This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Approximation by the Bessel–Riesz Quotient

Ikemefuna Agbanusi Department of Mathematics and Computer Science, Colorado College, [email protected]
Abstract

How large is the Bessel potential, Gα,μfG_{\alpha,\mu}f, compared to the Riesz potential, IαfI_{\alpha}f, of a given function? We prove that, for certain ff and pp,

G1,μfpω(I1f,1/μ)p,\|G_{1,\mu}f\|_{p}\approx\omega(I_{1}f,1/\mu)_{p},

where ω(f,t)p\omega(f,t)_{p} is the LpL^{p} modulus of continuity. However, for 0<α<10<\alpha<1,

Gα,μfpC(ω(Iαf,1/μ)p)αIαfp1α.\|G_{\alpha,\mu}f\|_{p}\leq C(\omega(I_{\alpha}f,1/\mu)_{p})^{\alpha}\cdot\|I_{\alpha}f\|^{1-\alpha}_{p}.

These estimates are obtained by studying the quotient of the two operators, Eα,μ:=(Δ)α/2(μ2Δ)α/2E_{\alpha,\mu}:=\frac{(-\Delta)^{\alpha/2}}{(\mu^{2}-\Delta)^{\alpha/2}}, and exploiting its approximation theoretic properties. Additionally, Gα,μf=𝒪(μα2)G_{\alpha,\mu}f=\mathcal{O}(\mu^{-\frac{\alpha}{2}}) if IαfI_{\alpha}f vanishes near a given point. This “localization” result is derived from kernel estimates of Eα,μE_{\alpha,\mu}.

1 Overview

This paper deals with some topics at the intersection of Fourier analysis, potential theory and approximation theory. Specifically, for μ>0\mu>0, we study the Bessel–Riesz quotient, Eα,μE_{\alpha,\mu}, which is the operator defined by the multiplier

mα,μ(ξ):=|ξ|α(μ2+|ξ|2)α2.m_{\alpha,\mu}(\xi):=\dfrac{\left|\xi\right|^{\alpha}}{(\mu^{2}+\left|\xi\right|^{2})^{\frac{\alpha}{2}}}. (1)

Here f^(ξ)\hat{f}(\xi) denotes the Fourier transform while Gα,μf^(ξ):=(μ2+|ξ|2)α2f^(ξ)\widehat{G_{\alpha,\mu}f}(\xi):=(\mu^{2}+|\xi|^{2})^{-\frac{\alpha}{2}}\hat{f}(\xi) and Iαf^(ξ):=|ξ|αf^(ξ)\widehat{I_{\alpha}f}(\xi):=|\xi|^{-\alpha}\hat{f}(\xi) define the family of Bessel and Riesz potentials respectively. We focus mainly on the case 0<α10<\alpha\leq 1.

The Riesz and Bessel potentials are indispensable in potential theory, the theory of Sobolev spaces, elliptic PDEs, and describing the fine structure of functions and sets to name just a few applications. Quantifying the relationship between these two classical operators is also an important problem. For the “standard” case μ=1\mu=1, this comparison has been expressed in the literature in several ways—in terms of their capacities as in Ziemer [15, pg. 67] and via fractional maximal functions as in Adams–Hedberg [1, Theorem 3.6.2].

Since Gα,μf=Eα,μIαfG_{\alpha,\mu}f=E_{\alpha,\mu}I_{\alpha}f, norm or pointwise estimates for Eα,μE_{\alpha,\mu} allow for a direct comparison of the two potentials. This is one reason for introducing and studying this operator. Moreover, the approach using the Bessel–Riesz quotient appears better at handling the behaviour as μ\mu\to\infty. This is because we can apply some approximation theoretic tools as discussed below.

At least two observations point to the connection with approximation theory. The first is the trivial fact that mα,μ(ξ)0m_{\alpha,\mu}(\xi)\to 0 pointwise as μ\mu\to\infty. The second observation starts with a formula from Stein [13, §5.3.2]

|ξ|α(μ2+|ξ|2)α2=(1μ2μ2+|ξ|2)α2=1j=1aα,j(1+|ξμ1|2)j,\dfrac{\left|\xi\right|^{\alpha}}{(\mu^{2}+\left|\xi\right|^{2})^{\frac{\alpha}{2}}}=\left(1-\frac{\mu^{2}}{\mu^{2}+\left|\xi\right|^{2}}\right)^{\frac{\alpha}{2}}=1-\sum_{j=1}^{\infty}a_{\alpha,j}(1+|\xi\mu^{-1}|^{2})^{-j},

for some positive coefficients with aα,j=1\sum a_{\alpha,j}=1. By Fourier inversion we obtain

Eα,μf(x)=f(x)Tα,μf(x),E_{\alpha,\mu}f(x)=f(x)-T_{\alpha,\mu}f(x), (2)

where Tα,μT_{\alpha,\mu} has the convolution kernel Aα,μ(z)A_{\alpha,\mu}(z) defined as

Aα,μ(z):=μdj=1aα,jG2j(μz).A_{\alpha,\mu}(z):=\mu^{-d}\sum_{j=1}^{\infty}a_{\alpha,j}G_{2j}(\mu z). (3)

Again, Gs(z)G_{s}(z) is the Bessel kernel whose Fourier transform is (1+|ξ|2)s2(1+|\xi|^{2})^{-\frac{s}{2}}. Their well known properties imply that Aα,μ(z)A_{\alpha,\mu}(z) is a positive, radial, integrable function with L1L^{1} norm Aα,μ1=j=1aα,jG2j1=1\|A_{\alpha,\mu}\|_{1}=\|\sum_{j=1}^{\infty}a_{\alpha,j}G_{2j}\|_{1}=1. Note that Aα,μ(z)A_{\alpha,\mu}(z) is also singular at the origin since lim|z|0G2j(z)=\lim_{|z|\to 0}G_{2j}(z)=\infty when 0<2jd0<2j\leq d. In any case, Tα,μT_{\alpha,\mu} is an approximate identity and, by (2), Eα,μE_{\alpha,\mu} is its approximation error.

With any approximation scheme, the main task is quantifying the error, Eα,μfp\|E_{\alpha,\mu}f\|_{p}. The error turns out to be related to the LpL^{p} modulus of continuity defined by

ω(f,t)p:=sup|h|tf(+h)f()p.\omega(f,t)_{p}:=\sup_{|h|\leq t}\|f(\cdot+h)-f(\cdot)\|_{p}. (4)

The exact dependence is our first main result and is summarized next. Here XYX\approx Y means that C1YXCYC^{-1}Y\leq X\leq CY for some C>0C>0.

Theorem 1.1.

Suppose fLp(d)f\in L^{p}(\mathbb{R}^{d}).

  1. (a)

    If α=1\alpha=1 and 1<p<1<p<\infty then

    E1,μfpω(f,1/μ)p.\|E_{1,\mu}f\|_{p}\approx\omega(f,1/\mu)_{p}. (5)
  2. (b)

    If α=1\alpha=1 and p=1p=1 there is a constant k=k(d)k=k(d) such that

    E1,μf1kω(f,1/μ)1logω(f,1/μ)1.\|E_{1,\mu}f\|_{{}_{1}}\leq k\omega(f,1/\mu)_{1}\log\omega(f,1/\mu)_{1}. (6)
  3. (c)

    If 0<α<10<\alpha<1 and 1p<1\leq p<\infty there is a constant C=C(α,d)C=C(\alpha,d) such that

    Eα,μfpC(ω(f,1/μ)p)αfp1α\|E_{\alpha,\mu}f\|_{p}\leq C(\omega(f,1/\mu)_{p})^{\alpha}\cdot\|f\|^{1-\alpha}_{p} (7)

The novelty here is the L1L^{1} estimate (6) and the interpolation type inequality (7). It turns out that (5) is implicit in Colzani [5] and Liu–Lu [10], but we give an independent proof tailored to our operator. We do not know if (6) or (7) is sharp.

Theorem 1.1 and the identity Gα,μf=Eα,μIαfG_{\alpha,\mu}f=E_{\alpha,\mu}I_{\alpha}f have the following corollary.

Corollary 1.2.
  1. (a)

    If 1<p<1<p<\infty and I1fLp(d)I_{1}f\in L^{p}(\mathbb{R}^{d}), we have

    G1,μfpω(I1f,1/μ)p.\|G_{1,\mu}f\|_{p}\approx\omega(I_{1}f,1/\mu)_{p}.
  2. (b)

    If I1fL1(d)I_{1}f\in L^{1}(\mathbb{R}^{d}), then

    G1,μf1kω(I1f,1/μ)1logω(I1f,1/μ)1.\|G_{1,\mu}f\|_{{}_{1}}\leq k\omega(I_{1}f,1/\mu)_{1}\log\omega(I_{1}f,1/\mu)_{1}.
  3. (c)

    If 0<α<10<\alpha<1, 1p<1\leq p<\infty and IαfLp(d)I_{\alpha}f\in L^{p}(\mathbb{R}^{d}), then

    Gα,μfpC(ω(Iαf,1/μ)p)αIαfp1α.\|G_{\alpha,\mu}f\|_{p}\leq C(\omega(I_{\alpha}f,1/\mu)_{p})^{\alpha}\cdot\|I_{\alpha}f\|^{1-\alpha}_{p}.

This gives the precise sense in which the Bessel potential “fine tunes” the Riesz potential. Note that the Hardy–Littlewood–Sobolev inequality gives conditions for the assumptions of Corollary 1.2 to hold. Corollary 1.2 also connects the Riesz potential and the truncated maximal function. The latter is defined as follows:

Mα,δf(x)=suprδ1rdα|xy|r|f(y)|𝑑y.M_{\alpha,\delta}f(x)=\sup_{r\leq\delta}\frac{1}{r^{d-\alpha}}\int_{|x-y|\leq r}|f(y)|\,dy. (8)
Corollary 1.3.

If p>1p>1 and I1fLp(d)I_{1}f\in L^{p}(\mathbb{R}^{d}),

ω(I1f,1/μ)pCM1,1μfp.\omega(I_{1}f,1/\mu)_{p}\leq C\|M_{1,\frac{1}{\mu}}f\|_{p}.

This is an immediate consequence of a result of Schechter [12, Theorem 3.5] which in turn refines the Muckenhoupt–Wheeden [11, Theorem 1] estimate for fractional integrals. Corollary 1.3 is likely a known result, but we have not found it stated anywhere. It would be nice to have a more direct proof.

Returning to the approximation properties of Eα,μE_{\alpha,\mu}, it is natural to seek the best or worst possible order of approximation. So called “saturation theorems” giving the best possible rate can be found in §3 when α=1\alpha=1. Finding the worst rate of approximation is tantamount to knowing how slow ω(f,t)p0\omega(f,t)_{p}\to 0 as t0t\to 0 as ff ranges in LpL^{p}. This is an open problem as far as we know though there are partial results due to Besov–Stechkin for L2()L^{2}(\mathbb{R}) in [3].

Theorem 1.1 also gives another characterization of Besov spaces (see section on Notation for the relevant definitions).

Corollary 1.4.

Fix 0<s<10<s<1. For 1<p<1<p<\infty and 0<q0<q\leq\infty, we have

|f|Bp,qsq1(μsE1,μfp)qdμμ.|f|^{q}_{{B}^{s}_{p,q}}\approx\int_{1}^{\infty}(\mu^{s}||E_{1,\mu}f||_{p})^{q}\,\frac{d\mu}{\mu}.

We now turn to issues of pointwise convergence. Let HL\mathcal{M}_{\text{HL}} denote the Hardy–Littlewood maximal operator. As Aα,1(z)A_{\alpha,1}(z) is radially decreasing, we have the maximal inequality supμ>0|(Aα,μf)(x)|CHLf(x)\sup_{\mu>0}|(A_{\alpha,\mu}\star f)(x)|\leq C\mathcal{M}_{\text{HL}}f(x) (see Stein [14, §2.2.1]). This implies a weak type (1,1)(1,1) boundedness from which convergence of Tα,μf(x)T_{\alpha,\mu}f(x) to f(x)f(x) for almost all xx follows.

However, to deal with convergence at a specified point, or to study the structure of the set where pointwise convergence holds, we need good kernel estimates. A direct attack on the series (3) defining Aα,μA_{\alpha,\mu} appears unwieldy. In fact, we could not make this approach work because the known estimates for Bessel kernels are not precise enough to be summed in an infinite series.

Note that if b(ξ)b(\xi) is either mα,μ(ξ)m_{\alpha,\mu}(\xi) or 1mα,μ(ξ)1-m_{\alpha,\mu}(\xi), then b(ξ)b(\xi) satisfies

|ξβb(ξ)|Cβ|ξ||β|;ξ0.|\partial^{\beta}_{\xi}b(\xi)|\leq C_{\beta}\left|\xi\right|^{-|\beta|};\quad\xi\neq 0. (9)

Incidentally, this implies the uniform LpL^{p} boundedness of Eα,μE_{\alpha,\mu} for 1<p<1<p<\infty, but still does not give fast enough decay at infinity for the kernel. A more detailed analysis actually shows that

|ξβb(ξ)|Cα,β|ξ|α|β|(μ2+|ξ|2)α2.|\partial^{\beta}_{\xi}b(\xi)|\leq C_{\alpha,\beta}|\xi|^{\alpha-|\beta|}(\mu^{2}+|\xi|^{2})^{-\frac{\alpha}{2}}. (10)

This refinement is the main ingredient in the following result.

Theorem 1.5.

Suppose b(ξ)Lb(\xi)\in L^{\infty} and satisfies (10) for 0<α10<\alpha\leq 1. Then its kernel B(x)B(x) satisfies

|xγB(x)|Cα,γ,d{|2μx|α2|x||γ|d;|μx|>1,(|μx|2+1)α2|x||γ|d;|μx|1.|\partial_{x}^{\gamma}B(x)|\leq C_{\alpha,\gamma,d}\begin{cases}|2\mu x|^{-\frac{\alpha}{2}}|x|^{-|\gamma|-d};&|\mu x|>1,\\ (|\mu x|^{2}+1)^{-\frac{\alpha}{2}}|x|^{-|\gamma|-d};&|\mu x|\leq 1.\end{cases}

Near the origin, this is the standard estimate for Calderon–Zygmund type kernels. However, the extra decay at infinity has the following quantitative localization principle as a consequence. The proof is short enough to be given here.

Corollary 1.6.

If ff in LpL^{p} vanishes near x0x_{0}, then Eα,μf(x0)=𝒪(μα2)E_{\alpha,\mu}f(x_{0})=\mathcal{O}(\mu^{-\frac{\alpha}{2}}). In particular, if IαfI_{\alpha}f is in LpL^{p} and vanishes near x0x_{0}, then Gα,μf(x0)=𝒪(μα2)G_{\alpha,\mu}f(x_{0})=\mathcal{O}(\mu^{-\frac{\alpha}{2}}).

Proof.

By translation invariance, we may assume that x0=0x_{0}=0, and δ>0\delta>0 is such that f=0f=0 for |x|<δ|x|<\delta. If μδ>1\mu\delta>1, Theorem 1.5 applied to mα,μ(ξ)m_{\alpha,\mu}(\xi) gives

|Eα,μf(0)|=||y|>δKα,μ(y)f(y)𝑑y|Cμα2|y|>δ|f(y)||y|d+α2𝑑y,|E_{\alpha,\mu}f(0)|=\left|\int_{|y|>\delta}K_{\alpha,\mu}(-y)f(y)\,dy\right|\leq\frac{C}{\mu^{\frac{\alpha}{2}}}\int_{|y|>\delta}\frac{|f(y)|}{|y|^{d+\frac{\alpha}{2}}}\,dy,

where Kα,μK_{\alpha,\mu} is its kernel. By Hölder’s inequality,

|Eα,μf(0)|Cd,pμα2δ(dp+α2)fpCd,δ,pμα2fp.|E_{\alpha,\mu}f(0)|\leq C_{d,p}\mu^{-\frac{\alpha}{2}}\delta^{-(\frac{d}{p}+\frac{\alpha}{2})}\|f\|_{p}\leq C_{d,\delta,p}\mu^{-\frac{\alpha}{2}}\|f\|_{p}.

The proof is completed by appealing to the identity Gα,μf=Eα,μIαfG_{\alpha,\mu}f=E_{\alpha,\mu}I_{\alpha}f. ∎

The main idea in this paper is ultimately a simple one—to compare two operators, examine their quotient! Unfortunately, it gets bogged down in a morass of notation and computations. Luckily they are fairly straightforward. In fact, the deepest results used are the (Hörmander–Mikhlin) multiplier theorem and the equivalence of KK-functionals as in [9].

It would be interesting to know if similar approximation and kernel estimates hold for, say, L(x,D)μ2+L(x,D)\frac{\sqrt{L(x,D)}}{\sqrt{\mu^{2}+L(x,D)}} where L(x,D)L(x,D) is a linear second order differential operator. Indeed, it may be worth seeking sharp kernel estimates for pseudo-differential operators, i.e those with non-constant coefficient symbols, satisfying appropriate variants of (10). We hope to treat these and other issues elsewhere.

Before leaving this section we mention that the Bessel–Riesz quotient also surfaced in the author’s unpublished study of transmission boundary value problems with a large parameter. Models of scattering by high contrast materials are good examples to keep in mind here. In these problems, the unknown Dirichlet and Neumann traces of the solution on the “interface” are related by operators similar to the Bessel–Riesz quotient. This was separate motivation for examining this operator in detail.

An outline of the paper now follows. We first introduce some notation and definitions in the upcoming subsection. The proof of Theorem 1.1 is presented in §2 along with some extensions to Hardy spaces. As indicated earlier, the Saturation Theorems are proven in §3. The “Fourier Analytic” proof of Theorem 1.5 is given in §4 and shown to apply to a wider class of symbols.

Notation

Everything takes place in dd–dimensional Euclidean space d\mathbb{R}^{d} and for 1p1\leq p\leq\infty, Lp=Lp(d)L^{p}=L^{p}(\mathbb{R}^{d}) are the usual Lebesgue spaces with norm denoted by fp\|f\|_{p}.

For a non-negative integer kk, the Sobolev space WpkW^{k}_{p} consists of LpL^{p} functions having distributional derivatives up to order kk in LpL^{p}. In WpkW^{k}_{p} we use the norm fWpk=|γ|kDγfp\|f\|_{W^{k}_{p}}=\sum_{|\gamma|\leq k}\|D^{\gamma}f\|_{p}, and seminorm |f|Wpk=|γ|=kDγfp|f|_{W^{k}_{p}}=\sum_{|\gamma|=k}\|D^{\gamma}f\|_{p}.

The direct and inverse Fourier transform of ff and g^\hat{g} respectively defined as

f^(ξ)=eixξf(x)𝑑x;gˇ(x)=(2π)deixξg^(ξ)𝑑ξ.\hat{f}(\xi)=\int e^{-ix\cdot\xi}f(x)\,dx;\qquad\check{g}(x)=(2\pi)^{-d}\int e^{ix\cdot\xi}\hat{g}(\xi)\,d\xi.

When convenient, we also use xξ\mathcal{F}_{x\to\xi} and ξx\mathcal{F}_{\xi\to x} for the direct and inverse transform. For suitable functions a(ξ)a(\xi), we associate the operator

a(D)f(x)=(2π)deixξa(ξ)f^(ξ)𝑑ξ.a(D)f(x)=(2\pi)^{-d}\int e^{ix\cdot\xi}a(\xi)\widehat{f}(\xi)\,d\xi.

The Bessel potential space ps\mathcal{L}_{p}^{s} is defined as ps:={Gsf:fLp}\mathcal{L}_{p}^{s}:=\{G_{s}\star f:f\in L^{p}\}. Here Gs(x)G_{s}(x) is the Bessel potential defined by

Gs(x)=ξx[(1+|ξ|2)s2].G_{s}(x)=\mathcal{F}_{\xi\to x}\left[(1+|\xi|^{2})^{-\frac{s}{2}}\right].

Occasionally we need the “dilated” Bessel kernel defined, for t>0t>0, by

𝒥s(x,t)=tdGs(tx).\mathcal{J}_{s}(x,t)=t^{d}G_{s}(tx).

For 0<s<10<s<1, 1p<1\leq p<\infty and 1q1\leq q\leq\infty, we define the Besov spaces Bp,qsB^{s}_{p,q} as those fLpf\in L^{p} for which the seminorm

|f|Bp,qs:=(01(tsω(f,t)p)qdtt)1/q<.|f|_{B^{s}_{p,q}}:=\left(\int_{0}^{1}(t^{-s}\omega(f,t)_{p})^{q}\frac{dt}{t}\right)^{1/q}<\infty.

Equipped with the norm fBp,qs=fp+|f|Bp,qs\|f\|_{B^{s}_{p,q}}=\|f\|_{p}+|f|_{B^{s}_{p,q}} this becomes a Banach space. Lip(s,p)\text{Lip}(s,p) is the set of LpL^{p} functions which satisfy ω(f,t)p=𝒪(ts)\omega(f,t)_{p}=\mathcal{O}(t^{s}). For 0<s<10<s<1, Lip(s,p)=Bp,s\text{Lip}(s,p)=B^{s}_{p,\infty} while Lip(1,p)=Wp1\text{Lip}(1,p)=W^{1}_{p}.

For more on these function spaces we refer the reader to [13].

2 Order of Approximation

2.1 The LpL^{p} Case when 0<α10<\alpha\leq 1

Recall from the Introduction that Aα,μ(z)>0A_{\alpha,\mu}(z)>0 and Aα,μ(z)1=1\|A_{\alpha,\mu}(z)\|_{1}=1. Thus

Eα,μf(x)=f(x)dAα,μ(xy)f(y)𝑑y=dAα,μ(xy)(f(x)f(y))𝑑y.E_{\alpha,\mu}f(x)=f(x)-\int_{\mathbb{R}^{d}}A_{\alpha,\mu}(x-y)f(y)\,dy=\int_{\mathbb{R}^{d}}A_{\alpha,\mu}(x-y)(f(x)-f(y))\,dy.

Minkowski’s inequality and a change of variables readily show

Eα,μfpdAα,1(y)f()f(y/μ)pdydAα,1(y)ω(f,|y|/μ)pdy.\|E_{\alpha,\mu}f\|_{p}\leq\int_{\mathbb{R}^{d}}A_{\alpha,1}(y)\|f(\cdot)-f(\cdot-y/\mu)\|_{p}\,dy\leq\int_{\mathbb{R}^{d}}A_{\alpha,1}(y)\omega(f,|y|/\mu)_{p}\,dy.

To simplify the notation we set Aα(z):=Aα,1(z)=j=1aα,jG2j(z)A_{\alpha}(z):=A_{\alpha,1}(z)=\sum_{j=1}^{\infty}a_{\alpha,j}G_{2j}(z). Some properties of aα,ja_{\alpha,j} and G2j(z)G_{2j}(z) are summarized next.

Lemma 2.1.
  1. (a)

    G2j(z)G_{2j}(z) is positive, radial and decreasing with G2jL1=1\|G_{2j}\|_{L^{1}}=1. Moreover, G2j(z)=12d+2j22πd2Γ(j)Kd2j2(|z|)|z|2jd2\displaystyle G_{2j}(z)=\frac{1}{2^{\frac{d+2j-2}{2}}\pi^{\frac{d}{2}}\Gamma\left(j\right)}K_{\frac{d-2j}{2}}(|z|)|z|^{\frac{2j-d}{2}}, where Kν(t)K_{\nu}(t) is the modified Bessel function of the third kind.

  2. (b)

    We have aα,j>0a_{\alpha,j}>0 and aα,jCαj1α2a_{\alpha,j}\leq C_{\alpha}j^{-1-\frac{\alpha}{2}} with j=1aα,j=1\sum_{j=1}^{\infty}a_{\alpha,j}=1.

  3. (c)

    dG2j(y)|y|s𝑑y=Cd,sΓ(j+s2)Γ(j)\displaystyle\int_{\mathbb{R}^{d}}G_{2j}(y)|y|^{s}\,dy=C_{d,s}\frac{\Gamma\left(j+\frac{s}{2}\right)}{\Gamma(j)} for some constant Cd,sC_{d,s}. In addition, for 0s<α0\leq s<\alpha, dAα(y)|y|s𝑑y\displaystyle\int_{\mathbb{R}^{d}}A_{\alpha}(y)|y|^{s}\,dy converges.

Proof.
  1. (a)

    These are proved in Aronszajn–Smith [2, pgs. 413–421].

  2. (b)

    We use the binomial expansion (1t)α/2=1j=1aα,jtj\left(1-t\right)^{\alpha/2}=1-\sum_{j=1}^{\infty}a_{\alpha,j}t^{j}, where

    aα,j=|(α/2j)|=1Γ(α2)j1+α2(1+o(1))Cαj1α2.a_{\alpha,j}=\left|\binom{\alpha/2}{j}\right|=\frac{1}{\Gamma(-\frac{\alpha}{2})j^{1+\frac{\alpha}{2}}}\left(1+o(1)\right)\leq C_{\alpha}j^{-1-\frac{\alpha}{2}}.

    It follows that j=1aα,jtj\sum_{j=1}^{\infty}a_{\alpha,j}t^{j} converges absolutely for |t|1\left|t\right|\leq 1. Plugging t=1t=1 into (1t)α/2\left(1-t\right)^{\alpha/2} shows that j=1aα,j=1\sum_{j=1}^{\infty}a_{\alpha,j}=1.

  3. (c)

    For |Re(ν)|<Re(β)|\text{Re}(\nu)|<\text{Re}(\beta), we use the formula [7, Eq. 10.43.19]:

    0tβ1Kν(t)𝑑t=2β2Γ(β+ν2)Γ(βν2).\int_{0}^{\infty}t^{\beta-1}K_{\nu}(t)\,dt=2^{\beta-2}\Gamma\left(\frac{\beta+\nu}{2}\right)\Gamma\left(\frac{\beta-\nu}{2}\right). (11)

    A switch to spherical coordinates combined with part (a) and (11) gives

    dG2j(y)|y|s𝑑y=22jd2Γ(d2)Γ(j)0tj+d2+s1Kjd2(t)𝑑t=2sΓ(d+s2)Γ(d2)Γ(j+s2)Γ(j).\int_{\mathbb{R}^{d}}G_{2j}(y)|y|^{s}\,dy=\frac{2^{2-j-\frac{d}{2}}}{\Gamma(\frac{d}{2})\Gamma(j)}\int_{0}^{\infty}t^{j+\frac{d}{2}+s-1}K_{j-\frac{d}{2}}(t)\,dt=\frac{2^{s}\Gamma(\frac{d+s}{2})}{\Gamma(\frac{d}{2})}\cdot\frac{\Gamma\left(j+\frac{s}{2}\right)}{\Gamma(j)}.

    Since Γ(x+a)Γ(x)xa\Gamma(x+a)\sim\Gamma(x)x^{a} for large xx, and ajCj1α/2a_{j}\leq Cj^{-1-\alpha/2} we have

    dAα(y)|y|s𝑑y=j=1aα,jdG2j(y)|y|s𝑑y=Cs,dj=1aα,jΓ(j+s2)Γ(j)Cj=1j(2+αs)2,\int_{\mathbb{R}^{d}}A_{\alpha}(y)|y|^{s}\,dy=\sum_{j=1}^{\infty}a_{\alpha,j}\int_{\mathbb{R}^{d}}G_{2j}(y)|y|^{s}\,dy=C_{s,d}\sum_{j=1}^{\infty}a_{\alpha,j}\frac{\Gamma\left(j+\frac{s}{2}\right)}{\Gamma(j)}\leq C\sum_{j=1}^{\infty}j^{-\frac{(2+\alpha-s)}{2}},

    which converges when 0s<α0\leq s<\alpha.

We warm–up with a computation in the case α=1\alpha=1. Set B(ξ,μ)=ln(1+|ξ/μ|2)B(\xi,\mu)=\ln(1+|\xi/\mu|^{2}). The inequality 1etmin{1,t}1-e^{-t}\leq\min\{1,t\} for t>0t>0 shows

m1,μ(ξ)=a1,j(1(1+|ξ/μ|2)j)=a1,j(1ejB)a1,jmin{1,jB},\displaystyle m_{1,\mu}(\xi)=\sum a_{1,j}\left(1-(1+|\xi/\mu|^{2})^{-j}\right)=\sum a_{1,j}\left(1-e^{-jB}\right)\leq\sum a_{1,j}\min\{1,jB\},

and split the resulting sum to see that

m1,μ(ξ)jB1a1,jjB+jB1a1,j.m_{1,\mu}(\xi)\leq\sum_{j\leq B^{-1}}a_{1,j}jB+\sum_{j\geq B^{-1}}a_{1,j}.

The bound for a1,ja_{1,j} in Lemma 2.1(b) and comparison with an integral now yields

m1,μ(ξ)\displaystyle m_{1,\mu}(\xi) B01B1t𝑑t+1B1t32𝑑t2B(ξ,μ).\displaystyle\leq B\int_{0}^{\frac{1}{B}}\frac{1}{\sqrt{t}}\,dt+\int_{\frac{1}{B}}^{\infty}\frac{1}{t^{\frac{3}{2}}}\,dt\leq 2\sqrt{B(\xi,\mu)}.

This suggests a Kolmogorov–Seliverstov–Plessner type result.

Proposition 2.2.

If ln(1+|ξ|2)f^(ξ)2<\|\sqrt{\ln(1+|\xi|^{2})}\widehat{f}(\xi)\|_{2}<\infty, then E1,μf2=𝒪((lnμ)12)\|E_{1,\mu}f\|_{{}_{2}}=\mathcal{O}((\ln\mu)^{-\frac{1}{2}}) as μ\mu\to\infty.

Proof.

By Plancherel’s formula

E1,μf22|ξ|2(|ξ|2+μ2)ln(1+|ξ|2)ln(1+|ξ|2)|f^(ξ)|2𝑑ξ.\displaystyle\|E_{1,\mu}f\|^{2}_{{}_{2}}\leq\int\frac{|\xi|^{2}}{(|\xi|^{2}+\mu^{2})\ln(1+|\xi|^{2})}\ln(1+|\xi|^{2})|\widehat{f}(\xi)|^{2}\,d\xi.

r2/((r2+μ2)ln(1+r2)){r^{2}}/((r^{2}+\mu^{2})\ln(1+r^{2})) attains its maximum when rμr\approx\mu completing the proof. ∎

Our next result extends this exercise and contains Theorem 1.1 (b) and (c).

Theorem 2.3.

Assume 1p<1\leq p<\infty.

  1. (i)

    If α=1\alpha=1, there is a C>0C>0 depending only on dd such that

    E1,μfpCω(f,1/μ)p(3+2ln(fp2ω(f,1/μ)p)).\|E_{1,\mu}f\|_{p}\leq C\omega(f,1/\mu)_{p}\left(3+2\ln\left(\frac{||f||_{p}}{2\omega(f,1/\mu)_{p}}\right)\right).
  2. (ii)

    If 0<α<10<\alpha<1, there is a constant CC depending only on dd and α\alpha such that

    Eα,μfpC(ω(f,1/μ)p+(ω(f,1/μ)p)αfp1α)\|E_{\alpha,\mu}f\|_{p}\leq C\left(\omega(f,1/\mu)_{p}+(\omega(f,1/\mu)_{p})^{\alpha}\cdot||f||_{p}^{1-\alpha}\right)
Proof.

Recall that Aα(z)=j=1aα,jG2j(z)\displaystyle A_{\alpha}(z)=\sum_{j=1}^{\infty}a_{\alpha,j}G_{2j}(z) and that Eα,μfpAα(y)ω(f,|y|/μ)p𝑑y\displaystyle\|E_{\alpha,\mu}f\|_{p}\leq\int A_{\alpha}(y)\omega(f,|y|/\mu)_{p}\,dy. Let R>0R>0 be a large number to be chosen shortly. It follows that

Eα,μfp\displaystyle\|E_{\alpha,\mu}f\|_{p} j=1aα,jdω(f,|y|/μ)pG2j(y)𝑑y\displaystyle\leq\sum_{j=1}^{\infty}a_{\alpha,j}\int_{\mathbb{R}^{d}}\omega\left(f,|y|/\mu\right)_{p}G_{2j}(y)\,dy
jRaα,jdω(f,|y|/μ)pG2j(y)𝑑y+j>Raα,jdω(f,|y|/μ)pG2j(y)𝑑y.\displaystyle\leq\sum_{j\leq R}a_{\alpha,j}\int_{\mathbb{R}^{d}}\omega\left(f,|y|/\mu\right)_{p}G_{2j}(y)\,dy+\sum_{j>R}a_{\alpha,j}\int_{\mathbb{R}^{d}}\omega\left(f,|y|/\mu\right)_{p}G_{2j}(y)\,dy.

For the first sum we use the inequality ω(f,γt)p(1+|γ|)ω(f,t)p\omega(f,\gamma t)_{p}\leq(1+|\gamma|)\omega(f,t)_{p}. In the second sum we use ω(f,t)p2fp\omega(f,t)_{p}\leq 2||f||_{p}. Altogether

Eα,μfpω(f,1/μ)pjRaα,jd(1+|y|)G2j(y)𝑑y+2fpj>Raα,jdG2j(y)𝑑y.\|E_{\alpha,\mu}f\|_{p}\leq\omega\left(f,1/\mu\right)_{p}\sum_{j\leq R}a_{\alpha,j}\int_{\mathbb{R}^{d}}(1+|y|)G_{2j}(y)\,dy+2\|f\|_{p}\sum_{j>R}a_{\alpha,j}\int_{\mathbb{R}^{d}}G_{2j}(y)\,dy.

By Lemma 2.1(c),

Eα,μfpcdω(f,1/μ)pjRaα,j(1+j12)+2fpjRaα,j.\|E_{\alpha,\mu}f\|_{p}\leq c_{d}\omega\left(f,1/\mu\right)_{p}\sum_{j\leq R}a_{\alpha,j}(1+j^{\frac{1}{2}})+2\|f\|_{p}\sum_{j\geq R}a_{\alpha,j}.

We can now split the argument into the two cases.

  1. (i)

    The case α=1\alpha=1.

    We know that a1,jcj3/2a_{1,j}\leq cj^{-3/2} from Lemma 2.1(b) and can compare sums to integrals to deduce

    E1,μfpcd(ω(f,1/μ)p(1+lnR)+2fpR12).\|E_{1,\mu}f\|_{p}\leq c_{d}\left(\omega\left(f,1/\mu\right)_{p}(1+\ln R)+2\|f\|_{p}R^{-\frac{1}{2}}\right). (12)

    The choice R=(fp/2ω(f,1/μ)p)2R=\left(||f||_{p}/2\omega(f,1/\mu)_{p}\right)^{2} minimizes (12) and completes the proof in this case.

  2. (ii)

    The case 0<α<10<\alpha<1.

    Here aα,jcαj1α2a_{\alpha,j}\leq c_{\alpha}j^{-1-\frac{\alpha}{2}} and this time the integral test yields

    Eα,μfpcα,d(ω(f,1/μ)p(1+R12α2)+fpRα2).\|E_{\alpha,\mu}f\|_{p}\leq c_{\alpha,d}\left(\omega\left(f,1/\mu\right)_{p}(1+R^{\frac{1}{2}-\frac{\alpha}{2}})+\|f\|_{p}R^{-\frac{\alpha}{2}}\right). (13)

    This is minimized by the choice R=(αfp(1α)ω(f,1/μ)p)2R=\left(\dfrac{\alpha||f||_{p}}{(1-\alpha)\omega(f,1/\mu)_{p}}\right)^{2}.

2.2 Proof of Theorem 1.1 (a)

In this and the next two sections, we take α=1\alpha=1 and to simplify the notation we set Eμ:=E1,μE_{\mu}:=E_{1,\mu}, Tμ:=T1,μT_{\mu}:=T_{1,\mu} and A(x):=A1(x)A(x):=A_{1}(x). To motivate the proof, let us quickly estimate the order of approximation for functions in Besov spaces. Assume first that fBp,s:=Lip(s,p)f\in B^{s}_{p,\infty}:=\text{Lip}(s,p) with 0<s<10<s<1 and 1p<1\leq p<\infty. Recall that functions in Lip(s,p)\text{Lip}(s,p) satisfy ω(f,t)p=𝒪(ts)\omega(f,t)_{p}=\mathcal{O}(t^{s}). Arguing as in the above proof, but using Lemma 2.1(c)

EμfpA(y)f()f(y/μ)pdy\displaystyle\|E_{\mu}f\|_{p}\leq\int A(y)\|f(\cdot)-f(\cdot-y/\mu)\|_{p}\,dy μs|f|Bp,sA(y)|y|s𝑑y\displaystyle\leq\mu^{-s}|f|_{B^{s}_{p,\infty}}\int A(y)|y|^{s}\,dy
μsCd,s|f|Bp,s.\displaystyle\leq\mu^{-s}C_{d,s}|f|_{B^{s}_{p,\infty}}.

As the embedding Bp,qsBp,sB^{s}_{p,q}\hookrightarrow B^{s}_{p,\infty} holds, the same estimate applies to fBp,qsf\in B^{s}_{p,q}.

To handle the case s=1s=1 we restrict to 1<p<1<p<\infty. Note that ω(f,t)p=O(t)\omega(f,t)_{p}=O(t) implies fp1=Wp1f\in\mathcal{L}_{p}^{1}=W^{1}_{p} (see [13, pgs. 135–139]). Thus, for some gLpg\in L^{p}, (1+|ξ|2)1/2f^(ξ)=g^(ξ)(1+|\xi|^{2})^{1/2}\widehat{f}(\xi)=\widehat{g}(\xi) and Eμf=μ1E1(𝒥1(,μ)g)E_{\mu}f=\mu^{-1}E_{1}(\mathcal{J}_{1}(\cdot,\mu)\star g) where 𝒥s(x,t)=tdGs(tx)\mathcal{J}_{s}(x,t)=t^{d}G_{s}(tx) is the dilated Bessel kernel defined in the Notation section. From this

Eμfp=μ1E1(𝒥1(,μ)g))pCμ1gpCμ1fWp1(d).\|E_{\mu}f\|_{p}=\mu^{-1}\|E_{1}(\mathcal{J}_{1}(\cdot,\mu)\star g))\|_{p}\leq C\mu^{-1}\|g\|_{p}\leq C\mu^{-1}\|f\|_{W^{1}_{p}(\mathbb{R}^{d})}.

The first inequality follows from the LpL^{p} boundedness of both E1E_{1} and convolution with 𝒥s(x,t)\mathcal{J}_{s}(x,t). This argument combined with Theorem 2.3 imply results for function in various spaces. For simplicity, we state those that apply to functions in the Lipschitz classes Lip(s,p)\text{Lip}(s,p).

Corollary 2.4.

If fLip(s,p)f\in\text{Lip}(s,p) then

EμfpC{μs,p>1,0<s1;μs,p=1,0<s<1;μ1ln(μ),p=1,s=1.\|E_{\mu}f\|_{p}\leq C\begin{cases}\mu^{-s},&p>1,\quad 0<s\leq 1;\\ \mu^{-s},&p=1,\quad 0<s<1;\\ \mu^{-1}\ln(\mu),&p=1,\quad s=1.\end{cases}

Corollary 2.4 improves on Theorem 2.3, but the sharp result is Theorem 1.1 as it asserts the equivalence Eμfpω(f,1/μ)p\|E_{\mu}f\|_{p}\approx\omega(f,1/\mu)_{p} for 1<p<1<p<\infty. The idea is to use KK–functionls as an intermediary. We establish the equivalence between the order of approximation and a KK–functional. Known relationships between KK–functionals and the modulus of continuity allow us to complete the proof.

As in Ditzian–Ivanov [6], we introduce the KK–functional

K(t,f,|D|)p:=infgLp|D|gLp(fgp+t|D|gp).K(t,f,|D|)_{p}:=\inf_{\begin{subarray}{c}g\in L^{p}\\ |D|g\in L^{p}\end{subarray}}\left(\|f-g\|_{p}+t\||D|g\|_{p}\right). (14)

Here is the main step in proving the equivalence theorem.

Lemma 2.5.

K(1/μ,f,|D|)pEμfpK(1/\mu,f,|D|)_{p}\approx\|E_{\mu}f\|_{p}, for 1<p<1<p<\infty.

Proof.

If both g,|D|gLpg,|D|g\in L^{p},

Eμgp=1μ𝒥1(,μ)|D|gp1μ|D|gpμ1|D|gp.\|E_{\mu}g\|_{p}=\frac{1}{\mu}\|\mathcal{J}_{1}(\cdot,\mu)\star|D|g\|_{p}\leq\frac{1}{\mu}\||D|g\|_{p}\leq\mu^{-1}\||D|g\|_{p}.

This in turn implies

EμfpEμ(fg)p+Eμgpfgp+μ1|D|gp.\|E_{\mu}f\|_{p}\leq\|E_{\mu}(f-g)\|_{p}+\|E_{\mu}g\|_{p}\leq\|f-g\|_{p}+\mu^{-1}\||D|g\|_{p}.

Taking the infimum over such gg gives EμfpK(1/μ,f,|D|)p\|E_{\mu}f\|_{p}\leq K(1/\mu,f,|D|)_{p} which is one direction of the result.

We turn to the opposite inequality. Set g=Tμfg=T_{\mu}f. We only need show that μ1|D|gp:=μ1|D|TμfpCfTμfp\mu^{-1}\||D|g\|_{p}:=\mu^{-1}\||D|T_{\mu}f\|_{p}\leq C\|f-T_{\mu}f\|_{p}. On the Fourier transform side

μ1|D|Tμf^(ξ)=|ξ|μ(1|ξ|(μ2+|ξ|2)12)f^(ξ)\displaystyle\mu^{-1}\widehat{|D|T_{\mu}f}(\xi)=\frac{|\xi|}{\mu}\left(1-\frac{|\xi|}{(\mu^{2}+|\xi|^{2})^{\frac{1}{2}}}\right)\widehat{f}(\xi) =μ((μ2+|ξ|2)12+|ξ|)|ξ|f^(ξ)(μ2+|ξ|2)12\displaystyle=\frac{\mu}{((\mu^{2}+|\xi|^{2})^{\frac{1}{2}}+|\xi|)}\cdot\frac{|\xi|\widehat{f}(\xi)}{(\mu^{2}+|\xi|^{2})^{\frac{1}{2}}}
:=r(ξ)Eμf^(ξ),\displaystyle:=r(\xi)\widehat{E_{\mu}f}(\xi),

and we only need show that r(ξ)r(\xi) defines a bounded operator on LpL^{p}. A direct computation shows that for ξ0\xi\neq 0

|rξk|=|μ((μ2+|ξ|2)12+|ξ|)2(ξk|ξ|+ξk(μ2+|ξ|2)12)|2|ξ|.\left|\frac{\partial r}{\partial\xi_{k}}\right|=\left|-\mu((\mu^{2}+|\xi|^{2})^{\frac{1}{2}}+|\xi|)^{-2}\cdot\left(\frac{\xi_{k}}{|\xi|}+\frac{\xi_{k}}{(\mu^{2}+|\xi|^{2})^{\frac{1}{2}}}\right)\right|\leq\frac{2}{|\xi|}.

For any multi-index γ\gamma, this can be extended to |γr(ξ)|Cγ|ξ||γ||\partial^{\gamma}r(\xi)|\leq C_{\gamma}|\xi|^{-|\gamma|}. The multiplier theorem (see [14]) shows that r(D)r(D) is LpL^{p} bounded for 1<p<1<p<\infty. Hence, for 1<p<1<p<\infty, we obtain μ1|D|TμfpCfTμfp\mu^{-1}\||D|T_{\mu}f\|_{p}\leq C\|f-T_{\mu}f\|_{p}. Thus

K(1/μ,f,|D|)pfTμfp+μ1|D|TμfpCEμfp,K(1/\mu,f,|D|)_{p}\leq\|f-T_{\mu}f\|_{p}+\mu^{-1}\||D|T_{\mu}f\|_{p}\leq C\|E_{\mu}f\|_{p},

concluding the proof. ∎

We need one additional result.

Lemma 2.6.

For 1<p<1<p<\infty, gWp1g\in W^{1}_{p} if and only if gg and |D|g|D|g are in LpL^{p}.

Proof.

Using the Riesz transforms RjR_{j}, we can write Djg=Rj(|D|g)D_{j}g=R_{j}(|D|g), and the boundedness of RjR_{j} implies that DjgLpD_{j}g\in L^{p} whenever |D|gLp|D|g\in L^{p} if 1<p<1<p<\infty. Note that RjR_{j} is unbounded in L1L^{1} and LL^{\infty} and so we cannot include the case p=1,p=1,\infty. For the converse, suppose that gWp1g\in W^{1}_{p}. Then g=G1hg=G_{1}\star h for some hLph\in L^{p}. By definition, |D|g^(ξ)=|ξ|(|ξ|2+1)12h^(ξ)\widehat{|D|g}(\xi)=|\xi|(|\xi|^{2}+1)^{-\frac{1}{2}}\widehat{h}(\xi), so that |D|g=E1h|D|g=E_{1}h. As E1E_{1} is LpL^{p} bounded, |D|gp<\||D|g\|_{p}<\infty. ∎

Proof of Theorem 1.1(a).

We apply the result of Johnen–Scherer, [9], on the equivalence of moduli of continuity and KK–functionals. If we define

K(t,f,Lp,Wp1)=infgWp1(fgp+tsup|γ|=1Dγgp),K(t,f,L^{p},W^{1}_{p})=\inf_{g\in W^{1}_{p}}\left(||f-g||_{p}+t\sup_{|\gamma|=1}||D^{\gamma}g||_{p}\right),

their result is that K(t,f,Lp,Wp1)ω(f,t)pK(t,f,L^{p},W^{1}_{p})\approx\omega(f,t)_{p} for 1p1\leq p\leq\infty. However, Lemma 2.6 shows that when gWp1g\in W^{1}_{p}, we have sup|γ|=1Dγgp|D|gp\sup_{|\gamma|=1}\|D^{\gamma}g\|_{p}\approx\||D|g\|_{p} for 1<p<1<p<\infty. This implies K(t,f,|D|)K(t,f,Lp,Wp1)K(t,f,|D|)\approx K(t,f,L^{p},W^{1}_{p}), and we have shown

EμfpK(1/μ,f,|D|)K(1/μ,f,Lp,Wp1)ω(f,1/μ)p.\|E_{\mu}f\|_{p}\approx K(1/\mu,f,|D|)\approx K(1/\mu,f,L^{p},W^{1}_{p})\approx\omega(f,1/\mu)_{p}.

2.3 Extension to HpH^{p} when α=1\alpha=1

We denote the real Hardy spaces by Hp(d)H^{p}(\mathbb{R}^{d}). For p>1p>1, they coincide with the LpL^{p} spaces. For 0<p10<p\leq 1, it is a normed space of distributions. We denote the norm by Hp\|\cdot\|_{H^{p}} in this section. A far more thorough exposition on these spaces can be found in [14, Chap. 3] and we only state the bare minimum required.

The analog of Theorem 1.1 for functions in Hardy spaces is the following.

Theorem 2.7.

For 0<p10<p\leq 1, we have EμfHpCpω(f,1/μ)Hp\displaystyle\|E_{\mu}f\|_{{H}^{p}}\leq C_{p}\omega(f,1/\mu)_{{H}^{p}}.

For ff in Hp(d)H^{p}(\mathbb{R}^{d}), ω(f,t)Hp:=sup|h|t||f(+h)f()||Hp\omega(f,t)_{{H}^{p}}:=\sup_{|h|\leq t}||f(\cdot+h)-f(\cdot)||_{H^{p}} is the HpH^{p} modulus of continuity. We require a result of Colzani [5, Theorem 4.1] in the proof and to state it we need some notation. Let ϕC(d)\phi\in C^{\infty}(\mathbb{R}^{d}) be supported in |ξ|2|\xi|\leq 2, with ϕ(ξ)=1\phi(\xi)=1 for |ξ|1|\xi|\leq 1. and define Φsf^(ξ)=ϕ(ξ/s)f^(ξ)\displaystyle\widehat{\Phi_{s}\star f}(\xi)=\phi\left(\xi/s\right)\widehat{f}(\xi).

Lemma 2.8.

fΦμfHpCω(f,1/μ)Hp||f-\Phi_{\mu}\star f||_{{H}^{p}}\leq C\omega(f,1/\mu)_{{H}^{p}},

We also repeatedly use the “multiplier theorem for Hardy spaces” [13, §7.4.9].

Lemma 2.9.

Let kk\in\mathbb{N}. Then m(ξ)m(\xi) is HpH^{p} bounded if mLm\in L^{\infty} and, for |β|k|\beta|\leq k and k>d/pk>d/p, satisfies

sup0<A<A2|β|dA<|ξ|2A|βm(ξ)|2𝑑ξB.\sup_{0<A<\infty}A^{2|\beta|-d}\int_{A<|\xi|\leq 2A}|\partial^{\beta}m(\xi)|^{2}\,d\xi\leq B.
Proof of Theorem 2.7.

EμE_{\mu} is Hp{H}^{p} bounded as the estimate |βmμ(ξ)|Cβ|ξ|β|\partial^{\beta}m_{\mu}(\xi)|\leq C_{\beta}|\xi|^{-\beta} and the pointwise boundedness mμ(ξ)m_{\mu}(\xi) are enough to verifiy the conditions of Lemma 2.9. From here we check that

EμfHp=Eμ(fΦμf+Φμf)Hp\displaystyle\|E_{\mu}f\|_{{H}^{p}}=\|E_{\mu}(f-\Phi_{\mu}\star f+\Phi_{\mu}\star f)\|_{{H}^{p}} CfΦμfHp+Eμ(Φμf)Hp\displaystyle\leq C\|f-\Phi_{\mu}\star f\|_{{H}^{p}}+\|E_{\mu}(\Phi_{\mu}\star f)\|_{{H}^{p}}
Cω(f,1/μ)Hp+Eμ(Φμf)Hp,\displaystyle\leq C\omega(f,1/\mu)_{{H}^{p}}+\|E_{\mu}(\Phi_{\mu}\star f)\|_{{H}^{p}},

where we used the HpH^{p} boundedness of EμE_{\mu} in the second line and Colzani’s result, Lemma 2.8, in the third line. Thus the proof reduces to showing that Eμ(Φμf)HpCω(f,1/μ)Hp\|E_{\mu}(\Phi_{\mu}\star f)\|_{{H}^{p}}\leq C\omega(f,1/\mu)_{{H}^{p}}.

To prove this, we modify another argument of Colzani [5, Theorem 5.1]. As in that paper, we can assume that f^\widehat{f} is supported in (R+¯)d={xk0; 1kd}(\overline{R_{+}})^{d}=\{x_{k}\geq 0;\,1\leq k\leq d\}. Let θSd1\theta\in S^{d-1} be a fixed unit vector with θ\theta in a proper conic subset of (R+)d(R_{+})^{d}. We can thus find an ϵ>0\epsilon>0, such that θ\theta^{\perp}, the plane through the origin orthogonal to θ\theta, lies outside the ϵ\epsilon conic neighborhood of (R+¯)d(\overline{R_{+}})^{d}. Let χ(ξ)\chi(\xi) be smooth, homogenous of degree 0, identically 1 on (R+¯)d(\overline{R_{+}})^{d} and vanish outside of the ϵ/2\epsilon/2 conic neighborhood of (R+¯)d(\overline{R_{+}})^{d}. We use Colzani’s trick and write

Eμ(Φμf)^(ξ)=χ(ξ)mμ(ξ)ϕ(ξ/μ)(eiξθμ1)1(eiξθμ1)f^(ξ)\widehat{E_{\mu}(\Phi_{\mu}\star f)}(\xi)=\chi(\xi)m_{\mu}(\xi)\phi\left(\xi/\mu\right)(e^{i\frac{\xi\cdot\theta}{\mu}}-1)^{-1}\cdot(e^{i\frac{\xi\cdot\theta}{\mu}}-1)\widehat{f}(\xi)

We exploit the homogeneity of χ\chi and define two auxilary functions

ψ(z):=|z|χ(z)ϕ(z)(|z|2+1)12(eizθ1),ψ~(z):=ψ(z/μ).\psi(z):=\frac{|z|\chi(z)\phi(z)}{(|z|^{2}+1)^{\frac{1}{2}}(e^{iz\cdot\theta}-1)},\quad\widetilde{\psi}(z):=\psi(z/\mu).

If ψ~(ξ)\widetilde{\psi}(\xi) were HpH^{p} bounded, we would have

Eμ(Φμf)Hp=ψ~(D)[f(+θ/μ)f()]HpCpf(+θ/μ)f()Hp,\|E_{\mu}(\Phi_{\mu}\star f)\|_{{H}^{p}}=\|\widetilde{\psi}(D)\left[f(\cdot+\theta/\mu)-f(\cdot)\right]\|_{{H}^{p}}\leq C_{p}\|f(\cdot+\theta/\mu)-f(\cdot)\|_{{H}^{p}},

which of course implies that Eμ(Φμf)HpCpω(f,1/μ)Hp\|E_{\mu}(\Phi_{\mu}\star f)\|_{{H}^{p}}\leq C_{p}\omega(f,1/\mu)_{{H}^{p}} by the definition of the HpH^{p} modulus of continuity. The proof thus reduces to checking that ψ~\widetilde{\psi} satisfies the conditions of Lemma 2.9. Those conditions are dilation invariant so it is enough to check them for ψ\psi. To show boundedness, we focus on the behavior near θ\theta^{\perp} where eizθ1=0e^{iz\cdot\theta}-1=0. Excluding the origin, χ\chi vanishes on θ\theta^{\perp} by construction. As limz0ψ(z)\lim_{z\to 0}\psi(z) exists, we see that ψ\psi is bounded. For the derivatives, we use the product rule

βψ(z)=β1++β4=β(ββ1β4)β1χβ2ϕβ3(1eizθ1)β4(|z|(1+|z|2)12).\partial^{\beta}\psi(z)=\sum_{\beta_{1}+\ldots+\beta_{4}=\beta}\binom{\beta}{\beta_{1}\cdots\beta_{4}}\partial^{\beta_{1}}\chi\partial^{\beta_{2}}\phi\partial^{\beta_{3}}\left(\frac{1}{e^{iz\cdot\theta}-1}\right)\partial^{\beta_{4}}\left(\frac{|z|}{(1+|z|^{2})^{\frac{1}{2}}}\right).

Straightforward estimates using the support of ϕ\phi, and the homogeneity of χ\chi, yield |βψ(z)|Cϵ,β|z|β|\partial^{\beta}\psi(z)|\leq C_{\epsilon,\beta}|z|^{-\beta}. Applying Lemma 2.9 completes the proof. ∎

3 Saturation Theorems for α=1\alpha=1

In this section we show there is a limit to how well T1,μT_{1,\mu} approximates functions in LpL^{p}. We also characterize the class of functions which achieve the optimal approximation rate. To see the main idea, note that

m1,μ(ξ)f^(ξ)=|ξ|(μ2+|ξ|2)12f^(ξ)=|ξ|f^(ξ)(|ξ|2+μ2)12.m_{1,\mu}(\xi)\widehat{f}(\xi)=\dfrac{\left|\xi\right|}{(\mu^{2}+\left|\xi\right|^{2})^{\frac{1}{2}}}\widehat{f}(\xi)=\frac{|\xi|\widehat{f}(\xi)}{(|\xi|^{2}+\mu^{2})^{\frac{1}{2}}}.

The numerator is bounded in LpL^{p} when f\nabla f is in LpL^{p} while the denominator is 𝒪(1/μ)\mathcal{O}(1/\mu). We thus expect that functions with a derivative in LpL^{p} should have approximation error that is 𝒪(1/μ)\mathcal{O}(1/\mu). Though not rigorous, it is not far from the truth. The arguments to follow dress this observation in functional analytic clothing. Parts are adapted from Butzer–Nessel [4]. Once again Eμ:=E1,μE_{\mu}:=E_{1,\mu}.

Theorem 3.1.

For 1p<1\leq p<\infty, Eμfp=o(1/μ)\|E_{\mu}f\|_{p}=o(1/\mu) implies f=0f=0 in LpL^{p}.

Proof.

The proof is broken into cases.

  1. (a)

    When p=1p=1, both f^(ξ)\widehat{f}(\xi) and Eμf^(ξ)\widehat{E_{\mu}f}(\xi) are continuous functions. By assumption, μ|Eμf^(ξ)|μEμf1\mu|\widehat{E_{\mu}f}(\xi)|\leq\mu\|E_{\mu}f\|_{1} so limμμ|Eμf^(ξ)|=0\lim_{\mu\to\infty}\mu|\widehat{E_{\mu}f}(\xi)|=0 holds pointwise. A calculus argument gives limμμEμf^(ξ)=|ξ|f^(ξ)\lim_{\mu\to\infty}\mu\widehat{E_{\mu}f}(\xi)=|\xi|\widehat{f}(\xi). This implies f^(ξ)=0\widehat{f}(\xi)=0 almost everywhere finishing the proof. Incidentally, we showed that limμμEμf=|D|f\lim_{\mu\to\infty}\mu E_{\mu}f=|D|f. This remark is used below.

  2. (b)

    For the case 1<p21<p\leq 2, we use the Hausdorff–Young inequality to arrive at μEμf^pμEμfp=o(1)\|\mu\widehat{E_{\mu}f}\|_{p^{\prime}}\leq\|\mu E_{\mu}f\|_{p}=o(1). If we define the sequence gn(ξ):=nEnf^(ξ)g_{n}(\xi):=n\widehat{E_{n}f}(\xi), we see that limgn=0\lim g_{n}=0 in LpL^{p^{\prime}} and some subsequence satisfies limgnk=0\lim g_{n_{k}}=0 almost everywhere. But the earlier calculus argument and the uniqueness of limits implies 0=limkgnk=limknkEnkf^=|ξ|f^(ξ)0=\lim_{k}g_{n_{k}}=\lim_{k}n_{k}\widehat{E_{n_{k}}f}=|\xi|\widehat{f}(\xi), and again f^(ξ)=0\widehat{f}(\xi)=0 almost everywhere.

  3. (c)

    For the case p>2p>2 we use a duality argument. As 1<p<21<p^{\prime}<2, we know that φWp1\varphi\in W^{1}_{p^{\prime}} implies |D|φLp|D|\varphi\in L^{p^{\prime}}. The functional on Wp1W^{1}_{p^{\prime}} defined by Lμ,f(φ):=μEμf(x)φ(x)𝑑xL_{\mu,f}(\varphi):=\int\mu E_{\mu}f(x)\varphi(x)\,dx is easily seen to be a bounded linear functional with norm o(1)o(1). Two application of the dominated convergence theorem shows that

    0=limμLμ,f(φ)=limμμEμfφ=limμfμEμφ=f|D|φ.0=\lim_{\mu\to\infty}L_{\mu,f}(\varphi)=\lim_{\mu\to\infty}\int\mu E_{\mu}f\varphi=\lim_{\mu\to\infty}\int f\mu E_{\mu}\varphi=\int f\cdot|D|\varphi.

    As a linear functional, ff vanishes on the dense subspace Wp1W^{1}_{p^{\prime}} so f=0f=0.

We now describe the functions which attain the optimal order Eμfp=𝒪(1/μ)\|E_{\mu}f\|_{p}=\mathcal{O}(1/\mu).

Theorem 3.2.

For fLpf\in L^{p}, Eμfp=𝒪(1/μ)\|E_{\mu}f\|_{p}=\mathcal{O}(1/\mu) holds if and only if

  1. (a)

    fWp1f\in W^{1}_{p} when 1<p<1<p<\infty.

  2. (b)

    |ξ|f^(ξ)=ν^(ξ)|\xi|\widehat{f}(\xi)=\widehat{\nu}(\xi), for some bounded measure ν\nu when p=1p=1.

Proof.
  1. (a)

    In view of Theorem 1.1, such a function satisfies ω(f,t)p=𝒪(t)\omega(f,t)_{p}=\mathcal{O}(t) for 1<p<1<p<\infty. This defines Wp1W^{1}_{p} when 1<p<1<p<\infty.

  2. (b)

    In proving Theorem 3.1, we showed that limμμEμf=|D|f\lim\limits_{\mu\to\infty}\mu E_{\mu}f=|D|f. If Eμf1=𝒪(1/μ)\|E_{\mu}f\|_{1}=\mathcal{O}(1/\mu), then dνn=nEnf(x)dxd\nu_{n}=nE_{n}f(x)\,dx is a bounded sequence of Radon measures. This sequence satisfies supn|νn|(d)<\sup_{n}|\nu_{n}|(\mathbb{R}^{d})<\infty. By weak compactness (see [8, pgs. 54–55]), we can extract a limit, call it ν\nu, also satisfying |ν|(d)<|\nu|(\mathbb{R}^{d})<\infty. As a result, |D|f=ν|D|f=\nu or |ξ|f^(ξ)=ν^(ξ)|\xi|\widehat{f}(\xi)=\widehat{\nu}(\xi) which is one direction of (b).

    If fL1f\in L^{1} and |D|f=ν|D|f=\nu, for some bounded measure ν\nu, then

    Eμf1=μ1(𝒥1(,μ)ν1μ1|ν|(d),\|E_{\mu}f\|_{1}=\mu^{-1}\|(\mathcal{J}_{1}(\cdot,\mu)\star\nu\|_{1}\leq\mu^{-1}|\nu|(\mathbb{R}^{d}),

    which proves the other half of (b).

4 Kernel Estimates and Localization

We prove Theorem 1.5 in this section. The proof is based on a Littlewood–Paley type argument as in Stein [14, pgs. 241-247]. We modify the standard argument by replacing the Hörmander–Mikhlin condition (9) with the condition (10):|βb(ξ)|Cβ|ξ|α|β|(μ2+|ξ|2)α2\,\,|\partial^{\beta}b(\xi)|\leq C_{\beta}|\xi|^{\alpha-|\beta|}(\mu^{2}+|\xi|^{2})^{-\frac{\alpha}{2}}.

Proof of Theorem 1.5.

Let 1=jϕ(2jξ)1=\sum_{j\in\mathbb{Z}}\phi(2^{-j}\xi) be a Littlewood–Paley partition of unity. Put

Bj(x)=eixξϕ(2jξ)b(ξ)𝑑ξB_{j}(x)=\int e^{ix\cdot\xi}\phi(2^{-j}\xi)b(\xi)\,d\xi

For any multi-indices β\beta and γ\gamma we see

|xβ(ix)γBj(x)|=|xβeixξξγϕ(2jξ)b(ξ)𝑑ξ||ξβ(ξγϕ(2jξ)b(ξ))|𝑑ξ.\left|x^{\beta}(-i\partial_{x})^{\gamma}B_{j}(x)\right|=\left|\int x^{\beta}e^{ix\xi}\xi^{\gamma}\phi(2^{-j}\xi)b(\xi)\,d\xi\right|\leq\int\left|\partial_{\xi}^{\beta}(\xi^{\gamma}\phi(2^{-j}\xi)b(\xi))\right|\,d\xi.

The product rule, (10), and direct integration gives

|xβ(ix)γBj(x)|Cγ,β,d2j(d+|γ||β|)2α(j1)(22(j1)+μ2)α/2,\left|x^{\beta}(-i\partial_{x})^{\gamma}B_{j}(x)\right|\leq C_{\gamma,\beta,d}2^{j(d+|\gamma|-|\beta|)}\frac{2^{\alpha(j-1)}}{(2^{2(j-1)}+\mu^{2})^{\alpha/2}},

which can be rearranged to the derivative estimate

|xγBj(x)|Cγ,M,d2j(d+|γ|M)2α(j1)(22(j1)+μ2)α/2|x|M.|\partial_{x}^{\gamma}B_{j}(x)|\leq C_{\gamma,M,d}2^{j(d+|\gamma|-M)}\frac{2^{\alpha(j-1)}}{(2^{2(j-1)}+\mu^{2})^{\alpha/2}}|x|^{-M}. (15)

We now split the sum as

xγB(x)=2j1|x|1xγBj(x)+2j1>|x|1xγBj(x).\partial_{x}^{\gamma}B(x)=\sum_{2^{j-1}\leq|x|^{-1}}\partial_{x}^{\gamma}B_{j}(x)+\sum_{2^{j-1}>|x|^{-1}}\partial_{x}^{\gamma}B_{j}(x).

To estimate the first sum, set M=0M=0 in (15) to find that

2j1|x|1|xγBj(x)|Cγ,d2j1|x|12j(d+|γ|)(1+(μ/2j1)2)α2.\sum_{2^{j-1}\leq|x|^{-1}}|\partial_{x}^{\gamma}B_{j}(x)|\leq C_{\gamma,d}\sum_{2^{j-1}\leq|x|^{-1}}\frac{2^{j(d+|\gamma|)}}{(1+\left(\mu/2^{j-1}\right)^{2})^{\frac{\alpha}{2}}}. (16)

When 2j1|x|12^{j-1}\leq|x|^{-1}, we see that (1+(μ/2j1)2)α2(1+|μx|2)α2(1+(\mu/2^{j-1})^{2})^{-\frac{\alpha}{2}}\leq(1+|\mu x|^{2})^{-\frac{\alpha}{2}} and summing the geometric series (16) we obtain

2j1|x|1|xγBj(x)|Cγ,d(1+|μx|2)α21|x|d+|γ|.\sum_{2^{j-1}\leq|x|^{-1}}|\partial_{x}^{\gamma}B_{j}(x)|\leq\frac{C_{\gamma,d}}{(1+|\mu x|^{2})^{\frac{\alpha}{2}}}\frac{1}{|x|^{d+|\gamma|}}. (17)

For the second sum, we set MM to be the smallest integer greater than |γ|+d+1/2|\gamma|+d+1/2 and arrive at

2j1>|x|1|xγBj(x)|\displaystyle\sum_{2^{j-1}>|x|^{-1}}|\partial_{x}^{\gamma}B_{j}(x)| Cγ,d,M|x|M2j1>|x|12j(d+|γ|M)2α(j1)(μ2j1(μ2j1+2j1μ))α2\displaystyle\leq C_{\gamma,d,M}|x|^{-M}\sum_{2^{j-1}>|x|^{-1}}2^{j(d+|\gamma|-M)}\frac{2^{\alpha(j-1)}}{\left(\mu 2^{j-1}\left(\frac{\mu}{2^{j-1}}+\frac{2^{j-1}}{\mu}\right)\right)^{\frac{\alpha}{2}}}
Cγ,d,M|x|Mμα/22j1>|x|12j(d+|γ|+α2M)(μ2j1+2j1μ)α2.\displaystyle\leq C_{\gamma,d,M}|x|^{-M}\mu^{-\alpha/2}\sum_{2^{j-1}>|x|^{-1}}\frac{2^{j(d+|\gamma|+\frac{\alpha}{2}-M)}}{\left(\frac{\mu}{2^{j-1}}+\frac{2^{j-1}}{\mu}\right)^{\frac{\alpha}{2}}}. (18)

Setting t=μ/2j1t=\mu/2^{j-1} and L=|μx|L=|\mu x|. If 2j1>|x|12^{j-1}>|x|^{-1}, we see that 0<tL0<t\leq L. A direct calculation shows

sup0<tL(t+t1)α2={2α2;L>1,(L+L1)α2;L1.\sup\limits_{0<t\leq L}(t+t^{-1})^{-\frac{\alpha}{2}}=\begin{cases}2^{-\frac{\alpha}{2}};&L>1,\\ (L+L^{-1})^{-\frac{\alpha}{2}};&L\leq 1.\end{cases}

We use this to sum the geometric series in (4) and obtain

2j1>|x|1|xγBj(x)|C{|2μx|α2|x||γ|d;|μx|>1,(|μx|2+1)α2|x||γ|d;|μx|1.\sum_{2^{j-1}>|x|^{-1}}|\partial_{x}^{\gamma}B_{j}(x)|\leq C\begin{cases}|2\mu x|^{-\frac{\alpha}{2}}|x|^{-|\gamma|-d};&|\mu x|>1,\\ (|\mu x|^{2}+1)^{-\frac{\alpha}{2}}|x|^{-|\gamma|-d};&|\mu x|\leq 1.\end{cases} (19)

Combining (19) with the earlier estimate (17) completes the proof. ∎

A similar argument establishes a Hörmander condition which we include here for completeness and to contrast with the usual one.

Theorem 4.1.

If b(ξ)b(\xi) satisfies (10), then its kernel satisfies

|x|2|y||B(x+y)B(x)|𝑑xC{|2μy|1/2;|μy|>1,(|μy|2+1)1/2;|μy|1.\int\limits_{|x|\geq 2|y|}\left|B(x+y)-B(x)\right|\,dx\leq C\begin{cases}|2\mu y|^{-1/2};&|\mu y|>1,\\ (|\mu y|^{2}+1)^{-1/2};&|\mu y|\leq 1.\end{cases} (20)

We present a refinement of Corollary 1.6 from the Introduction.

Theorem 4.2.

Fix δ>0\delta>0 and suppose fLf\in L^{\infty} vanishes for |x|<δ|x|<\delta. If σ<δ\sigma<\delta, there is a constant C=C(d,σ,δ)C=C(d,\sigma,\delta) such that the “uniform maximal” estimate holds:

sup|x|σsupμ1|Eα,μf(x)|CfL.\sup_{|x|\leq\sigma}\sup_{\mu\geq 1}|E_{\alpha,\mu}f(x)|\leq C\|f\|_{{}_{L^{\infty}}}. (21)

The behavior of the constant in (21) is of interest here. Our proof gives a logarithmic dependence on the distance δσ\delta-\sigma which is independent of α\alpha

Proof.

Let Kα,μ(z)K_{\alpha,\mu}(z) be the associated kernel. For any |x|<σ|x|<\sigma it is enough to estimate

|y|δ|Kα,μ(xy)|𝑑y|xy|δσ|Kα,μ(xy)|𝑑y.\int_{|y|\geq\delta}|K_{\alpha,\mu}(x-y)|\,dy\leq\int_{|x-y|\geq\delta-\sigma}|K_{\alpha,\mu}(x-y)|\,dy.

Since we are taking a supremum in μ\mu, we cannot assume that 1/μ1/\mu is small compared to δσ\delta-\sigma. Instead we split the integral over two regions: |xy|1/μ|x-y|\geq 1/\mu and δσ|xy|1/μ\delta-\sigma\leq|x-y|\leq 1/\mu. By Theorem 1.5

|y|δ|Kα,μ(xy)|𝑑y|xy|1μμα2|xy|d+α2𝑑y+δσ|xy|1μ(1+(μ|xy|)2)α2|xy|d𝑑y.\int\limits_{|y|\geq\delta}|K_{\alpha,\mu}(x-y)|\,dy\leq\int\limits_{|x-y|\geq\frac{1}{\mu}}\frac{\mu^{-\frac{\alpha}{2}}}{|x-y|^{d+\frac{\alpha}{2}}}\,dy\,+\int\limits_{\delta-\sigma\leq|x-y|\leq\frac{1}{\mu}}\frac{(1+(\mu|x-y|)^{2})^{-\frac{\alpha}{2}}}{|x-y|^{d}}\,dy.

After the integration dust settles, we see that

supμ1|y|δ|Kα,μ(xy)|𝑑yCsupμ1(1+|lnμ(δσ)|(1+(μ(δσ))2)α2)C(1+|ln(δσ)|).\sup_{\mu\geq 1}\int\limits_{|y|\geq\delta}|K_{\alpha,\mu}(x-y)|\,dy\leq C\sup_{\mu\geq 1}\left(1+\frac{|\ln\mu(\delta-\sigma)|}{(1+(\mu(\delta-\sigma))^{2})^{\frac{\alpha}{2}}}\right)\leq C(1+|\ln(\delta-\sigma)|).

Acknowledgement

I would like to thank Professor L. Colzani for clarifying some points in [5]. This improved the argument in §2.2. Professor A. Larrain–Hubach also made helpful comments on an earlier draft. Any remaining errors are, of course, mine.

References

  • [1] D. Adams and L. Hedberg, Function Spaces and Potential Theory, Springer Berlin Heidelberg, 1999.
  • [2] N. Aronszajn and K.T Smith, Theory of Bessel Potentials. I, Annales de l’institut Fourier 11 (1961), 385–475.
  • [3] O. V. Besov and S. B. Stechkin, Description of Moduli of Continuity in L2L_{2}, Proceedings of the Steklov Institute of Mathematics 134 (1975).
  • [4] P. L. Butzer and R. J. Nessel, Contributions To The Theory Of Saturation For Singular Integrals In Several Variables. I: General Theory, Indagationes Mathematicae (Proceedings) 69 (1966), 513–531.
  • [5] L. Colzani, Jackson Theorems in Hardy Spaces and Approximation by Riesz Means, Journal of Approximation Theory 49 (1987), 240–251.
  • [6] Z. Ditzian and K.G Ivanov, Strong Converse Inequalities, Journal d’Analyse Mathématique 62 (1993), 61–111.
  • [7] NIST Digital Library of Mathematical Functions, http://dlmf.nist.gov/, Release 1.1.8 of 2022-12-15, F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
  • [8] L. C. Evans and R. F. Gariepy, Measure theory and fine property of functions, CRC Press, Florida, USA, 1992.
  • [9] H. Johnen and K. Scherer, On the equivalence of the KK-functional and moduli of continuity and some applications, Constructive Theory of Functions of Several Variables (W. Schempp and K. Zeller, eds.), Springer Berlin Heidelberg, 1977, pp. 119–140.
  • [10] Z. Liu and S. Lu, Applications of Hörmander multiplier theorem to approximation in real Hardy spaces, Harmonic Analysis (MT. Cheng, DG. Deng, and XW. Zhou, eds.), Lecture Notes in Mathematics, Springer, Berlin, 1991, pp. 119–129.
  • [11] B. Muckenhoupt and R. Wheeden, Weighted Norm Inequalities for Fractional Integrals, Trans. Amer. Math. Soc. 192 (1974), 261–274.
  • [12] M. Schecter, The Spectrum of the Schrodinger Operator, Trans. Amer. Math. Soc. 312 (1989), no. 1, 115–128.
  • [13] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, 1970.
  • [14]  , Harmonic Analysis: Real-Variable Methods, Orthogonality and Oscillatory integrals, Princeton University Press, 1993.
  • [15] W. P. Ziemer, Weakly Differentiable Functions, Springer Berlin Heidelberg, 1989.