TY - JOUR AU - AB - Abstract We provide an improvement of the Hörmander multiplier theorem in which the Sobolev space |${{L^{r}_{s}}}(\mathbb R^{n})$| with integrability index r and smoothness index s > n/r is replaced by the Sobolev space with smoothness s built upon the Lorentz space |$L^{n/s,1}(\mathbb R^{n})$|⁠. 1 Introduction Given a bounded function σ on |$\mathbb R^{n}$|⁠, we define a linear operator \begin{align*} T_{\sigma}(f)(x) = \int_{\mathbb R^{n}} \widehat{f}(\xi) \sigma(\xi) e^{2\pi i x\cdot \xi}\ \text{d}\xi \end{align*} acting on Schwartz functions f on |$\mathbb R^{n}$|⁠; here |$\widehat {f}(\xi ) = \int _{\mathbb R^{n}} f(x) e^{-2\pi i x\cdot \xi }\ \text {d}x$| is the Fourier transform of f. An old problem in harmonic analysis is to find optimal sufficient conditions on σ to be an Lp Fourier multiplier, that is, for the operator Tσ to admit a bounded extension from |$L^{p}(\mathbb R^{n})$| to itself for a given |$p\in (1,\infty )$|⁠. Mikhlin’s [13] classical multiplier theorem states that if the condition \begin{align} |\partial^{\alpha} \sigma(\xi)|\leq C_{\alpha} |\xi|^{-| \alpha|}, \qquad \xi\neq 0, \end{align} (1.1) holds for all multi-indices α with size |α|≤ [n/2] + 1, then Tσ admits a bounded extension from |$L^{p}(\mathbb R^{n})$| to itself for all |$1 0 let (I−Δ)s/2 denote the operator given on the Fourier transform by multiplication by (1 + 4π2|ξ|2)s/2 and let Ψ be a Schwartz function whose Fourier transform is supported in the annulus {ξ : 1/2 < |ξ| < 2} and which satisfies |$\sum _{j\in \mathbb Z} \widehat {\Psi }(2^{-j}\xi )=1$| for all |$\xi \neq 0$|⁠. If for some 1 ≤ r ≤ 2 and s > n/r, σ satisfies \begin{align} \sup_{k\in \mathbb Z} \left\|(I-\Delta)^{s/2} \left[ \widehat{\Psi}\sigma (2^{k} \cdot)\right] \right\|_{L^{r}(\mathbb R^{n}) }<\infty, \end{align} (1.2) then Tσ admits a bounded extension from |$L^{p}(\mathbb R^{n})$| to itself for all |$1 n, then the condition |$\left | \frac 1p -\frac 12 \right | <\frac sn$| is essentially optimal for assumption (1.2). Observe also that the condition rs > n is dictated by the embedding of |${{L^{r}_{s}}}(\mathbb R^{n}) \hookrightarrow L^{\infty }(\mathbb R^{n})$|⁠. It is still unknown to us if Lp boundedness holds on the line |$\left | \frac 1p -\frac 12 \right | =\frac sn$|⁠. Positive endpoint results on Lp and on H1 involving Besov spaces can be found in Seeger [16], [17], [18]. Unlike the Mikhlin multiplier theorem, the Hörmander and Calderón–Torchinsky theorems can treat multipliers whose derivatives have infinitely many singularities, such as the multiplier \begin{align} \sigma(x)=\sum_{k\in \mathbb Z} \phi(2^{-k}x) |2^{-k}x-a_{k}|^{\beta}, \end{align} (1.3) where β > 0, ϕ is a smooth function supported in the set |$\{x\in \mathbb R^{n}: \frac {1}{2}<|x|<2\}$| and, for every |$k\in \mathbb N$|⁠, |$a_{k}\in \mathbb R^{n}$| belongs to the same set. In this paper, we improve the result of [2, Theorem 4.6] by replacing the Lebesgue space |$L^{r}(\mathbb R^{n})$|⁠, |$r>\frac {n}{s}$|⁠, in condition (1.2) by the locally larger Lorentz space |$L^{\frac {n}{s},1}(\mathbb R^{n})$|⁠, defined in terms of the norm \begin{align*} \|f\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}=\int_{0}^{\infty} f^{\ast}(r)r^{\frac{s}{n}-1}\,\text{d}r. \end{align*} Here, f* stands for the nonincreasing rearrangement of the function f, namely, for the unique nonincreasing left-continuous function on |$(0,\infty )$| equimeasurable with f, given by the explicit expression \begin{align*} f^{\ast}(t)= \inf \left\{r\ge 0:\,\, |\{y\in \mathbb R^{n}:\,\, |f(y)|>r\}| < t \right\}\,. \end{align*} We point out that the Lorentz space |$L^{\frac {n}{s},1}(\mathbb R^{n})$| appears naturally in this context, since it is known to be, at least for integer values of s, locally the largest rearrangement-invariant function space such that membership of |$(I-\Delta )^{\frac {s}{2}}f$| to this space forces f to be bounded, see [20, 3]. Theorem 1.1. Let Ψ be a Schwartz function on |$\mathbb R^{n}$| whose Fourier transform is supported in the annulus 1/2 < |ξ| < 2 and satisfies |$\sum _{j\in \mathbb Z} \widehat {\Psi } ( 2^{-j} \xi ) =1$|⁠, |$\xi \neq 0$|⁠. Let |$p\in (1,\infty )$|⁠, |$n\in \mathbb N$|⁠, and let s ∈ (0, n) satisfy \begin{align*} \left|\frac{1}{p}-\frac{1}{2}\right|<\frac{s}{n}. \end{align*} Then for all functions f in the Schwartz class of |$\mathbb R^{n}$| we have the a priori estimate \begin{align} \|T_{\sigma} f\|_{L^{p}(\mathbb R^{n})} \leq C \sup_{j\in \mathbb Z} \left\|(I-\Delta)^{\frac{s}{2}}\left[\widehat\Psi \sigma(2^{j}\cdot)\right]\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \|f\|_{L^{p}(\mathbb R^{n})}. \end{align} (1.4) As an application of Theorem 1.1 we show that the function σ from (1.3) continues to be an Lp Fourier multiplier for any |$p\in (1,\infty )$| if |2−kx − ak| is replaced by |$(\log \frac{e4^n}{|2^{-k}x-a_k|^n})^{-1}$|⁠. In fact, we can even allow an arbitrary iteration of logarithms in this example. Example 1.2. Assume that |$n\in \mathbb N$|⁠, n ≥ 2, and β < 0. Let ϕ be a smooth function supported in the set |$A = \{x\in \mathbb R^{n}: 1/2<|x|<2\}$| and let ak ∈ A, |$k\in \mathbb Z$|⁠. Then the function \begin{align} \sigma(x)=\sum_{k\in \mathbb Z} \phi(2^{-k}x) \left(\log \frac{e4^{n}}{|2^{-k}x-a_{k}|^{n}}\right)^{\beta} \end{align} (1.5) is an Lp Fourier multiplier for any |$p\in (1,\infty )$|⁠. To verify the statement of Example 1.2, we fix a positive integer s0} f^{\ast}(t) t^{\frac{1}{p}}. \end{align*} It can be shown that \begin{align*} \|f\|_{L^{p,1}(\mathbb R^{n})}=p\int_{0}^{\infty} |\{x\in \mathbb R^{n}: |f(x)|>\lambda\}|^{\frac{1}{p}}\,\text{d}\lambda \end{align*} and \begin{align*} \|f\|_{L^{p,\infty}(\mathbb R^{n})}=\sup_{\lambda>0} \lambda |\{x\in \mathbb R^{n}: |f(x)|>\lambda\}|^{\frac{1}{p}}. \end{align*} The space |$L^{p^{\prime},\infty }(\mathbb R^{n})$|⁠, where |$p^{\prime}=\frac {p}{p-1}$|⁠, is a kind of a measure-theoretic dual of the space |$L^{p,1}(\mathbb R^{n})$|⁠, in the sense that the following form of Hölder’s inequality \begin{align*} \int_{\mathbb R^{n}} |fg| \leq \|f\|_{L^{p,1}(\mathbb R^{n})} \|g\|_{L^{p^{\prime},\infty}(\mathbb R^{n})} \end{align*} holds. In what follows, B(x, r) denotes the ball centered at point x and having the radius r. If a ball of radius r is centered at the origin, we shall denote it simply by Br. Let q ≥ 1 be a real number. We consider the centered maximal operator |$M_{L^{q}}$| defined by \begin{align*} M_{L^{q}} f(x) = \sup_{r>0} \left(\frac{1}{|B(x,r)|} \int_{B(x,r)} |f(y)|^{q}\,\text{d}y\right)^{\frac{1}{q}}. \end{align*} Observe that \begin{align*} M_{L^{q}} f=(M |f|^{q})^{\frac{1}{q}}, \end{align*} where M stands for the classical Hardy–Littlewood maximal operator. The crucial step toward proving Theorem 2.2 is the following lemma, which can be understood as a sharp variant of [8, Theorem 2.1.10]. Lemma 2.1. Assume that |$n\in \mathbb N$|⁠, s ∈ (0, n), and |$q>\frac {n}{s}$|⁠. Then there is a positive constant C depending on n, s, and q such that for any |$j\in \mathbb Z$| and any measurable function f on |$\mathbb R^{n}$|⁠, \begin{align} \left\|\frac{f(x+2^{-j}y)}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} \leq C M_{L^{q}} f(x), \quad x\in \mathbb R^{n}. \end{align} (2.7) Proof. We may assume, without loss of generality, that j = 0 and x = 0. Indeed, setting g(y) = f(x + 2−jy), we obtain \begin{align} \left\|\frac{f(x+2^{-j}y)}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} = \left\|\frac{g(y)}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} \end{align} (2.8) and \begin{align} M_{L^{q}} f(x) &=\sup_{r>0} \left(\frac{1}{|B(x,r)|} \int_{B(x,r)} |f(y)|^{q}\,\text{d}y\right)^{\frac{1}{q}}\\ \nonumber &=\sup_{r>0} \left(\frac{1}{2^{jn}|B(x,r)|} \int_{B(0,2^{j} r)} |f(x+2^{-j}z)|^{q}\,\text{d}z\right)^{\frac{1}{q}}\\ \nonumber &=\sup_{r^{\prime}>0} \left(\frac{1}{|B(0,r^{\prime})|} \int_{B(0,r^{\prime})} |g(y)|^{q}\,\text{d}y\right)^{\frac{1}{q}}\\ \nonumber &=M_{L^{q}} g(0). \end{align} (2.9) Hence, it suffices to show that for any measurable function g on |$\mathbb R^{n}$|⁠, \begin{align} \left\|\frac{g(y)}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} \leq C M_{L^{q}} g(0). \end{align} (2.10) If |$M_{L^{q}} g(0)=\infty $|⁠, then inequality (2.10) holds trivially, so we can assume in what follows that |$M_{L^{q}} g(0)<\infty $|⁠. Since the case |$M_{L^{q}} g(0)=0$| is trivial as well (as g needs to vanish a.e. in this case), dividing the function g by the positive constant |$M_{L^{q}} g(0)$|⁠, we can in fact assume that |$M_{L^{q}} g(0)=1$|⁠. Fix any a > 0 and |$k\in \mathbb N_{0}$|⁠. Then \begin{align*} |\{y\in B_{2^{k+1}}\setminus B_{2^{k}}: |g(y)|>a\}| &\leq \frac{1}{a^{q}}\int_{B_{2^{k+1}}\setminus B_{2^{k}}} |g(y)|^{q}\,\text{d}y\\[12pt] &\leq \frac{|B_{2^{k+1}}| }{a^{q}}\cdot \frac{1}{|B_{2^{k+1}}|} \int_{B_{2^{k+1}}} |g(y)|^{q}\,\text{d}y \leq \frac{\omega_{n} 2^{(k+1)n}}{a^{q}}, \end{align*} where ωn denotes the volume of the unit ball in |$\mathbb R^{n}$|⁠. Combining this with the trivial estimate \begin{align*} |\{y\in B_{2^{k+1}}\setminus B_{2^{k}}: |g(y)|>a\}| \leq \omega_{n} 2^{(k+1)n}, \end{align*} we deduce that \begin{align*} &\left|\left\{y\in \mathbb R^{n}: \frac{|g(y)|}{(1+|y|)^{s}}>a\right\}\right|\\[12pt] &\quad=\left|\left\{y\in B_{1}: \frac{|g(y)|}{(1+|y|)^{s}}>a\right\}\right| +\sum_{k=0}^{\infty} \left|\left\{y\in B_{2^{k+1}}\setminus B_{2^{k}}: \frac{|g(y)|}{(1+|y|)^{s}}>a\right\}\right|\\[12pt] &\quad\leq \left|\left\{y\in B_{1}: |g(y)|>a\right\}\right| + \sum_{k=0}^{\infty} \left|\left\{y\in B_{2^{k+1}}\setminus B_{2^{k}} : |g(y)|> 2^{ks} a\right\}\right|\\[12pt] &\quad\leq \left|\left\{y\in B_{1}: |g(y)|>a\right\}\right| + \sum_{k=0}^{\infty} \omega_{n} 2^{(k+1)n} \min\left\{\frac{1}{2^{ksq} a^{q}}, 1 \right\}\\[12pt] &\quad\leq \left|\left\{y\in B_{1}: |g(y)|>a\right\}\right| + \sum_{k\in \mathbb N_{0}: 2^{k}<\frac{1}{a^{{1}/{s}}}} \omega_{n} 2^{n} \cdot 2^{kn} + \sum_{k\in \mathbb N_{0}: 2^{k} \geq \frac{1}{a^{{1}/{s}}}} \frac{\omega_{n} 2^{n}}{a^{q}} \cdot 2^{k(n-sq)}\\[12pt] &\quad\leq \left|\left\{y\in B_{1}: |g(y)|>a\right\}\right| + \frac{C}{a^{\frac{n}{s}}}. \end{align*} Notice that in the last inequality we have used the fact that n − sq < 0. Hence, \begin{align*} \left\|\frac{g(y)}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} &=\sup_{a>0} a \left|\left\{y\in \mathbb R^{n}: \frac{|g(y)|}{(1+|y|)^{s}}>a\right\}\right|{}^{\frac{s}{n}}\\[12pt] &\leq \sup_{a>0} a \left|\left\{y\in B_{1}: |g(y)|>a\right\}\right|{}^{\frac{s}{n}} +C\\[12pt] &=\|g\|_{L^{\frac{n}{s},\infty}(B_{1})} +C\\[12pt] &\leq C^{\prime} \|g\|_{L^{q}(B_{1})}+C\\[12pt] &\leq C^{\prime} \omega_{n}^{\frac{1}{q}} M_{L^{q}}g(0) +C \leq C^{^{\prime\prime}}, \end{align*} where C′ > 0 is the constant from the embedding |$L^{q}(B_{1}) \hookrightarrow L^{\frac {n}{s},\infty }(B_{1})$|⁠. Since |$M_{L^{q}}g(0)=1$|⁠, this proves (2.10), and in turn (2.7) as well. Theorem 2.2. Let |$p\in (1,\infty )$|⁠, |$n\in \mathbb N$|⁠, |$s\in (\frac {n}{2},n)$|⁠. Let Ψ be as in Theorem 1.1. Then \begin{align} \|T_{\sigma} f\|_{L^{p}(\mathbb R^{n})} \leq C \sup_{j\in \mathbb Z} \left\|(I-\Delta)^{\frac{s}{2}}\left[\widehat\Psi \sigma(2^{j}\cdot)\right]\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \|f\|_{L^{p}(\mathbb R^{n})}. \end{align} (2.11) Proof. Let \begin{align*} K=\sup_{j\in \mathbb Z} \left\| (I-\Delta)^{\frac{s }{2}} \left[\widehat{\Psi} \sigma(2^{j}\cdot)\right]\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} <\infty\, . \end{align*} Introduce the function Θ satisfying \begin{align*} \widehat{\Theta}(\xi)=\widehat{\Psi}(\xi/2)+\widehat{\Psi}(\xi)+\widehat{\Psi}(2\xi), \end{align*} and observe that |$\widehat {\Theta }$| is equal to 1 on the support of the function |$\widehat {\Psi }$|⁠. Let us denote by Δj and |$\Delta _{j}^{\Theta }$| the Littlewood–Paley operators associated with Ψ and Θ, respectively. If f is a Schwartz function on |$\mathbb R^{n}$|⁠, then standard manipulations yield \begin{align*} \Delta_{j} T_{\sigma} (f)(x) &= \int_{\mathbb R^{n}} \widehat{f}(\xi) \widehat{\Psi}(2^{-j} \xi) \sigma(\xi) e^{2\pi i x\cdot \xi}\text{ d}\xi = \int_{\mathbb R^{n}} \left(\Delta_{j}^{\Theta}f \right){\widehat{\phantom{0}}}{}\,(\xi) \widehat{\Psi}(2^{-j} \xi) \sigma(\xi) e^{2\pi i x\cdot \xi} \text{d}\xi\\[6pt] &=2^{jn} \int_{\mathbb R^{n}} \left(\Delta_{j}^{\Theta}f \right)\widehat{\phantom{0}}\,(2^{j} \xi^{\prime}) \widehat{\Psi}( \xi^{\prime}) \sigma(2^{j}\xi^{\prime}) e^{2\pi i x\cdot 2^{j}\xi^{\prime}} \text{d}\xi^{\prime} \\[6pt] &=\int_{\mathbb R^{n}} \left(\Delta_{j}^{\Theta}f \right)( x+2^{-j}y ) \left[\widehat{\Psi} \sigma(2^{j}\cdot)\right]\widehat{\phantom{0}} (y) \, \text{d}y\\[6pt] &=\int_{\mathbb R^{n}} \frac{ \left(\Delta_{j}^{\Theta}f \right)( x+2^{-j}y )}{(1+|y|)^{s}} (1+|y|)^{s}\left[\widehat{\Psi} \sigma(2^{j}\cdot)\right]\widehat{\phantom{0}}\, ( y) \, \text{d}y. \end{align*} By the Hölder inequality in Lorentz spaces, we therefore obtain \begin{align*} |\Delta_{j} T_{\sigma} (f)(x)|\leq \left\|\frac{\left(\Delta_{j}^{\Theta}f \right)( x+2^{-j}y )}{(1+|y|)^{s}}\right\|_{L^{\frac{n}{s},\infty}(\mathbb R^{n})} \left\|(1+|y|)^{s}\left[\widehat{\Psi} \sigma(2^{j}\cdot)\right]\widehat{\phantom{0}}\, ( y)\right\|_{L^{\left(\frac{n}{s}\right)^{\prime},1}(\mathbb R^{n})}. \end{align*} Since |$\frac {n}{s}<2$|⁠, we can find a real number q such that |$\frac {n}{s}1$|⁠), \begin{align*} \left\| T_{\sigma}(f) \right\|_{L^{p}(\mathbb {R}^{n})} &\le C \left\| \left( \sum_{j \in \mathbb {Z}} |\Delta_{j} T_{\sigma} (f) |^{2} \right)^{\frac{1}{2}} \right\|_{L^{p}(\mathbb {R}^{n})} \leq C K \left\| \left( \sum_{j \in \mathbb {Z}} |M_{L^{q}}\left(\Delta_{j}^{\Theta} f\right) |^{2} \right)^{\frac{1}{2}} \right\|_{L^{p}(\mathbb {R}^{n})}\\ &=C K \left\|\left( \sum_{j \in \mathbb {Z}} M \left(\left|\Delta_{j}^{\Theta} f\right|^{q}\right)^{\frac{2}{q}}\right)^{\frac{q}{2}} \right\|_{L^{\frac{p}{q}}(\mathbb {R}^{n})}^{\frac{1}{q}} \leq C K \left\|\left( \sum_{j \in \mathbb {Z}} \left|\Delta_{j}^{\Theta} f\right|{}^{q \cdot \frac{2}{q}}\right)^{\frac{q}{2}} \right\|_{L^{\frac{p}{q}}(\mathbb {R}^{n})}^{\frac{1}{q}}\\ &= C K \left\|\left( \sum_{j \in \mathbb {Z}} \left|\Delta_{j}^{\Theta} f\right|{}^{2}\right)^{\frac{1}{2}} \right\|_{L^{p}(\mathbb {R}^{n})} \leq C K \|f\|_{L^{p}(\mathbb {R}^{n})}. \end{align*} If p ∈ (1, 2) then the result follows by duality. 3 Interpolation Our main goal in this section will be to discuss the following result. Theorem 3.1. Suppose that |$1 0 and 0 < a < π such that for all |$t\in \mathbb R$| we have \begin{align*} 0< A_{\tau}(t) \le \exp \{ A e^{a |t|} \} \, . \end{align*} Then for 0 < θ < 1 we have \begin{align*} \left\vert F(\theta )\right\vert\le \exp\left\{ \dfrac{\sin(\pi \theta)}{2}\int_{-\infty}^{\infty}\left[ \dfrac{\log |A_{0}(t ) | }{\cosh(\pi t)-\cos(\pi\theta)} + \dfrac{\log | A_{1}(t )| }{\cosh(\pi t)+\cos(\pi\theta)} \right]\text{d}t \right\}. \end{align*} We also need the following lemma, whose standard proof is omitted. Lemma 3.3. Let |$1 0, there exist smooth functions |$h_{j}^{\varepsilon }$|⁠, |$j=1,\dots , N_{\varepsilon }$|⁠, supported in cubes on |$\mathbb R^{n}$| with pairwise disjoint interiors, and nonzero complex constants |$c_{j}^{\varepsilon }$| such that \begin{align} f_{z}^{\varepsilon} = \sum_{j=1}^{N_{\varepsilon}} |c_{j}^{\varepsilon}|^{\frac p{2} (1-z) + \frac p{p_{1}} z} \, h_{j}^{\varepsilon}, \end{align} (3.14) |$ \|{f_{\theta }^{\varepsilon }-f} \|_{L^{2}(\mathbb R^{n})}\!<\! \varepsilon , $||$ \|{f_{it}^{\varepsilon }}\|_{L^{2}(\mathbb R^{n})} \!\leq\! \left (\|f \|_{L^{p}(\mathbb R^{n})} +\varepsilon \right )^{\frac {p}{2}}$|⁠, and |$ \|{f_{1+it}^{\varepsilon }}\|_{L^{p_{1}}(\mathbb R^{n})}\!\leq\! \left (\|f \|_{L^{p}(\mathbb R^{n})} +\varepsilon \right )^{\frac {p}{p_{1}}} . $| The next three lemmas generalize results which are well known in the context of Lebesgue spaces to the setting of Lorentz spaces |$L^{p,1}(\mathbb R^{n})$|⁠. Lemma 3.4. Let 0 < s < n. Then \begin{align*} \|(I-\Delta)^{-\frac{s}{2}} f\|_{L^{\infty}(\mathbb R^{n})} \leq C(n) \frac{s}{n-s} \|f\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}. \end{align*} Proof. Let Gs be the function defined for any |$x\in \mathbb R^{n}$| by \begin{align*} G_{s}(x)=\frac{1}{(4\pi)^{\frac{s}{2}}\Gamma\left(\frac{s}{2}\right)} \int_{0}^{\infty} e^{-\frac{\pi|x|^{2}}{\delta}} e^{-\frac{\delta}{4\pi}} \delta^{\frac{-n+s}{2}} \frac{\,\text{d}\delta}{\delta}. \end{align*} It is not difficult to show that |$G_{s}(x)\leq C(n)\frac {s}{n-s} |x|^{-n+s}$|⁠. Therefore, \begin{align*} |(I-\Delta)^{-\frac{s}{2}} f(x)|=|G_{s}\ast f(x)| \leq \int_{\mathbb R^{n}} G_{s}(y) |f(x-y)|\,\text{d}y \leq C(n) \frac{s}{n-s} \|f\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}. \end{align*} Lemma 3.5. Let |$1 0, and let Ψ be as in Theorem 1.1. Then we have the a priori estimate \begin{align} \|(I-\Delta)^{\frac{s}{2}}[\widehat{\Psi}f]\|_{L^{p,1}(\mathbb R^{n})} \leq C(n,s,p,\Psi) \|(I-\Delta)^{\frac{s}{2}}f\|_{L^{p,1}(\mathbb R^{n})}. \end{align} (3.15) Proof. Pick real numbers p0, p1 satisfying |$1 0 and the set \begin{align*} M=\{y\in (0,\infty): \sup_{y\leq r<\infty} f^{\ast}(r) r^{\frac{s-a}{n}}> f^{\ast}(y) y^{\frac{s-a}{n}}\} \end{align*} is open. Hence, M is a countable union of open intervals, namely, |$M=\bigcup _{k\in S} (a_{k},b_{k})$|⁠, where S is a countable set of positive integers. Also, observe that if y ∈ (ak, bk), then |$\sup _{y\leq r<\infty } f^{\ast }(r) r^{\frac {s-a}{n}}=f^{\ast }(b_{k}) b_{k}^{\frac {s-a}{n}}$|⁠. We have \begin{align*} \int_{0}^{\infty} \left(f^{\ast}(r) r^{\frac{s-a}{n}}\right)^{\ast}(y) y^{\frac{a}{n}-1}\,\text{d}y &\leq \int_{0}^{\infty} \sup_{y\leq r<\infty} f^{\ast}(r) r^{\frac{s-a}{n}} y^{\frac{a}{n}-1}\,\text{d}y\\ &=\int_{(0,\infty)\setminus \cup_{k\in S} (a_{k},b_{k})} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y +\sum_{k\in S} f^{\ast}(b_{k}) b_{k}^{\frac{s-a}{n}} \int_{a_{k}}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y. \end{align*} Furthermore, for every k ∈ S, \begin{align*} f^{\ast}(b_{k}) b_{k}^{\frac{s-a}{n}} \int_{a_{k}}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y &\leq f^{\ast}(b_{k}) b_{k}^{\frac{s-a}{n}}\int_{\max\left(a_{k},\frac{b_{k}}{2}\right)}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y \cdot \frac{\int_{0}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y}{\int_{\frac{b_{k}}{2}}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y}\\ &=\frac{1}{1-\left(\frac{1}{2}\right)^{\frac{a}{n}}} f^{\ast}(b_{k}) b_{k}^{\frac{s-a}{n}} \int_{\max\left(a_{k},\frac{b_{k}}{2}\right)}^{b_{k}} y^{\frac{a}{n}-1}\,\text{d}y\\ &\leq \frac{2^{\frac{s-a}{n}}}{1-\left(\frac{1}{2}\right)^{\frac{a}{n}}} \int_{a_{k}}^{b_{k}} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y\\ &\leq \frac{C(n)}{a} \int_{a_{k}}^{b_{k}} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y. \end{align*} Therefore, \begin{align*} \int_{0}^{\infty} \left(f^{\ast}(r) r^{\frac{s-a}{n}}\right)^{\ast}(y) y^{\frac{a}{n}-1}\,\text{d}y &\leq \int_{0}^{\infty} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y +\frac{C(n)}{a} \sum_{k\in S} \int_{a_{k}}^{b_{k}} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y\\ &\leq \frac{C(n)}{a} \int_{0}^{\infty} f^{\ast}(y) y^{\frac{s}{n}-1}\,\text{d}y. \end{align*} To prove Theorem 3.1 we will also need the notion of a measure preserving transformation. We say that a mapping |$h: \mathbb R^{n} \rightarrow (0,\infty )$| is measure preserving if, whenever E is a measurable subset of |$(0,\infty )$|⁠, the set |$h^{-1}E=\{x\in \mathbb R^{n}: h(x)\in E\}$| is a measurable subset of |$\mathbb R^{n}$| and the n-dimensional Lebesgue measure of h−1E is equal to the one-dimensional Lebesgue measure of E. For more details on measure preserving transformations, see, for example, [1, Chapter 2, Section 7]. Proof of Theorem 3.1. We first observe that, by (3.13), we have |$p_{1}\neq 2$|⁠. In fact, we can assume that 1 < p1 < 2 and 1 < p ≤ 2, otherwise the result will follow by duality. Further, if p = 2 then Theorem 3.1 is a consequence of Plancherel’s theorem and of the Sobolev embedding from Lemma 3.4, so it is sufficient to focus on the case p < 2 in what follows. Define \begin{align*} \theta=\frac{\frac{1}{p}-\frac{1}{2}}{\frac{1}{p_{1}}-\frac{1}{2}}. \end{align*} The assumption (3.13) yields |$\theta \in (0,\frac {s}{s_{1}})$|⁠, and therefore \begin{align*} \theta = \frac{s-s_{0}}{s_{1}-s_{0}} \end{align*} for some s0 ∈ (0, s). Fix a function σ satisfying \begin{align} \sup_{j\in \mathbb Z} \|(I-\Delta)^{\frac{s}{2}}[\widehat{\Psi} \sigma(2^{j}\cdot)]\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}<\infty, \end{align} (3.17) and denote |$\varphi _{j}=(I-\Delta )^{\frac {s}{2}}[\widehat {\Psi } \sigma (2^{j}\cdot )]$|⁠, |$j\in \mathbb Z$|⁠. Thanks to (3.17), we have |$\lim _{r\to \infty } \varphi _{j}^{\ast }(r)=0$|⁠. By [1, Chapter 2, Corollary 7.6], there is a measure preserving transformation |$h_{j}: \mathbb R^{n} \rightarrow (0,\infty )$| such that |$|\varphi _{j}|=\varphi _{j}^{\ast } \circ h_{j}$|⁠. For a complex number z with |$0\leq \Re (z)\leq 1$|⁠, we define \begin{align} \sigma_{z}(\xi) =\sum_{j\in \mathbb Z} (I - \Delta)^{-\frac{s_{0}(1-z)+s_{1}z}{2}} \left[\varphi_{j} h_{j}^{\frac{s-(1-z)s_{0}-zs_{1}}{n}}\right](2^{-j}\xi) \widehat{\Phi}(2^{-j}\xi), \end{align} (3.18) where |$\widehat {\Phi }$| is a Schwartz function supported in the set |$\{\xi \in \mathbb R^{n}: \frac {1}{4} \leq |\xi | \leq 4\}$| and |$\widehat {\Phi }\equiv 1$| on the support of |$\widehat {\Psi }$|⁠. Fix |$f, g\in \mathscr C_{0}^{\infty }$|⁠. Given ε > 0, let |$f_{z}^{\varepsilon }$| and |$g_{z}^{\varepsilon }$| be functions having the form (3.14), with f replaced by g and with p replaced by p′ in the latter case, satisfying |$ \left \|{f_{\theta }^{\varepsilon }-f}\right \|_{L^{2}(\mathbb R^{n})}<\varepsilon $|⁠, |$ \left \|{g_{\theta }^{\varepsilon }-g}\right \|_{L^{2}(\mathbb R^{n})}<\varepsilon , $| and \begin{align} & \left\|{f_{it}^{\varepsilon}}\right\|_{L^{2}(\mathbb R^{n})}\le \left( \left\|{f}\right\|_{L^{p}(\mathbb R^{n})}+\varepsilon\right)^{\frac p{2}},\quad \left\|{f_{1+it}^{\varepsilon}}\right\|_{L^{p_{1}}(\mathbb R^{n})}\le \left( \left\|{f}\right\|_{L^{p}(\mathbb R^{n})}+\varepsilon\right)^{\frac p{p_{1}}},\\ \nonumber & \left\|{g_{it}^{\varepsilon}}\right\|_{L^{2}(\mathbb R^{n})}\le \left( \left\|{g}\right\|_{L^{p^{\prime}}(\mathbb R^{n})}+\varepsilon\right)^{\frac{p^{\prime}}{2}},\quad \left\|{g_{1+it}^{\varepsilon}}\right\|_{L^{p_{1}^{\prime}}(\mathbb R^{n})}\le \left( \left\|{g}\right\|_{L^{p^{\prime}}(\mathbb R^{n})}+\varepsilon\right)^{\frac{p^{\prime}}{p_{1}^{\prime}}}. \end{align} (3.19) For a complex number z with |$0\leq \Re (z) \leq 1$|⁠, define \begin{align*} F(z) =& \int_{\mathbb R^{n}} T_{\sigma_{z}}\left(f_{z}^{\varepsilon}\right) g_{z}^{\varepsilon}\; \text{d}x =\int_{\mathbb R^{n}} \sigma_{z}(\xi) \widehat{f^{\varepsilon}_{z}}(\xi) \widehat{g^{\varepsilon}_{z}}(\xi)\, \text{d}\xi. \end{align*} It is straightforward (but rather tedious) to verify that F is analytic on the strip |$S=\{z\in \mathcal C: 0 <\Re (z)<1\}$| and continuous on its closure. Let us write z = τ + it, 0 ≤ τ ≤ 1 and |$t\in \mathbb R$|⁠, and denote sτ = s0(1 − τ) + s1τ. Then, applying Lemmas 3.4 and 3.5 and using the fact that hj is measure preserving, we obtain \begin{align*} \|\sigma_{z}\|_{L^{\infty}(\mathbb R^{n})} &\leq C(n) \sup_{j\in \mathbb Z} \left\|(I-\Delta)^{-\frac{s_{0}(1-z)+s_{1}z}{2}}\left[\varphi_{j} h_{j}^{\frac{s-(1-z)s_{0}-zs_{1}}{n}}\right]\right\|_{L^{\infty}(\mathbb R^{n})}\\ &\leq C(n) \frac{s_{\tau}}{n-s_{\tau}} \sup_{j\in \mathbb Z} \left\|(I-\Delta)^{-\frac{s_{0}(-it)+s_{1}it}{2}}\left[\varphi_{j} h_{j}^{\frac{s-(1-\tau-it)s_{0}-(\tau+it)s_{1}}{n}}\right]\right\|_{L^{\frac{n}{s_{\tau}},1}(\mathbb R^{n})}\\ &\leq C(n,s_{0},s_{1}) \frac{s_{\tau}}{n-s_{\tau}} (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \left\|\varphi_{j} h_{j}^{\frac{s-(1-\tau-it)s_{0}-(\tau+it)s_{1}}{n}}\right\|_{L^{\frac{n}{s_{\tau}},1}(\mathbb R^{n})}\\ &\leq C(n,s_{0},s_{1}) (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \left\|\varphi_{j}^{\ast}(r) r^{\frac{s-(1-\tau)s_{0}-\tau s_{1}}{n}}\right\|_{L^{\frac{n}{s_{\tau}},1}(0,\infty)}\\ &\leq C(n,s_{0},s_{1}) (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \left\|\varphi_{j}^{\ast}\right\|_{L^{\frac{n}{s},1}(0,\infty)}\\ &\leq C(n,s_{0},s_{1}) (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \|\varphi_{j}\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}. \end{align*} Notice that if τ ∈ [0, θ), then the penultimate inequality follows from Lemma 3.7. Thus, \begin{align} |F(z)|&\leq \|\sigma_{z}\|_{L^{\infty}(\mathbb R^{n})} \left\|f^{\varepsilon}_{z}\right\|_{L^{2}(\mathbb R^{n})} \left\|g^{\varepsilon}_{z}\right\|_{L^{2}(\mathbb R^{n})}\\ \nonumber &\leq C(n,s_{0},s_{1}) (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \|\varphi_{j}\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \left\|f^{\varepsilon}_{z}\right\|_{L^{2}(\mathbb R^{n})} \left\|g^{\varepsilon}_{z}\right\|_{L^{2}(\mathbb R^{n})}. \end{align} (3.20) Since |$\|f^{\varepsilon }_{z}\|_{L^{2}(\mathbb R^{n})} \|g^{\varepsilon }_{z}\|_{L^{2}(\mathbb R^{n})}$| can be bounded from above by a constant independent of z, the previous estimate yields \begin{align} |F(z)|\leq C(n,s_{0},s_{1},p,p_{1},\varepsilon,f,g) (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \|\varphi_{j}\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \leq \exp\{A e^{a|t|}\} \end{align} (3.21) for a suitable choice of constants A > 0 and a ∈ (0, π). Also, if z = it, |$t\in \mathbb R$|⁠, then (3.20) combined with (3.19) yield \begin{align} |F(it)|\leq C(n,s_{0},s_{1}) (1+|t|)^{\frac{n}{2}+1} \left( \left\|{f}\right\|_{L^{p}(\mathbb R^{n})}+\varepsilon\right)^{\frac p{2}} \big( \left\|{g}\right\|_{L^{p^{\prime}}(\mathbb R^{n})}+\varepsilon\big)^{\frac{p^{\prime}}{2}} \sup_{j\in \mathbb Z} \|\varphi_{j}\|_{L^{\frac{n}{s},1}(\mathbb R^{n})}. \end{align} (3.22) Finally, by Hölder’s inequality and by (3.12), \begin{align*} |F(1+it)|&\leq \left\|T_{\sigma_{1+it}}(f^{\varepsilon}_{1+it})\right\|_{L^{p_{1}}(\mathbb R^{n})} \left\|g^{\varepsilon}_{1+it}\right\|_{L^{p^{\prime}_{1}}(\mathbb R^{n})}\\ &\leq C\sup_{j\in \mathbb Z} \left\|(I-\Delta)^{\frac{s_{1}}{2}} \left[\widehat{\Psi} \sigma_{1+it}(2^{j}\cdot )\right]\right\|_{L^{\frac{n}{s_{1}},1}(\mathbb R^{n})} \left\|f^{\varepsilon}_{1+it}\right\|_{L^{p_{1}}(\mathbb R^{n})} \left\|g^{\varepsilon}_{1+it}\right\|_{L^{p^{\prime}_{1}}(\mathbb R^{n})}. \end{align*} Notice that |$\widehat {\Psi } \sigma _{1+it}(2^{k}\cdot )$| picks up only those terms j of (3.18) which differ from k by at most two units. For simplicity, we may therefore take j = k in the calculation below. We have \begin{align*} \Bigg\|(I-\Delta)^{\frac{s_{1}}{2}}& \left[\widehat{\Psi}(I - \Delta)^{-\frac{s_{1}+it(s_{1}-s_{0})}{2}} \left[\varphi_{j} h_{j}^{\frac{s-s_{1}+it(s_{0}-s_{1})}{n}}\right]\right]\Bigg\|_{L^{\frac{n}{s_{1}},1}(\mathbb R^{n})}\\ &\leq C\left \|(I-\Delta)^{\frac{s_{1}}{2}} \left[(I - \Delta)^{-\frac{s_{1}+it(s_{1}-s_{0})}{2}} \left[\varphi_{j} h_{j}^{\frac{s-s_{1}+it(s_{0}-s_{1})}{n}}\right]\right]\right\|_{L^{\frac{n}{s_{1}},1}(\mathbb R^{n})}\\ &\leq C\left\|(I - \Delta)^{-\frac{it(s_{1}-s_{0})}{2}} \left[\varphi_{j} h_{j}^{\frac{s-s_{1}+it(s_{0}-s_{1})}{n}}\right]\right\|_{L^{\frac{n}{s_{1}},1}(\mathbb R^{n})}\\ &\leq C (1+|t|)^{\frac{n}{2}+1}\left\|\varphi_{j} h_{j}^{\frac{s-s_{1}}{n}} \right\|_{L^{\frac{n}{s_{1}},1}(\mathbb R^{n})} =C (1+|t|)^{\frac{n}{2}+1} \left\|\varphi_{j}^{\ast}(r) r^{\frac{s-s_{1}}{n}}\right\|_{L^{\frac{n}{s_{1}},1}(0,\infty)}\\ &=C (1+|t|)^{\frac{n}{2}+1} \left\|\varphi_{j}^{\ast} \right\|_{L^{\frac{n}{s},1}(0,\infty)} = C (1+|t|)^{\frac{n}{2}+1} \|\varphi_{j} \|_{L^{\frac{n}{s},1}(\mathbb R^{n})}. \end{align*} Notice that in the previous estimate we consecutively used Lemmas 3.6 and 3.5 and the fact that hj is measure preserving. Therefore, \begin{align} |F(1+it)|\leq C (1+|t|)^{\frac{n}{2}+1} \sup_{j\in \mathbb Z} \|\varphi_{j} \|_{L^{\frac{n}{s},1}(\mathbb R^{n})} (\|f\|_{L^{p}(\mathbb R^{n})}+\varepsilon)^{\frac{p}{p_{1}}} (\|g\|_{L^{p^{\prime}}(\mathbb R^{n})}+\varepsilon)^{\frac{p^{\prime}}{p_{1}^{\prime}}}. \end{align} (3.23) A combination of (3.21), (3.22), (3.23), and Lemma 3.2 yields \begin{align} |F(\theta)|\leq C\sup_{j\in \mathbb Z} \|\varphi_{j} \|_{L^{\frac{n}{s},1}(\mathbb R^{n})} (\|f\|_{L^{p}(\mathbb R^{n})}+\varepsilon) (\|g\|_{L^{p^{\prime}}(\mathbb R^{n})}+\varepsilon). \end{align} (3.24) Observe that |$ F(\theta ) = \int _{\mathbb R^{n}} \sigma (\xi ) \widehat {f_{\theta }^{\varepsilon }}(\xi ) \widehat { g_{\theta }^{\varepsilon }}(\xi ) \, \text {d}\xi $| as for every |$\xi \neq 0$|⁠, \begin{align*} \sigma_{\theta}(\xi) &=\sum_{j\in \mathbb Z} (I-\Delta)^{-\frac{s}{2}}\left[(I-\Delta)^{\frac{s}{2}}\left[\sigma(2^{j}\cdot)\widehat{\Psi}\right]\right](2^{-j}\xi) \widehat{\Phi}(2^{-j}\xi)\\ &=\sum_{j\in \mathbb Z} \sigma(\xi) \widehat{\Psi}(2^{-j}\xi) \widehat{\Phi}(2^{-j}\xi) =\sum_{j\in \mathbb Z} \sigma(\xi) \widehat{\Psi}(2^{-j}\xi) =\sigma(\xi). \end{align*} Notice that \begin{align*} &\left| \int_{\mathbb R^{n}} \sigma(\xi) \widehat{f_{\theta}^{\varepsilon}}(\xi) \widehat{ g_{\theta}^{\varepsilon}}(\xi) \, \text{d}\xi - \int_{\mathbb R^{n}} \sigma(\xi) \widehat{f }(\xi) \widehat{ g }(\xi) \, \text{d}\xi \right| \\ &\quad= \left| \int_{\mathbb R^{n}} \sigma(\xi) \left[ \widehat{f_{\theta}^{\varepsilon}}(\xi) \left(\widehat{ g_{\theta}^{\varepsilon}}(\xi)-\widehat{ g }(\xi) \right) + \widehat{g }(\xi) \left(\widehat{ f_{\theta}^{\varepsilon}}(\xi)-\widehat{ f }(\xi) \right) \right]\, \text{d}\xi \right| \\ &\quad\le \|\sigma\|_{L^{\infty}(\mathbb R^{n})} \left[ \left\|f_{\theta}^{\varepsilon}\right\|_{L^{2}(\mathbb R^{n})} \left\|{g}_{\theta}^{\varepsilon} -g\right\|_{L^{2}(\mathbb R^{n})} + \|g\|_{L^{2}(\mathbb R^{n})} \left\|{f}_{\theta}^{\varepsilon} -f\right\|_{L^{2}(\mathbb R^{n})} \right] \\ &\quad\le C\sup_{j\in \mathbb Z} \left\|(I-\Delta)^{\frac{s}{2}}\left[\widehat{\Psi}\sigma(2^{j}\cdot)\right]\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \left[ \left\|f^{\varepsilon}_{\theta} \right\|_{L^{2}(\mathbb R^{n})} \left\|{g}_{\theta}^{\varepsilon}\! -g\!\right\|_{L^{2}(\mathbb R^{n})}\! +\! \|g \|_{L^{2}(\mathbb R^{n})} \left\|{f}_{\theta}^{\varepsilon} -f\right\|_{L^{2}(\mathbb R^{n})} \right]\, . \end{align*} Recall that |${f}_{\theta }^{\varepsilon } -f$| and |${g}_{\theta }^{\varepsilon } -g$| converge to zero in |$L^{2}(\mathbb R^{n})$| as ε → 0. Letting ε → 0 in (3.24) yields \begin{align*} \left|\int_{\mathbb R^{n}} \sigma(\xi) \widehat{f}\ (\xi)\ \widehat{g}\ (\xi) \, \text{d}\xi \right|\le C\, \sup_{j\in\mathbb Z} \left\|{(I-\Delta)^{\frac{s}2}\left[\sigma(2^{j}\cdot)\widehat{\Psi}\right]}\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \left\|{f}\right\|_{L^{p}(\mathbb R^{n})} \|g\|_{L^{p^{\prime}}(\mathbb R^{n})}. \end{align*} Taking the supremum over all functions |$g\in L^{p^{\prime}}(\mathbb R^{n})$| with |$\| g\|_{L^{p^{\prime}}(\mathbb R^{n})} \le 1$| we obtain \begin{align*} \left\|{T_{\sigma }(f)}\right\|_{L^{p}(\mathbb R^{n})}\le C\, \sup_{j\in\mathbb Z} \left\|{(I-\Delta)^{\frac{s}2}\left[\sigma(2^{j}\cdot)\widehat{\Psi}\right]}\right\|_{L^{\frac{n}{s},1}(\mathbb R^{n})} \left\|{f}\right\|_{L^{p}(\mathbb R^{n})}. \end{align*} This completes the proof of Theorem 3.1. Acknowledgments We would like to thank the referee who pointed out that Theorem 3.1 in Section 3 could also be obtained by an argument using the idea in Connett and Schwartz [5], [6]. The first author acknowledges the support of the Simons Foundation and of the University of Missouri Research Board and Research Council. References [1] Bennett , C. and R. Sharpley . Interpolation of Operators . Boston : Academic Press , 1988 . Google Preview WorldCat COPAC [2] Calderón , A. P. and A. Torchinsky . “ Parabolic maximal functions associated with a distribution, II .” Adv. Math. 24 ( 1977 ): 101 – 71 . Google Scholar Crossref Search ADS WorldCat [3] Cianchi , A. and L. Pick . “ Sobolev embeddings into BMO, VMO and |$L^\infty $| .” Ark. Mat. 36, ( 1998 ): 317 – 40 . [4] Cianchi , A. , L. Pick , and L. Slavíková . “ Higher-order Sobolev embeddings and isoperimetric inequalities .” Adv. Math. 273 , ( 2015 ): 568 – 650 . Google Scholar Crossref Search ADS WorldCat [5] Connett , W. C. and A. L. Schwartz . “ The theory of ultraspherical multipliers .” Mem. Amer. Math. Soc. 9 , no. 183 ( 1977 ): iv+92 . WorldCat [6] Connett , W. C. and A. L. Schwartz . “ A remark about Calderón’s upper s method of interpolation .” Interpolation Spaces and Allied Topics in Analysis (Lund, 1983), 48–53, Lecture Notes in Math., 1070 , Berlin : Springer , 1984 . Google Preview WorldCat COPAC [7] Edmunds , D. E. and B. Opic . “ Boundedness of fractional maximal operators between classical and weak type Lorentz spaces .” Dissertationes Math. (Rozprawy Mat.) 410 , ( 2002 ): 50 . WorldCat [8] Grafakos , L. Classical Fourier Analysis, 3rd ed GTM 249 , NY : Springer , 2014 . Google Preview WorldCat COPAC [9] Grafakos , L. , D. He , P. Honzík , and H. V. Nguyen . ”The Hörmander multiplier theorem I: the linear case revisited.” Illinois J . Math. In press. [10] Hirschman , I. I. Jr. “ A convexity theorem for certain groups of transformations .” J. Analyse Math. 2 , ( 1953 ): 209 – 18 . Google Scholar Crossref Search ADS WorldCat [11] Hirschman , I. I. Jr. “ On multiplier transformations .” Duke Math. J. 26 , ( 1959 ): 221 – 42 . Google Scholar Crossref Search ADS WorldCat [12] Hörmander , L. “ Estimates for translation invariant operators in Lp spaces .” Acta Math. 104 , ( 1960 ): 93 – 139 . [13] Mikhlin , S. G. “ On the multipliers of Fourier integrals .” (Russian) Dokl. Akad. Nauk SSSR (N.S.) 109 , ( 1956 ): 701 – 03 . WorldCat [14] Miyachi , A. “ On some Fourier multipliers for |$H^p(\mathbb R^n)$| .” J. Fac. Sci. Univ. Tokyo Sect. IA Math. 27 , ( 1980 ): 157 – 79 . WorldCat [15] Miyachi , A. and N. Tomita . “ Minimal smoothness conditions for bilinear Fourier multipliers .” Rev. Mat. Iberoam. 29 , ( 2013 ): 495 – 530 . Google Scholar Crossref Search ADS WorldCat [16] Seeger , A. “ A limit case of the Hörmander multiplier theorem .” Monatsh. Math. 105 , ( 1988 ): 151 – 60 . Google Scholar Crossref Search ADS WorldCat [17] Seeger , A. “ Estimates near L1 for Fourier multipliers and maximal functions .” Arch. Math. (Basel) 53 , ( 1989 ): 188 – 93 . Google Scholar Crossref Search ADS WorldCat [18] Seeger , A. “ Remarks on singular convolution operators .” Studia Math. 97 , no. 2 ( 1990 ): 91 – 114 . Google Scholar Crossref Search ADS WorldCat [19] Stein , E. M. Singular Integral and Differentiability Properties of Functions. Princeton Univ . Princeton, NJ : Press , 1970 . Google Preview WorldCat COPAC [20] Stein , E. M. ”Editor’s note: the differentiability of functions in |$\mathbb {R}^n$| .” Ann. Math. (2) 113 , no. 2 ( 1981 ): 383 – 85 . [21] Wainger , S. “ Special trigonometric series in k-dimensions .” Mem. Amer. Math. Soc. 59 , ( 1965 ): 1 – 102 . WorldCat Communicated by Prof. Assaf Naor © The Author(s) 2018. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permission@oup.com. This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model) TI - A Sharp Version of the Hörmander Multiplier Theorem JF - International Mathematics Research Notices DO - 10.1093/imrn/rnx314 DA - 2019-08-02 UR - https://www.deepdyve.com/lp/oxford-university-press/a-sharp-version-of-the-h-rmander-multiplier-theorem-nyYYevsVwg SP - 4764 VL - 2019 IS - 15 DP - DeepDyve ER -