (D )− =1. || 2 || || 1 || By Lemma 3.3, expansive matrices J and D2 are equivalent. 8 1 Define the matrix B such that S− BS is a block diagonal matrix where each Jordan block 1 J of S− AS is replaced by a matrix D2. This procedure is done for complex eigenvalues λ with corresponding Jordan blocks of the form (3.2). If λ R is a real negative eigenvalue, then we replace k k Jordan block J by J. Finally, if λ∈is a positive eigenvalue, then we do nothing to J. × − 1 By the construction we have defined a block diagonal matrix S− BS for which two complex Jordan blocks of size k for a conjugate pair λ and λ of complex eigenvalues of A correspond to two Jordan blocks of size k for the eigenvalue λ of B. In the case of a negative eigenvalue λ of A, a Jordan block of size k corresponds to a| Jordan| block of B of the same size, but for positive eigenvalue λ . This shows that (ii)-(iv) hold. Finally, to prove| (i)| observe that (3.3) implies that for any m Z, ∈ 1 m m 1 m 1 m S− A− B S = (S− AS)− (S− BS) =1 || || || || Hence, m m 1 sup A− B S S− < . m Z || || ≤ || |||| || ∞ ∈ By Lemma 3.3, A and B are equivalent. 

We are now ready to give the proof of Theorem 3.1.

Proof of Theorem 3.1. Let A be an expansive matrix. Let B be the equivalent matrix with all positive eigenvalues which is guaranteed by Lemma 3.6. By [9, Theorem 1] there exists a real matrix X such that B = exp(X). Since all eigenvalues of B are > 1, all eigenvalues of X are positive. For t R define one-parameter group of dilations At = exp(P ln t), t > 0, 1 ∈ where P = c X and c = tr(X). By Lemma 3.4 dilations At, t > 1, are all equivalent with B = Ac, which in turn is equivalent with A. Finally, the uniqueness of P follows from Lemma 3.5. Indeed, suppose that A′ = { t exp(P ′ ln t) is another one-parameter group of dilations with a generator P ′ satisfy- }t>0 ing (i) and such that B is equivalent to At′ for some t > 1. By Lemma 3.4, B is equivalent to At′ for all t> 1. Choose t0 > 1 such that

det(A) = det(B) = det(A′ ) = exp(ln t tr(P ′)) = t . | | t0 0 0

Since At0 and At′0 are equivalent, have all positive eigenvalues, and det(At0 ) = det(At′0 ), by

Lemma 3.5 we have At0 = At′0 . Likewise, by Lemma 3.4, At and At′ are equivalent, have all positive eigenvalues, and det(At) = det(At′ ) for all t> 1. Thus, At = At′ for all t> 0, which shows the uniqueness. 

As an immediate corollary of Theorem 3.1 we have the following result. A similar result to Corollary 3.7 was observed by Cheshmavar and F¨uhr in [8, Remark 7.11].

Corollary 3.7. Let ρ1 and ρ2 be the quasi-norms associated to dilations A1 and A2, respec- tively. Let P1 and P2 be the generators of one-parameter groups of dilations as in Theorem 3.1 corresponding to A1 and A2, respectively. Then, ρ1 and ρ2 are equivalent if and only if P1 = P2. We can now state the main result of our paper. 9 Theorem 3.8. Let A be an anisotropic dilation. Then, there exists an expansive continuous group At t>0 such that its generator has all positive eigenvalues and discrete and continuous { } p p anisotropic Hardy spaces coincide H = H . That is, for f ′, we have A At { } ∈ S

f Hp f Hp . k k A ≃k k {At} p With this theorem, we are able to associate an anisotropic Hardy space HA with the parabolic PDE (1.1), in the sense that the fundamental solution to (1.1) can be used as a kernel in the radial maximal characterization of Hp . In light of Theorem 2.3 it is tempting At to conclude that Theorem 3.1 already accomplishes{ this.} However, Theorem 3.1 implies only p p that for each t> 1, HA and HAt , both as anisotropic Hardy spaces (with respect to dilations A and At), are equivalent. What we need is slightly stronger: the anisotropic Hardy space Hp , defined via discrete maximal functions, is the same as the parabolic Hardy space Hp , A At defined via a continuous maximal function as shown below. { } Lemma 3.9. Let A be an expansive continuous group and let A = A , where t > 1. { t}t>0 t0 0 Then for any f ′, we have ∈ S f Hp f Hp . k k A ≃k k {At}

Proof. Without loss of generality, by rescaling we can assume that generator P of At t>0 satisfies tr(P ) = 1. We start with the inclusion Hp Hp . Let f Hp ,{ that} is, A At At ⊇ { } ∈ { } ˜ p p Mϕf L . To show that f HA, by Theorem 2.2 it suffices to establish the pointwise inequality∈ ∈ (3.4) M 0f(x) M˜ f(x). ϕ ≤ ϕ Note that for simplicity, we majorize the (discrete) radial maximal operator with a non- tangential (continuous) maximal operator, which follows immediately from the definitions. Indeed, we convert the continuous dilation (on the right) to the discrete dilation (on the left) by setting t =(t )k, where k Z. Since b = det A = t we have 0 ∈ | | 0 k k b− ϕ(A− y)=ϕ ˜(t0)k (y). Taking the supremum over k Z on the left-hand side and over t > 0 andρ ˜(x y) < t, yields (3.4). ∈ − N p 0 p For the reverse inclusion, let N be large enough so that if f HA, then M f L , where ∈ ∈ F ∈ α N = N = ϕ : ϕ α,N := sup ∂ ϕ(x) (1 + x ) 1, α N . F F { ∈ S k k x Rn | | | | ≤ | | ≤ } ∈ Consider a continuous variant of radial and maximal function given by

0 M˜ f(x) = sup sup f ϕ˜t(x) . F ϕ t>0 | ∗ | ∈F In light of Theorem 2.2, it suffices to show that for any ϕ , there exist positive constants ∈F c1 and c2, independent of f, such that

0 0 (3.5) M˜ ϕf(x) c1M˜ f(x) c2M (x). ≤ F ≤ F 10 To establish the first inequality in (3.5) note that for any x Rn we have ∈

M˜ϕf(x) = sup sup f ϕ˜t(y) = sup f ϕ˜t(x + Atz) : t> 0, ρ˜(z) < 1 t>0 ρ˜(x y) 0, φ( )= ϕ( + z) for someρ ˜(z) < 1 . {| ∗ t | · · } The semi-norms of φ can be crudely estimated as

α N N N α N N N φ α,N = sup ∂ ϕ(x) (1+ x z ) 2 (1+ z ) sup ∂ ϕ(x) (1+ x ) 2 (1+ z ) . || || x Rn | | | − | ≤ | | x Rn | | | | ≤ | | ∈ ∈ n Taking supremum over z R such thatρ ˜(z) < 1 shows that φ c1 for some constant c1, which yields the first inequality.∈ ∈ F To establish the second inequality in (3.5), we first show that it holds if t [1, t0] and n ∈ then we extend it to all possible values of t> 0. Fix ϕ . For x R and t [1, t0], we write ∈F ∈ ∈

1 1 (f ϕ˜t)(x) = f(x z)t− ϕ(At− z)dz = f ψ(x) , | ∗ | Rn − | ∗ | Z 1 1 α where ψ(z) = t− ϕ(At− z). Observe that by chain rule, the partial derivatives ∂ ψ, are β 1 controlled by ∂ ϕ, where β α as well as norms of matrices At− , where t [1, t0]. Indeed, by the chain rule, see| |≤| [4, Lemma| 5.5], there exists a constant C > 0 such that∈

1 1 α α 1 N ψ α,N Ct− (At)− | | sup sup ∂ φ((At)− x) (1 + x ) || || ≤ || || β α x Rn | | | | | |≤| | ∈ 1 1 α α N 1 1 α N = Ct− (At)− | | sup sup ∂ φ(x) (1 + Atx ) Ct− (At)− | | At . || || β α x Rn | | | | ≤ || || || || | |≤| | ∈

Since t [1, t0] we deduce that ψ c with c depending on t0. ∈ ∈ F k k+1 Now let t > 0 be arbitrary. Let k Z be such that t [(t0) , (t0) ]. If we define k ∈ ∈ t˜ = t/(t ) , then t˜ [1, t ] and we have the identities A = A˜ k = A˜A k . Then, 0 ∈ 0 t t(t0) t (t0)

1 1 1 k 1 f ϕ˜t(x)= f(x z)t− ϕ((At)− z)dz = f(x z)(t0)− ϕ((At˜(t0)k )− z)dz ∗ Rn − t˜ Rn − Z Z 1 1 = f(x z)ψk(z)dz = (f ψk)(x), t˜ Rn − t˜ ∗ Z 1 k k where ψ(z) = ϕ((At˜)− z) c and ψk(z) = b− ψ(A− z). Therefore, taking the supremum of the continuous dilation∈ overF t > 0, with respect to a test function ϕ , is equivalent to taking the supremum of the discrete dilation over k Z, with respect∈ F to another test function ψ c . Hence, we obtain the second inequality∈ of (3.5). This completes our proof. ∈ F 

Proof of Theorem 3.8. Let A be an expansive dilation. By Theorem 3.1 we can find an associated continuous expansive group A such that its generator has all positive eigen- { t}t>0 values and A is equivalent with dilation At0 for some/all t0 > 1. By Theorem 2.3, discrete anisotropic Hardy spaces Hp and Hp coincide with equivalent quasi-norms. Hence, we A At0 obtain the required conclusion by Lemma 3.9.  11 4. Equivalence of dilations up to linear transformations In this section, we provide the classification of anisotropic Hardy spaces by correcting and expanding the results shown in [2]. The following result was shown by the first author [2, Theorem 10.3].

Theorem 4.1. Let ρ1 and ρ2 be the quasi-norms associated to dilations A1 and A2, respec- tively. If ρ1 and ρ2 are equivalent, then for all r > 1 and all m =1, 2,... (4.1) span ker(A λI)m = span ker(A λI)m, 1 − 2 − λ =rǫ λ =r | [| | [| where (4.2) ε = ε(A , A ) = ln det A / ln det A . 1 2 | 1| | 2| n In (4.1), we treat A1 and A2 as linear maps on C and λ varies over their complex eigen- values. In [2] it was incorrectly claimed the converse to Theorem 4.1 also holds. Cheshmavar and F¨uhr [8, Remark 7.4] have given an example showing that the converse is actually false. To illustrate this, we can use Corollary 3.7: two dilations are equivalent exactly when their corresponding generators (of trace 1) are exactly the same. Consider the following example. Example 4.1. For any c R, consider 2 2 dilation ∈ × 2 c A = . c 0 2   One can easily compute that ln 2 c/2 A = exp . c 0 ln 2    Hence, the generator Pc of a one-parameter group of dilations from Theorem 3.1 correspond- ing to Ac is given by 1/2 c/(2 ln 2) P = . c 0 1/2   By Corollary 3.7, dilations Ac are not equivalent to each other for different choices of c R. Obviously, the choice of c = 0 corresponds to the classical isotropic setting. In general,∈ by Lemma 3.5, matrices of the form 2 ∗2 ∗ ∗ (4.3)  .∗ ∗ ..  ∗  2 are equivalent if and only if all entries above the diagonal are identical. Example 4.1 suggests that it is rare when two dilations are equivalent. The situation changes drastically when we identify dilations up to a similarity. In this scenario dilations 1 A and S− AS, where S GLn(R) are not distinguished. Hence, we are interested in equivalence of quasi-norms∈ up to a linear transformation, see [2, Definition 10.9]. As a consequence of Corollary 4.3 we will see that the number of non-equivalent quasi-norms 12 corresponding to n n matrices of the form (4.3) is actually finite and equal to the partition function p(n). × This is a consequence of the following result, see [2, Theorem 10.10]. Since the original proof of this fact relied on incorrect formulation of [2, Theorem 10.5(ii)] due to the above mentioned problem with the converse of Theorem 4.1, we need to give a corrected proof.

n Theorem 4.2. Suppose we have two dilations A1 and A2 on R . The following are equiva- lent:

(i) the quasi-norms ρ1 and ρ2 associated to A1 and A2, respectively, are equivalent up to a linear transformation, i.e., there is a constant c > 0 and an invertible n n matrix S such that × 1/cρ (x) ρ (Sx) cρ (x) for all x Rn. 1 ≤ 2 ≤ 1 ∈ (ii) for all r> 1 and m =1, 2,... we have

m m ln det A1 (4.4) dim ker(A1 λI) = dim ker(A2 λI) , where ε = | |. − − ln det A2 λ =rε λ =r | X| |X| | | Proof of Theorem 4.2. Suppose that two quasi-norms ρ1 and ρ2 are equivalent up to a linear 1 transformation. Note that ρ (S ) is a quasi-norm associated with the dilation S− A S since 2 · 2 1 1 ρ (S(S− A Sx)) = det A ρ (Sx)= det(S− A S) ρ (x). 2 2 | 2| 2 | 2 | 2 Hence, the quasi-norms ρ1 and ρ2 are equivalent up to a linear transformation if and only if 1 A is equivalent to S− A S for some S GL (R). 1 2 ∈ n Since the quasi-norms ρ1 and ρ2(S ) are equivalent, Theorem 4.1 implies that for any r> 1, m =1, 2,... ·

m 1 m span ker(A λI) = span ker(S− A S λI) 1 − 2 − λ =rǫ λ =r | [| | [| 1 m 1 m = span ker(S− (A λI) S)= S− span (ker(A rId) ) . 2 − 2 − λ =r  λ =r  | [| | [| Hence, (4.4) holds. It remains to show the converse implication (ii) = (i). By Lemma 3.6 there exist ⇒ expansive dilations B1 and B2 with positive eigenvalues which are equivalent to A1 and A2, respectively. Since original dilations A1 and A2 satisfy (4.4), by Lemma 3.6(iv) their positive eigenvalue counterparts B1 and B2 satisfy dim ker(B rεI)m = dim ker(B rI)m, for all r> 1, m =1, 2,... 1 − 2 − Let ρ1′ and ρ2′ be quasi-norms associated to B1 and B2, respectively. Since ρi and ρi′ are equivalent for i = 1, 2, it suffices to show that ρ1′ and ρ2′ are equivalent up to a linear transformation. Using Theorem 3.1 we can rescale one of dilations, say B1, to an equivalent dilation B1′ so that det B1′ = det B2, without affecting the conclusion (i). More precisely, we consider the unique one-parameter group of dilations exp(P ln t) such that B = exp(P ln t ) for { }t>0 1 1 t1 = det B1 > 1, where P is the generator as in Theorem 3.1. Define B1′ = exp(P ln t2), where t2 = det B2. Then, B1′ is equivalent to B1 and det B1′ = det B2. 13 Moreover, we claim that m m (4.5) dim ker(B′ rI) = dim ker(B rI) for all r> 1, m =1, 2,... 1 − 2 − Indeed, if the normal Jordan form of P has a Jordan block of size k corresponding to an eigenvalue λ > 0, then exp(tP ) has in its normal Jordan form a block of the same size corresponding to an eigenvalue eλt for any t R. In other words, ∈ dim ker(P λI)m = dim ker(exp(P ln t) tλI)m for all t,λ > 0, m =1, 2,... − − λ Take any r > 1 and write it as r = (t2) for some λ > 0. Since ε = ln t1/ ln t2 we have ε λ r =(t1) . Hence, dim ker(exp(P ln t ) rI)m = dim ker(P λI)m = dim ker(exp(P ln t ) rεI)m. 2 − − 1 − This shows (4.5). Finally, observe that for any matrix B, the number of Jordan blocks of size m corre- sponding to an eigenvalue r is equal to ≥ m m 1 dim ker(B rI) dim ker(B rI) − . − − − Hence, (4.5) implies that the number of Jordan blocks of size m corresponding to an eigen- value r is the same for both B1′ and B2. Therefore, the matrices B1′ and B2 have the same Jordan normal form. In other words, there is an invertible n n matrix S such that 1 × B1′ = S− B2S. This implies that the quasi-norm associated to B1′ , which is equivalent to ρ1′ , is ρ′ (S ). This proves that quasi-norms ρ′ and ρ′ are equivalent up to linear transformations 2 · 1 2 and so are ρ1 and ρ2.  Theorem 4.2 implies the following classification of expansive dilations according to their Jordan normal form. Corollary 4.3. For any k n, take any ≤ (4.6) n ,...,n N n + ... + n = n, 1 k ∈ 1 k (4.7) 1 <λ <...<λ < 2 λn1 λnk =2. 1 k 1 ··· k i and partitions π , i =1,...,k of ni, i.e., (4.8) πi =(πi ... πi ) Nmi πi + ... + πi = n . 1 ≥ ≥ mi ∈ 1 mi i For specified parameters (4.6), (4.7), and (4.8), define the corresponding block diagonal ma- trix 1 k (4.9) A(λ1,...,λk; π ,..., π ) i consisting of Jordan blocks for eigenvalues λi, i = 1,...,k, and sizes πj, j = 1,...,mi. Then, any expansive dilation A in Rn is equivalent up to a linear transformation to some 1 k dilation A(λ1,...,λk; π ,..., π ). Moreover, this correspondence is 1-to-1. That is, dilations 1 k of the form (4.9) for distinct choices of eigenvalues λ1,...,λk and partitions π ,..., π are not equivalent up to linear transformations. Proof. By Theorem 3.1, we can assume replace A by an equivalent dilation B with all positive eigenvalues. In addition, by rescaling we can assume that det B = 2. Then, B has a Jordan normal form (4.9) for some eigenvalues (4.7) and the corresponding Jordan blocks of sizes 14 given by partitions (4.8). Now, if we modify any of the partitions πi, i =1,...,k, or any of the eigenvalues λi, then we necessarily change the value of dim ker(A(λ ,...,λ ; π1,..., πk) λI)m 1 k − for some λ> 1 and m =1, 2,... Theorem 4.2 guarantees that two different dilations of the 1 k form A(λ1,...,λk; π ,..., π ) are not mutually equivalent up to linear transformation.  p Rn p Rn Recall that two anisotropic Hardy spaces HA1 ( ) and HA2 ( ) are equivalent up to linear transformations if there exists an invertible n n matrix S such that the dilation operator 1 × p Rn p Rn f f(S− ) defines an between HA1 ( ) and HA2 ( ), see [2, Definition 7→ · 1 10.9]. This happens precisely if A1 and S− A2S are equivalent dilations, see the proof of [2, Theorem 10.10]. As a consequence of Corollary 4.3 dilations of the form (4.9) classify p Rn anisotropic Hardy spaces HA( ) up to linear transformations. n Theorem 4.4. Suppose we have two dilations A1 and A2 on R . The following are equiva- lent: 1 (i) A1 and S− A2S are equivalent for some n n invertible matrix S, (ii) (4.4) holds for all r > 1 and m =1, 2,..., × p Rn p Rn (iii) HA1 ( ) and HA2 ( ) are equivalent up to linear transformations for all 0

0 5.1. Parabolic Setting. To set the context for these questions, we fix an expansive con- tinuous group A , which defines the associated parabolic differential equation (1.1): { t} ∂u 1 1 = (D− ∆D )u = L u. ∂t t · t t t In the frequency domain, the fundamental solution Φ is given in a simple form. Proposition 5.1 ([6, Section 1.3]). Let Φ be defined by ∈ S (5.1) Φ(ˆ ξ) = exp[ 4π2 Bξ,ξ ], − h i 1 1 where B is a norm-inducing matrix B for A∗ . Then Φ˜ (x) = t− Φ(A− x) satisfies the { t } t t differential equation (1.1). Moreover, if f ′, then u(x, t) = f Φ˜ t(x) also satisfies the same equation. ∈ S ∗ Proposition 5.1 is an elementary result, but we include the proof for completeness. Its proof does not require the assumption (2.1) made in [6]. However, it does require the following property of a norm-inducing matrix B for A∗ , which we state again: { t } d 1 1 (5.2) BA∗x, A∗x = (BP ∗ + P B)A∗x, A∗x = A∗x, A∗x . dth t t i t h t t i t h t t i

Proof. If Φ˜ t is a solution for the PDE (1.1), then by taking the Fourier transform we obtain 2 ∂ 4π At∗ξ, At∗ξ (5.3) [Φ(ˆ A∗ξ)] = − h iΦ(ˆ A∗ξ). ∂t t t t If Φ is given by (5.1), then a simple calculation using (5.2) shows that it satisfies (5.3). Conversely, observe that for fixed ξ Rn, (5.3) can be seen as an ordinary differential equation of the form ∈ 2 d 4π A∗ξ, A∗ξ h(t)= − h t t ih(t). dt t s(t) 2 Hence, its solution must be of the form h(t)= c0e , where s′(t)= 4π At∗ξ, At∗ξ /t. By the 2 −ˆ h i property (5.2), we have s(t)=( 4π ) BAt∗ξ, At∗ξ . Since h(t) = Φ(At∗ξ), the fundamental − 2 h i n solution Φ satisfies Φ(ˆ ξ) = exp[ 4π Bξ,ξ ]. Consequently, if f ′, then F : R (0, ) − h i ∈ S × ∞ → C, defined by F (x, t)= f Φ˜ (x), is also a solution to (1.1).  ∗ t 1 Now observe that the parabolic PDE, associated with the operator Lt = (Dt− ∆Dt)/t, depends on t, for which there is no viable semigroup theory. Indeed, if we naively set 16 T f = Φ˜ f, a computation with the Fourier transform gives t t ∗ 2 (TtTrf)∧(ξ) = exp( 4π A t BAt∗ + A r BAr∗)ξ, At∗+rξ ). − h t+r t+r i J

We will have semigroup exactly if the J-term satisfies the identity A t BAt∗ + A r BAr∗ = | {z } t+r t+r tLt p BAt∗+r, which is not likely. So we cannot make sense of the operation e− f, nor use HLt to denote Hp . This leads to the following problem. Given a dilation matrix A and At { } its associated expansive group At , and having established the equivalence between the anisotropic and parabolic Hardy{ spaces} Hp Hp , investigate the connection with the A At ≃ { } theory of Hardy associated with operators. 5.2. Alternative Parabolic Approach. By seeking a semigroup structure associated with the expansive group A , we formulate an alternative definition of Hardy space. To do this, { t} we fix a continuous group At t>0. Letρ ˜ be the quasi-norm associated with At∗ , and define the alternative fundamental{ } solution∗ { } (5.4) Φ(ˆ ξ) = exp( 4π2ρ2(ξ)). − ∗ Then we have a semigroup T f = Φ˜ f(x), as made apparent by the Fourier transform t √t ∗ and the homogeneity propertyρ ˜ (At∗ξ)= tρ˜ (ξ), ∗ ∗ \ 2 2 2 2 ˆ 2 2 √ TtTsf(ξ) = exp( 4π ρ˜ (A√∗ tξ)) exp( 4π ρ˜ (A√∗ sξ))f(ξ) = exp( 4π ρ (ξ s + t)) − ∗ − ∗ − ∗ \ = Ts+tf(ξ).

Furthermore, Φ given by (5.4) is the fundamental solution of ∂tu = Lu, defined in frequency by 2 2 (5.5) ∂tuˆ(ξ)= Lu(ξ)= 4π ρ˜ (ξ) uˆ(ξ). − ∗ The operator L is the infinitesimal generator of the semigroup Tt , which is defined formally Ttf f { } c + for f in the domain of L, by Lf = limt 0 t− . Then we are now in a position to define p → the space H as all tempered distributions f ′ such that L ∈ S 0 tL ˜ p Φf = sup e− f = sup Φ√t f L . M t>0 | | t>0 | ∗ | ∈ We can now attempt to place the pseudo-differential operator L from (5.5) among existing p literature. We start with a basic question concerning the nature of these HL spaces. Given A and its dual A∗ , we can have more than one homogeneous norm ρ∗ to use in defining { t} { t } the fundamental solution Φ.ˆ For example, even in the diagonal case, fix δ > 0, and define tδ 0 δ 0 A = with P = . t 0 tδ 0 δ     Since At = At∗, we do not need to make a distinction between the homogeneous norms ρ or ρ∗. Associated with this group are two natural choices of quasi-norms. The first 1 is the canonical quasi-norm, from solving the identity A− x = 1, which gives the norm | t | ρ˜ (x) = x 1/δ. The second is given byρ ˜ (x) = x 2/δ + x 2/δ, so that it satisfies the 1 | | 2 | 1| | 2| homogeneityρ ˜(Atx)= tρ˜(x). Their geometries do differ:ρ ˜1(x) = 1 exactly when x is on the p 2/δ boundary of the Euclidean unit ball, whileρ ˜2(x)=1 for x on the boundary of ℓ unit ball 2/δ 2 2/δ given by x ℓ2/δ = j=1 xj = 1. k k | | 17 P A fundamental question is to consider whether the resulting Hardy spaces, from the two norms, agree. Consider the example when δ = 2, with the homogeneous normsρ ˜ (ξ)= ξ 1/2 1 | | andρ ˜ (ξ)= ξ + ξ . Their respective fundamental solutions are Φ(1)(ξ) = exp( 4π2 ξ ) 2 | 1| | 2| − | | (2) 2 2 and Φ (ξ) =p exp( 4π ξ1 ) exp( 4π ξ2 ), which under the inverse Fourier transform, are − | | (1) − | | d given by the Φ (x) = P2(x) and the product of two Poisson kernels in R, (2) 2 (n+1)/2 whichd we denote by Q(x) = Φ (x)= P1(x1)P1(x2), where Pn(x)= cn(1+ x )− . We proceed along the classical arguments, see [19, Chapter III.1] or [13, Chapter| | 2.1]. Observe that the first norm leads to the classical Hardy space. Denote the Hardy space p 0 associated with the kernel Q to be H , with the norm f p = f for f ′, defined Q HQ Q p 2 k k kM k ∈ S formally for f ′(R ) for which f Qs(x) is defined for s > 0. We can readily establish the inclusion H∈p S Hp by decomposing∗ Q along each variable, and obtain ⊆ Q

∞ (k+j) k (j,k) x1 x2 Q(x)= Q (x )Q (x )= 2− 2− ϕ , , 1 1 2 2 2j 2k j,kX=0   (j,k) (k) (j) (k) where ϕ (x1, x2) = Φ (x1)Φ (x2), and Φ are smooth cutoff functions in R and bounded in (R2). Then we can majorize the radial maximal function of Q with respect to S p p a grand maximal function, and obtain the inclusion that H HQ. However, the reverse inclusion is unknown. Classically, one⊆ defines a test function by Ψ(x)= ∞ η(s)P (x)dx, where η is smooth on [1, ), and satisfies 1 s ∞

R ∞ ∞ η(s)ds =1, skη(s)ds = 0 for k N. ∈ Z1 Z1 The Ψ is shown to be in , and majorizes the maximal function associated with the Poisson kernel. However, when weS use Q in this construction, Ψ cannot be shown to be smooth, given the lack of differentiability of Qˆ along the ξ1 and ξ2 axes. This leaves open the question p p of whether we do have HQ H , or, more generally, if two homogeneous norms to the same continuous group leads to the⊆ same Hardy space. It is worth adding that if p> 1, then the p p p Hardy space H and HQ actually coincide with L by [19, Chapter II.4]. In general, the p p fact the Hardy spaces HL coincide with L spaces for p > 1 is related to the boundedness of maximal functions along curves due to Stein and Wainger [20]. However, the following problem remains open.

Question 1. Given an essentially continuous expansive group A∗ , with two homogeneous { t } quasi-normsρ ˜1∗ andρ ˜2∗, consider the resulting fundamental solution and PDE, given by the differential operators L1 and L2, respectively. Do they result in the same Hardy space, that is, Hp = Hp , 0

0, ⊂ 2 tL dist(U1, U2) e− f , f C exp f f , |h 1 2i| ≤ − ct k 1k2k 2k2   tL where e− f = Φ˜ f and dist(U , U ) = inf ρ˜(x y): x U , x U ? 1 √t ∗ 1 2 { − 1 ∈ 1 2 ∈ 2} References [1] P. Auscher and J. M. Martell. Weighted norm inequalities, off-diagonal estimates and elliptic operators. III. of elliptic operators. J. Funct. Anal., 241(2):703–746, 2006. [2] M. Bownik. Anisotropic Hardy spaces and wavelets. Mem. Amer. Math. Soc., 164(781):vi+122, 2003. [3] Marcin Bownik. Atomic and molecular decompositions of anisotropic Besov spaces. Math. Z., 250(3):539–571, 2005. [4] Marcin Bownik, Baode Li, and Jinxia Li. Variable anisotropic singular integral operators. preprint arXiv:2004.09707. [5] The Anh Bui, Jun Cao, Luong Dang Ky, Dachun Yang, and Sibei Yang. Musielak-Orlicz-Hardy spaces associated with operators satisfying reinforced off-diagonal estimates. Anal. Geom. Metr. Spaces, 1:69– 129, 2013. [6] A. Calder´on and A. Torchinsky. Parabolic maximal functions associated with a distribution. Advances in Math., 16:1–64, 1975. [7] A. Calder´on and A. Torchinsky. Parabolic maximal functions associated with a distribution. II. Advances in Math., 24(2):101–171, 1977. [8] Jahangir Cheshmavar and Hartmut F¨uhr. A classification of anisotropic Besov spaces. Appl. Comput. Harmon. Anal., 49(3):863–896, 2020. [9] Walter J. Culver. On the existence and uniqueness of the real logarithm of a matrix. Proceedings of the American Mathematical Society, 17(5):1146–1151, 1966. [10] Shai Dekel, Pencho Petrushev, and Tal Weissblat. Hardy spaces on Rn with pointwise variable anisotropy. J. Fourier Anal. Appl., 17(5):1066–1107, 2011. [11] Xuan Thinh Duong and Lixin Yan. Duality of Hardy and BMO spaces associated with operators with bounds. J. Amer. Math. Soc., 18(4):943–973 (electronic), 2005. [12] C. Fefferman and E. Stein. Hp spaces of several variables. Acta Math., 129(3-4):137–193, 1972. [13] L. Grafakos. Modern Fourier analysis, volume 250 of Graduate Texts in Mathematics. Springer, New York, second edition, 2009. [14] Steve Hofmann, Guozhen Lu, Dorina Mitrea, Marius Mitrea, and Lixin Yan. Hardy spaces associated to non-negative self-adjoint operators satisfying Davies-Gaffney estimates. Mem. Amer. Math. Soc., 214(1007):vi+78, 2011. [15] Steve Hofmann and Svitlana Mayboroda. Hardy and BMO spaces associated to divergence form elliptic operators. Math. Ann., 344(1):37–116, 2009. [16] Jos´eMar´ıaMartell and Cruz Prisuelos-Arribas. Weighted Hardy spaces associated with elliptic op- erators. Part I: Weighted norm inequalities for conical square functions. Trans. Amer. Math. Soc., 369(6):4193–4233, 2017. [17] Jos´eMar´ıaMartell and Cruz Prisuelos-Arribas. Weighted Hardy spaces associated with elliptic opera- H1 w tors. Part II: Characterizations of L( ). Publ. Mat., 62(2):475–535, 2018. [18] Pierre Portal. Maximal and quadratic Gaussian Hardy spaces. Rev. Mat. Iberoam., 30(1):79–108, 2014. [19] E. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III. [20] E. Stein and S. Wainger. Problems in harmonic analysis related to curvature. Bull. Amer. Math. Soc., 84(6):1239–1295, 11 1978. [21] Qingying Xue, Yong Ding, and KˆozˆoYabuta. Parabolic Littlewood-Paley operators. Math. Nachr., 282(1):125–141, 2009. 19 [22] Dachun Yang, Junqiang Zhang, and Ciqiang Zhuo. Variable Hardy spaces associated with operators satisfying Davies-Gaffney estimates. Proc. Edinb. Math. Soc. (2), 61(3):759–810, 2018.

Department of Mathematics, University of Oregon Email address: [email protected] Department of Mathematics and Statistics, Sam Houston State University Email address: [email protected]

20