\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 170, pp. 1--24.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/170\hfil Generalized Leray-alpha equations] {Local and global low-regularity solutions to generalized Leray-alpha equations} \author[N. Pennington \hfil EJDE-2015/170\hfilneg] {Nathan Pennington} \address{Nathan Pennington \newline Department of Mathematics, Creighton University, 2500 California Plaza, Omaha, NE 68178, USA} \email{nathanpennington@creighton.edu} \thanks{Submitted May 15, 2015. Published June 18, 2015.} \subjclass[2010]{76D05, 35A02, 35K58} \keywords{Leray-alpha model; Besov space; fractional Laplacian} \begin{abstract} It has recently become common to study approximating equations for the Navier-Stokes equation. One of these is the Leray-$\alpha$ equation, which regularizes the Navier-Stokes equation by replacing (in most locations) the solution $u$ with $(1-\alpha^2\Delta)u$. Another is the generalized Navier-Stokes equation, which replaces the Laplacian with a Fourier multiplier with symbol of the form $-|\xi|^\gamma$ ($\gamma=2$ is the standard Navier-Stokes equation), and recently in \cite{taolog} Tao also considered multipliers of the form $-|\xi|^\gamma/g(|\xi|)$, where $g$ is (essentially) a logarithm. The generalized Leray-$\alpha$ equation combines these two modifications by incorporating the regularizing term and replacing the Laplacians with more general Fourier multipliers, including allowing for $g$ terms similar to those used in \cite{taolog}. Our goal in this paper is to obtain existence and uniqueness results with low regularity and/or non-$L^2$ initial data. We will also use energy estimates to extend some of these local existence results to global existence results. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{corollary}[theorem]{Corollary} \allowdisplaybreaks \section{Introduction} The incompressible form of the Navier-Stokes equation is given by \begin{equation}\label{NS} \begin{gathered} \partial_t u + (u\cdot \nabla)u=\nu\Delta u-\nabla p, \\ u(0,x)=u_0(x), \quad \operatorname{div}(u)=0 \end{gathered} \end{equation} where $u:I\times \mathbb{R}^n\to \mathbb{R}^n$ for some time strip $I=[0,T)$, $\nu>0$ is a constant due to the viscosity of the fluid, $p:I\times\mathbb{R}^n\to \mathbb{R}^n$ denotes the fluid pressure, and $u_0:\mathbb{R}^n\to\mathbb{R}^n$. The requisite differential operators are defined by $\Delta=\sum_{i=1}^n \frac{\partial^2}{\partial_{x_i}^2}$ and $\nabla=(\frac{\partial} {\partial_{x_i}},\dots ,\frac{\partial}{\partial_{x_n}})$. In dimension $n=2$, local and global existence of solutions to the Navier-Stokes equation are well known (see \cite{lady356}; for a more modern reference, see \cite[Chapter 17]{T3}). For dimension $n\geq 3$, the problem is significantly more complicated. There is a robust collection of local existence results, including \cite{Kato}, in which Kato proves the existence of local solutions to the Navier-Stokes equation with initial data in $L^n(\mathbb{R}^n)$; \cite{KP}, where Kato and Ponce solve the equation with initial data in the Sobolev space $H^{n/p-1,p}(\mathbb{R}^n)$; and \cite{KT}, where Koch and Tataru establish local existence with initial data in the space $BMO^{-1}(\mathbb{R}^n)$ (for a more complete accounting of local existence theory for the Navier-Stokes equation, see \cite{nsbook}). In all of these local results, if the initial datum is assumed to be sufficiently small, then the local solution can be extended to a global solution. However, the issue of global existence of solutions to the Navier-Stokes equation in dimension $n\geq 3$ for arbitrary initial data is one of the most challenging open problems remaining in analysis. Because of the intractability of the Navier-Stokes equation, many approximating equations have been studied. One of these is the Leray-$\alpha$ model, which is %\label{leray1} \begin{gather*} \partial_t(1-\alpha^2\Delta )u+\nabla_u (1-\alpha^2 \Delta) u-\nu\Delta (1-\alpha^2 \Delta )u=-\nabla p, \\ u(0,x)=u_0(x), ~~ \operatorname{div}u_0=\operatorname{div}u=0, \end{gather*} where we recall that $\nabla_u v=(u\cdot \nabla)v$. Note that setting $\alpha=0$ returns the standard Navier-Stokes equation. Like the Lagrangian Averaged Navier-Stokes (LANS) equation (which differs from the Leray-$\alpha$ in the presence of an additional nonlinear term), the system \eqref{leray1} compares favorably with numerical data; see \cite{holmleray}, in which the authors compared the Reynolds numbers for the Leray-$\alpha$ equation and the LANS equation with the Navier-Stokes equation. Another commonly studied equation is the generalized Navier-Stokes equation, given by \begin{gather*} \partial_t u + (u\cdot \nabla)u=\nu\mathcal{L} u-\nabla p,\\ u(0,x)=u_0(x), \quad \operatorname{div}(u)=0 \end{gather*} where $\mathcal{L}$ is a Fourier multiplier with symbol $m(\xi)=-|\xi|^\gamma$ for $\gamma>0$. Choosing $\gamma=2$ returns the standard Navier-Stokes equation. In \cite{Wu}, Wu proved (among other results) the existence of unique local solutions for this equation provided the data is in the Besov space $B^s_{p,q}(\mathbb{R}^n)$ with $s=1+n/p-\gamma$ and $1<\gamma\leq 2$. If the norm of the initial data is sufficiently small, these local solutions can be extended to global solutions. It is well known that if $\gamma\geq \frac{n+2}{2}$, then this equation has a unique global solution. In \cite{taolog}, Tao strengthened this result, proving global existence with the symbol $m(\xi)=-|\xi|^\gamma/g(|\xi|)$, with $\gamma\geq \frac{n+2}{2}$ and $g$ a non-decreasing, positive function that satisfies \[ \int_1^\infty \frac{ds}{sg(s)^2}=+\infty. \] Note that $g(|x|)=\log^{1/2}(2+|x|^2)$ satisfies the condition. Similar types of results involving $g$ terms that are, essentially, logarithms have been proven for the nonlinear wave equation; see \cite{taolog} for a more detailed description. Here we consider a combination of these two models, called the generalized Leray-$\alpha$ equation, which is \begin{equation} \label{leray1} \begin{gathered} \partial_t(1-\alpha^2 \mathcal{L}_2 )u+\nabla_u (1-\alpha^2 \mathcal{L}_2) u -\nu\mathcal{L}_1 (1-\alpha^2 \mathcal{L}_2 )u=-\nabla p, \\ u(0,x)=u_0(x), \quad \operatorname{div}u_0=\operatorname{div}u=0, \end{gathered} \end{equation} with the operators $\mathcal{L}_i$ defined by \[ \mathcal{L}_iu(x)=\int -\frac{|\xi|^{\gamma_i}}{g_i(\xi)}\hat{u}(\xi) e^{ix\cdot \xi}d\xi, \] where $g_i$ are radially symmetric, nondecreasing, and bounded below by $1$. Note that if $g_2=1$ and $\gamma_2=0$, then $\mathcal{L}_2 u(x)=-u(x)$, so choosing $g_1=g_2=1$, $\gamma_1=2$, and $\gamma_2=0$ returns the Navier-Stokes equation (after absorbing $(1+\alpha^2)^{-1}$ into the pressure function $p$). Choosing $g_1=g_2=1$ and $\gamma_1=\gamma_2=2$ gives the Leray-$\alpha$ equation, and choosing $g_2=\gamma_2=1$ returns the generalized Navier-Stokes equation. In \cite{barbato}, the authors proved the existence of a smooth global solution to the generalized Leary-$\alpha$ equation with smooth initial data provided $\gamma_1+\gamma_2\geq n/2+1$, $g_2=1$, and $g_1$ is in a category similar to, though inclusive of, the type of $g$ required in Tao's argument in \cite{taolog}. In \cite{kazuo}, Yamazaki obtains a unique global solution to equation \eqref{leray1} in dimension three provided $(1-\alpha^2\mathcal{L}_2)u_0$ is in the Sobolev space $H^{m,2}(\mathbb{R}^3)$, where $u_0$ is the initial data, $m>\max\{5/2, 1+2\gamma_1\}$, and provided $\gamma_1$ and $\gamma_2$ satisfy the inequality $2\gamma_1+\gamma_2\geq 5$ and that $g_1$ and $g_2$ satisfy \begin{equation}\label{g restriction} \int_1^\infty \frac{ds}{s g_1^2(s)g_2(s)} =\infty. \end{equation} The goal of this article is to obtain a much wider array of existence results, specifically existence results for initial data with low regularity and for initial data outside the $L^2$ setting. We will also, where applicable, use the energy bound from \cite{kazuo} to extend these local solutions to global solutions. Our plan is to follow the general contraction-mapping based procedure outlined by Kato and Ponce in \cite{KP} for the Navier-Stokes equation, with two key modifications. First, the approach used in \cite{KP} relies heavily on operator estimates for the heat kernel $e^{t\Delta}$. We will require similar estimates for our solution operator $e^{t\mathcal{L}_1}$ and some operator estimates for $(1-\alpha^2\mathcal{L}_2)$, and establishing these estimates is the purpose of Section \ref{L operator estimates} and Section \ref{operator est}. This will require some technical restrictions on the choices of $g_1$ and $g_2$ that will be more fully addressed below. We also note that these estimates should allow the application of this general technique to other similar equations, like the MHD equation in \cite{kazuomhd} and \cite{kazuo}, the generalized MHD equation found in \cite{globalmhd} (see \cite{genmdh} for a general study of the generalized MHD equation), the logarithmically super critical Boussinesq system in \cite{hmidi}, and the Navier-Stokes like equation studied in \cite{olsontitialphalike}. The second modification is in how we will deal with the nonlinear term. For the first set of results, we will use the standard Leibnitz-rule estimate to handle the nonlinear terms. Our second set of results rely on a product estimate (due to Chemin in \cite{chemin}) which will allow us to obtain lower regularity existence but will (among other costs) require us to work in Besov spaces. The advantages and disadvantages of each approach will be detailed later in this introduction. The product estimates themselves are stated as Proposition \ref{product est 1} and Proposition $\ref{product est 2}$ in Section \ref{Besov space}. We will also need bounds on the terms $(1-\mathcal{L}_2)$ and $(1-\mathcal{L}_2)^{-1}$, and establishing these bounds is the subject of Section \ref{L operator estimates}. The rest of this paper is organized as follows. The remainder of this introduction is devoted to stating and contextualizing the main results of the paper. Section \ref{Besov space} reviews the basic construction of Besov spaces and states some foundational results, including our two product estimates. In Section \ref{type 1} we carry out the existence argument using the standard product estimate, and in Section \ref{type 2} we obtain existence results using the other product estimate. As stated above, Sections \ref{L operator estimates} and \ref{operator est} contain the proofs of the operator estimates that are central to the arguments used in Sections \ref{type 1} and \ref{type 2}. Our last task before stating the main results is to establish some notation. First, we denote Besov spaces by $B^s_{p,q}(\mathbb{R}^n)$, with norm denoted by $\|\cdot\|_{B^s_{p,q}}=\|\cdot\|_{s,p,q}$ (a complete definition of these spaces can be found in Section \ref{Besov space}). We define the space \[ C^T_{a;s,p,q}=\{f\in C((0,T):B^s_{p,q}(\mathbb{R}^n)):\|f\|_{a;s,p,q}<\infty\}, \] where \[ \|f\|_{a;s,p,q}=\sup\{t^a\|f(t)\|_{s,p,q}:t\in (0,T)\}, \] $T>0$, $a\geq 0$, and $C(A:B)$ is the space of continuous functions from $A$ to $B$. We let $\dot{C}^T_{a;s,p,q}$ denote the subspace of $C^T_{a;s,p,q}$ consisting of $f$ such that \[ \lim_{t\to 0^+}t^a f(t)=0 \quad \text{(in } B^s_{p,q}(\mathbb{R}^n)). \] Note that while the norm $\|\cdot\|_{a;s,p,q}$ lacks an explicit reference to $T$, there is an implicit $T$ dependence. We also say $u\in BC(A:B)$ if $u\in C(A:B)$ and $\sup_{a\in A}\|u(a)\|_{B}<\infty$. Now we are ready to state the existence results. As expected in these types of arguments, the full result gives unique local solutions provided the parameters satisfy a large collection of inequalities. Here we state special cases of the full results. Our first Theorem uses the standard product estimate (Proposition \ref{product est 1} in Section \ref{Besov space}). \begin{theorem}\label{old style short} Let $\gamma_1>1$, $\gamma_2>0$, $q\geq 1$, $p\geq 2$, and let $s_1$, $s_2$ be real numbers such that $s_2>\gamma_2$, $01$, $\gamma_2>0$, $q\geq 1$, $p\geq 2$, $s_1$, and $s_2$ satisfy \begin{gather*} 0\gamma_2-n/p-1, \\ \gamma_1\geq 2s_2-s_1-\gamma_2+n/p+1,\\ n/p>\gamma_2/2, \\ s_2\geq \gamma_2/2. \end{gather*} We also assume that $g_1$ and $g_2$ are Mikhlin multipliers (see inequality \eqref{mikhlin condition don}). Then for any divergence free $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$, there exists a unique local solution $u$ to the generalized Leray-alpha equation \eqref{leray1}, with \[ u\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}^T_{a;s_2,p,q}, \] where $a=(s_2-s_1)/\gamma_1$. $T$ can be chosen to be a non-increasing function of $\|u_0\|_{B^{s_1}_{p, q}}$ with $T=\infty$ if $\|u_0\|_{B^{s_1}_{p, q}}$ is sufficiently small. \end{theorem} We remark that in the first theorem, $\gamma_2$ can be arbitrarily large, but $s_1>-1$, while in the second theorem $\gamma_2<2n/p$, but for sufficiently large $\gamma_1$ and sufficiently small $\gamma_2$, $s_1>\gamma_2-n/p-1$ can be less than $-1$. Thus the non-standard product estimate allows us to obtain existence results for initial data with lower regularity, but requires $\gamma_2$ to be small and requires the use of Besov spaces. We also note that if we set $\gamma_2=0$ and $g_2(|\xi|)=1$ (and thus are back in the case of the generalized Navier-Stokes equation), then these techniques would recover the results of Wu in \cite{Wu} for the generalized Navier-Stokes equation. As was stated above, these results will hold if the $g_i$ are Mikhlin multipliers. However, there are interesting choices of $g_i$ (specifically $g_i$ being, essentially, a logarithm) which are not Mikhlin multipliers. For this case we have analogous, but slightly weaker, results. In what follows, we let $r^-$ indicate a number arbitrarily close to, but strictly less than, $r$ (and similarly let $r^+$ be a number arbitrarily close to, but strictly greater than, $r$). \begin{theorem}\label{old style short log} Let $\gamma_1>1$, $\gamma_2>0$, $q\geq 1$, $p\geq 2$, and $s_1$, $s_2$ be real numbers such that $s_2>\gamma_2$, $00$ and $|g_i^{(k)}(r)|\leq Cr^{-k}$ for $1\leq k\leq n/2+1$. Then for any divergence free $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$, there exists a unique local solution $u$ to the generalized Leray-alpha equation \eqref{leray1}, with \[ u\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}^T_{a;s_2,p,q}, \] where $a=(s_2-s_1)/\gamma_1^-$ for arbitrarily small $\varepsilon>0$. $T$ can be chosen to be a non-increasing function of $\|u_0\|_{B^{s_1}_{p, q}}$ with $T=\infty$ if $\|u_0\|_{B^{s_1}_{p, q}}$ is sufficiently small. \end{theorem} \begin{theorem}\label{special case 1 log} Let $\gamma_1>1$, $\gamma_2>0$, $q\geq 1$, $p\geq 2$, $s_1$ and $s_2$ satisfy \begin{gather*} 0\gamma^-_2-n/p-1, \\ \gamma^-_1\geq 2s_2-s_1-\gamma^-_2+n/p+1,\\ n/p>\gamma^-_2/2,\\ s_2\geq \gamma^-_2/2. \end{gather*} We also assume that, for $i=1,2$, $|g_i(r)|\leq Cr^\delta$ for any $\delta>0$ and $|g_i^{(k)}(r)|\leq Cr^{-k}$ for $1\leq k\leq n/2+1$. Then for any divergence free $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$, there exists a unique local solution $u$ to the generalized Leray-alpha equation \eqref{leray1}, with \[ u\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}^T_{a;s_2,p,q}, \] where $a=(s_2-s_1)/\gamma^-_1$. $T$ can be chosen to be a non-increasing function of $\|u_0\|_{B^{s_1}_{p, q}}$ with $T=\infty$ if $\|u_0\|_{B^{s_1}_{p, q}}$ is sufficiently small. \end{theorem} Incorporating the additional constraints from the energy bound in \cite{kazuo}, we can now state the global existence result. \begin{corollary} \label{coro1.1} Let $p=2$ and let $n=3$. Then, for any of our local existence results, if we additionally assume that \begin{gather*} 2\gamma_1+\gamma_2\geq 5, \\ \int_1^\infty \frac{ds}{s g_1^2(s)g_2(s)} =\infty, \end{gather*} then the local solutions can be extended to global solutions. \end{corollary} Note that if $g_1$ and $g_2$ are Mikhlin multipliers, then all of the constraints on $g_1$ and $g_2$ are satisfied. Also, if $g_1$ and $g_2$ are logarithms, then the corollary extends the appropriate local solutions from Theorem \ref{old style short log} and \ref{special case 1 log} to global solutions. The corollary follows directly from the smoothing effect of the operator $e^{t\mathcal{L}_1}$, which ensures that, for any $t>0$, our local solution $u(t,\cdot)\in B^r_{2,q}(\mathbb{R}^3)$ for any $r\in \mathbb{R}$. This provides the smoothness necessary to use the energy bound from \cite{kazuo} to obtain a uniform-in-time bound on the $B^{s_1}_{2,q}(\mathbb{R}^3)$ norm of the solution, and then a standard bootstrapping argument completes the proof of global existence. In Section \ref{Higher regularity for the local existence result}, we include an argument detailing this smoothing effect for the solution to Theorem \ref{old style short}. Finally, we remark that extending the local solutions to global solutions for $p\neq 2$ and $n>3$ will be the subject of future work. Handling $n>3$ should follow by tweaking the argument used in \cite{kazuo}. Obtaining global solutions for $p\neq 2$ is significantly more complicated, and the argument will follow the interpolation based argument used by Gallagher and Planchon \cite{galplan} for the two dimensional Navier-Stokes equation. \section{Besov spaces}\label{Besov space} We begin by defining the Besov spaces $B^s_{p,q}(\mathbb{R}^n)$. Let $\psi_0$ be an even, radial, Schwartz function with Fourier transform $\hat{\psi_0}$ that has the following properties: \begin{gather*} \hat{\psi_0}(x)\geq 0,\\ \operatorname{support}\hat{\psi_0}\subset A_0 :=\{\xi\in \mathbb{R}^n:2^{-1}<|\xi|<2\},\\ \sum_{j\in\mathbb{Z}} \hat{\psi_0}(2^{-j}\xi)=1, \quad \text{for all } \xi\neq 0. \end{gather*} We then define $\hat{\psi_j}(\xi)=\hat{\psi}_0(2^{-j}\xi)$ (from Fourier inversion, this also means $\psi_j(x)=2^{jn}\psi_0(2^jx)$), and remark that $\hat{\psi_j}$ is supported in $A_j:=\{\xi\in\mathbb{R}^n:2^{j-1}<|\xi|<2^{j+1}\}$. We also define $\Psi$ by \begin{equation}\label{low freq part} \hat{\Psi}(\xi)=1-\sum_{k=0}^\infty \hat{\psi}_k(\xi). \end{equation} We define the Littlewood Paley operators $\Delta_j$ and $S_j$ by \[ \Delta_j f=\psi_j\ast f, \quad S_jf=\sum_{k=-\infty}^{j}\Delta_k f, \] and record some properties of these operators. Applying the Fourier Transform and recalling that $\hat{\psi}_j$ is supported on $2^{j-1}\leq |\xi|\leq2^{j+1}$, it follows that \begin{equation}\label{besovlemma1} \begin{gathered} \Delta_j\Delta_k f= 0, \quad |j-k|\geq 2, \\ \Delta_j (S_{k-3}f\Delta_{k}g)= 0 \quad |j-k|\geq 4, \end{gathered} \end{equation} and, if $|i-k|\leq 2$, then \begin{equation}\label{besovpieces67} \Delta_j(\Delta_kf\Delta_i g)=0 \quad j>k+4. \end{equation} For $s\in\mathbb{R}$ and $1\leq p,q\leq \infty$ we define the space $\tilde{B}^s_{p,q}(\mathbb{R}^n)$ to be the set of distributions such that \[ \|u\|_{\tilde{B}^s_{p,q}}=\Big(\sum_{j=0}^\infty (2^{js}\|\Delta_j u\|_{L^p})^q\Big)^{1/q}<\infty, \] with the usual modification when $q=\infty$. Finally, we define the Besov spaces $B^s_{p,q}(\mathbb{R}^n)$ by the norm \[ \|f\|_{B^s_{p,q}}=\|\Psi*f\|_p+\|f\|_{\tilde{B}^s_{p,q}}, \] for $s>0$. For $s>0$, we define $B^{-s}_{p',q'}$ to be the dual of the space $B^s_{p,q}$, where $p',q'$ are the Holder-conjugates to $p,q$. These Littlewood-Paley operators are also used to define Bony's paraproduct. We have \begin{equation}\label{lp start} fg=\sum_{k} S_{k-3}f\Delta_k g + \sum_{k}S_{k-3}g\Delta_k f + \sum_{k}\Delta_k f\sum_{l=-2}^2 \Delta_{k+l} g. \end{equation} The estimates \eqref{besovlemma1} and \eqref{besovpieces67} imply that \begin{equation} \label{bony256} \begin{aligned} \Delta_j (fg)&= \sum_{k=-3}^3 \Delta_j (S_{j+k-3}f\Delta_{j+k} g) + \sum_{k=-3}^3 \Delta_j (S_{j+k-3}g\Delta_{j+k} f)\\ &\quad +\sum_{k>j-4}\Delta_j \Big(\Delta_k f\sum_{l=-2}^2 \Delta_{k+l}g\Big). \end{aligned} \end{equation} Now we turn our attention to establishing some basic Besov space estimates. First, we let $1\leq q_1\leq q_2\leq \infty$, $\beta_1\leq \beta_2$, $1\leq p_1\leq p_2\leq\infty$, $\alpha>0$, and set $\tilde{p}=np/(n-\alpha p)$ with $\alpha0$, we have \begin{equation} \|f\|_{L^{\tilde{p}}}\leq \|f\|_{B^\varepsilon_{\tilde{p},q}} \leq \|f\|_{B^{\alpha+\varepsilon}_{p,q}}, \end{equation} where we used the definition of Besov spaces for the first inequality and \eqref{besov embedding3} for the second. Next we record our two Leibnitz-rule type estimate. The first is the standard estimate, which can be found in (among many other places) \cite[Lemma 2.2]{chae}. See also \cite[Proposition 1.1]{TT}. \begin{proposition}\label{product est 1} Let $s>0$ and $q\in [1,\infty]$. Then \[ \|fg\|_{B^s_{p,q}}\leq C\big(\|f\|_{L^{p_1}}\|g\|_{B^s_{p_2,q}} +\|f\|_{B^s_{q_1,q}}\|g\|_{L^{q_2}}\big), \] where $1/p=1/p_1+1/p_2=1/q_1+1/q_2$ and $p_i, q_i \in [1,\infty]$ for $i=1,2$. \end{proposition} Our second product estimate is less common. The estimate originated in \cite{chemin}; another proof can be found in \cite{besovpaper2}. \begin{proposition}\label{product est 2} Let $f\in B^{s_1}_{p_1,q}(\mathbb{R}^n)$ and let $g\in B^{s_2}_{p_2,q}(\mathbb{R}^n)$. Then, for any $p$ such that $1/p\leq 1/p_1+1/p_2$ and with $s=s_1+s_2-n(1/p_1+1/p_2-1/p)$, we have \[ \|fg\|_{B^s_{p,q}}\leq C\|f\|_{B^{s_1}_{p_1,q}}\|g\|_{B^{s_2}_{p_2,q}}, \] provided $s_10$. \end{proposition} \section{Local existence by Proposition \ref{product est 1}}\label{type 1} Theorem \ref{old style short} and Theorem \ref{old style short log} are both proven using the standard product estimate (Proposition \ref{product est 1}). These theorems are both special cases of more general theorems, and the primary task of this section is to prove the theorem which implies Theorem \ref{old style short log}. There is a similar result associated with Theorem \ref{old style short}, and it will be discussed at the end of the section. \begin{theorem}\label{full version 2 log} Let $\gamma_1>1$, $\gamma_2>0$, $q\geq 1$, and $p\geq 2$. Assume $g_1$ and $g_2$ satisfy $|g_i(r)|\leq C(1+r)^\delta$ for any $\delta>0$ and $|g_i^{(k)}(r)|\leq Cr^{-k}$ for $1\leq k\leq n/2+1$. Let $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$ be divergence-free. Then there exists a unique local solution $u$ to the generalized Leray-alpha equation \eqref{leray1}, with \[ u\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}^T_{a;s_2,p,q}, \] where $a=(s_2-s_1)/(\gamma_1-\varepsilon)$ for any sufficiently small $\varepsilon>0$ if there exists $k>0$ such that the parameters satisfy \eqref{full 2 list}. $T$ can be chosen to be a non-increasing function of $\|u_0\|_{B^{s_1}_{p, q}}$ with $T=\infty$ if $\|u_0\|_{B^{s_1}_{p, q}}$ is sufficiently small. \end{theorem} We begin by re-writing equation \eqref{leray1} as \begin{equation} \label{leray2} \begin{gathered} \partial_t u+P(1-\alpha^2\mathcal{L}_2)^{-1}\operatorname{div}(u\otimes (1-\alpha^2\mathcal{L}_2)u)-\nu\mathcal{L}_1 u=0, \\ u(0,x)=u_0(x), \quad \operatorname{div}u_0=\operatorname{div}u=0, \end{gathered} \end{equation} where $P$ is the Hodge projection onto divergence free vector fields and an application of the divergence free condition shows $\nabla_u (1-\alpha^2 \mathcal{L}_2) u=\operatorname{div}(u\otimes (1-\alpha^2\mathcal{L}_2)u)$, where $v\otimes w$ is the matrix with $ij$ entry equal to the product of the $i^{th}$ coordinate of $v$ and the $j^{th}$ coordinate of $w$. Setting $\alpha=1$ and $\nu=1$ for notational simplicity and applying Duhamel's principle, we obtain that $u$ is a solution to the equation if and only if $u$ is a fixed point of the map $\Phi$ given by \[ \Phi(u)=e^{t\mathcal{L}_1}u_0+\int_0^t e^{(t-s)\mathcal{L}_1}(W(u(s)))ds, \] where $W(u,v)=-P(1-\mathcal{L}_2)^{-1}\operatorname{div}(u(s)\otimes (1-\mathcal{L}_2)v(s))$. To simplify notation, we will also set $W(u,u)=W(u)$. Our goal is to show that $\Phi$ is a contraction in the space \begin{align*} X_{T,M}=\Big\{&f\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}_{a;s_2,p,q} \text{and} \\ &\sup_t \|f(t)-e^{t\mathcal{L}_1}u_0\|_{B^{s_1}_{p,q}} +\sup_t t^{a}\|u(t)\|_{B^{s_2}_{p,q}}0$ to be chosen later. Following the arguments outlined in \cite{KP} and \cite{sobpaper}, $\Phi$ will be a contraction if we can show that \begin{equation}\label{main parts 1} \begin{gathered} \sup_{t}t^{a}\|e^{t\mathcal{L}_1}u_0\|_{B^{s_2}_{p,q}}0$. To complete the argument, we need to bound this by $\|u\|_{B^{s_2}_{p,q}}^2$. To facilitate this, we define $\varepsilon$ by setting $\gamma_2-\gamma_2^-=\varepsilon$, choose $r+1+\varepsilon=s_2$ (which forces $s_2>\gamma_2$), and $q_2=p_2=p$ (which forces $p_1=q_1$). Applying these choices and using the Besov embedding \eqref{besov embedding1}, inequality \eqref{2mal} becomes \begin{equation}\label{2mali} \begin{aligned} \|W(u(s))\|_{B^{r}_{p^*,q}} &\leq C\|u\|_{L^{p_1}}\|u\|_{B^{s_2}_{p_2,q}} +C\|u\|_{B^{s_2-\gamma_2}_{q_1,q}}\|u\|_{B^{\gamma_2^+}_{q_2,q}}\\ &\leq C\|u\|_{B^{s_2}_{p,q}}\big(\|u\|_{L^{p_1}}+\|u\|_{B^{s_2-\gamma_2}_{p_1,q}}\big). \end{aligned} \end{equation} Finally, choosing $p_1=np/(n-kp)$ for some $k\gamma_2, \quad s_2=r+1+\varepsilon, \\ k\leq \gamma_2, \quad kp2(s_2-s_1)/(\gamma_1^-)\\ 0\leq -(s_1-r+n/p^*-n/p)/(\gamma_1^-)-2(s_2-s_1)/(\gamma_1^-)+1. \end{gathered} \end{equation} The first inequality in this list ensures that the $|t-s|$ term is integrable as $s$ goes to $t$, the second inequality does the same for the $s^{-2a}$ term as $s$ goes to $0$, and the last inequality makes the power on the post-integration $t$ positive. The last inequality follows by recalling that $T\leq 1$ and by choosing a sufficiently small $M$. For the last term in \eqref{main parts 1}, a similar argument gives \begin{align*} &\sup_{t}t^{a}\|\int_0^t e^{(t-s){\mathcal{L}_1}} W(u(s))ds\|_{B^{s_2}_{p,q}}\\ &\leq \sup_{t}t^{a}\int_0^t |t-s|^{-(s_2-r+n/p^*-n/p)/(\gamma_1^-)} \|W(u(s))\|_{B^{r}_{p^*,q}}ds \\ &\leq \sup_{t}t^{a}\int_0^t |t-s|^{-(s_2-r+n/p^*-n/p)/(\gamma_1^-)}s^{-2a} s^{2a}\|u(s)\|^2_{B^{s_2}_{p,q}}ds \\ &\leq C\|u\|^2_{a;s_2, p, q}\sup_t t^{a}t^{-(s_2-r+n/p^*-n/p) /(\gamma_1^-)-2(s_2-s_1)/(\gamma_1^-)+1} \\ &\leq CM^2T^{-(s_2-r)/(\gamma_1^-)-(s_2-s_1)/(\gamma_1^-)+1}2(s_2-s_1)/(\gamma_1^-), \\ 0\leq -(s_2-r+n/p^*-n/p)/(\gamma_1^-)-(s_2-s_1)/(\gamma_1^-)+1. \end{gathered} \end{equation} Combining \eqref{2ty1}, and \eqref{2ty2} (and removing redundancies) gives \begin{equation}\label{pac 1} \begin{gathered} s_1>r, \\ (\gamma_1^-)/2>s_2-s_1,\\ 0\leq s_2-r+n/p^*-n/p<(\gamma_1^-),\\ (\gamma_1^-)\geq 2s_2-r+n/p^*-n/p-s_1. \end{gathered} \end{equation} Incorporating \eqref{product 1 term}, and observing that, since $s_2>s_1$, the last inequality in \eqref{pac 1} implies the third inequality, we obtain \begin{equation}\label{full 2 list} \begin{gathered} s_2>\gamma_2\geq k,\\ kp 0$). To get Theorem \ref{old style short} or Theorem \ref{old style short log}, choose $k$ to be an arbitrarily small positive number, which removes $k$ from the last inequality and forces $\gamma_2>0$. \section{Local existence using Proposition \ref{product est 2}}\label{type 2} In this section we prove the following local existence result, which implies Theorem \ref{special case 1}. We address Theorem \ref{special case 1 log} at the end of the section. \begin{theorem}\label{full generality} Let $\gamma_1>1$, $\gamma_2>0$, $q\geq 1$, $p\geq 2$, and assume $g_1$ and $g_2$ satisfy the Mikhlin condition (see inequality \eqref{mikhlin condition don}). Let $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$ be divergence-free. Then there exists a unique local solution $u$ to the generalized Leray-alpha equation \eqref{leray1}, with \[ u\in BC([0,T):B^{s_1}_{p,q}(\mathbb{R}^n))\cap \dot{C}^T_{a;s_2,p,q}, \] where $a=(s_2-s_1)/\gamma_1$, if there exists $r$, $r_1$ and $r_2$ such that all the parameters satisfy \eqref{final list 2}. $T$ can be chosen to be a non-increasing function of $\|u_0\|_{B^{s_1}_{p, q}}$ with $T=\infty$ if $\|u_0\|_{B^{s_1}_{p, q}}$ is sufficiently small. \end{theorem} With the same set-up as the previous section, our goal is to show that \begin{equation}\label{2main parts 1} \begin{gathered} \sup_{t}t^{a}\|e^{t\mathcal{L}_1}u_0\|_{B^{s_2}_{p,q}}0, \\ r_1, r_22(s_2-s_1)/\gamma_1,\\ 0\leq -(s_1-r)/\gamma_1-2(s_2-s_1)/\gamma_1+1. \end{gathered} \end{equation} Estimating the last term of \eqref{main parts 1} in a similar fashion, we have \begin{equation} \begin{aligned} &\sup_{t}t^{a}\|\int_0^t e^{(t-s){\mathcal{L}_1}} W(u(s))ds\|_{B^{s_2}_{p,q}}\\ &\leq \sup_{t}t^{a}\int_0^t |t-s|^{-(s_2-r)/\gamma_1} \|W(u(s))\|_{B^{r}_{p,q}}ds\\ &\leq \sup_{t}t^{a}\int_0^t |t-s|^{-(s_2-r)/ \gamma_1}s^{-2a}s^{2a}\|u(s)\|^2_{B^{s_2}_{p,q}}ds\\ &\leq C\|u\|^2_{a;s_2,p,q}\sup_t t^{a}t^{-(s_2-r)/\gamma_1-2(s_2-s_1)/\gamma_1+1}\\ &\leq CM^2T^{-(s_2-r)/\gamma_1-(s_2-s_1)/\gamma_1+1}2(s_2-s_1)/\gamma_1,\\ 0\leq -(s_2-r)/\gamma_1-(s_2-s_1)/\gamma_1+1. \end{gathered} \end{equation} Our final task is to unify the conditions on the parameters. The sets of inequalities from equations \eqref{ty1} and \eqref{ty2} can be simplified to \begin{equation} \begin{gathered} 0s_2-\gamma_1,\\ \gamma_1\geq (s_2-s_1)+(s_2-r). \end{gathered} \end{equation} Incorporating the inequalities from \eqref{aldi}, we have \begin{equation}\label{vague list} \begin{gathered} 0s_2-\gamma_1,\\ \gamma_1\geq (s_2-s_1)+(s_2-r),\\ r+1-\gamma_2\leq r_1+r_2-n/p,\\ r_1+r_2>0,\\ r_1, r_20$ and $r_1, r_2s_2-\gamma_1,\\ \gamma_1\geq (s_2-s_1)+(s_2-r),\\ r=-1+\gamma_2+R-n/p, \\ n/p>\gamma_2/2+R, \\ s_2\geq \gamma_2/2+R. \end{gathered} \end{equation} Using the fourth line to eliminate $r$ from the other inequalities, and then removing extraneous inequalities, we finally get \begin{equation}\label{final list 2} \begin{gathered} 0\gamma_2/2+R, \\ s_2\geq \gamma_2/2+R. \end{gathered} \end{equation} Eliminating the free parameter $R$ weakens this to \begin{equation}\label{final list 3} \begin{gathered} 0\gamma_2-n/p-1,\\ \gamma_1\geq 2s_2-s_1-\gamma_2+n/p+1,\\ n/p>\gamma_2/2, \\ s_2\geq \gamma_2/2, \end{gathered} \end{equation} which finishes Theorem \ref{special case 1}. The analog of Theorem \ref{special case 1} for logarithmic $g$ is obtained by replacing $\gamma_1$ and $\gamma_2$ with $\gamma_1^-$ and $\gamma_2^-$. \section{Operator estimates for $\mathcal{L}^g_\gamma$}\label{L operator estimates} In this section, we define the Fourier multiplier $\mathcal{L}^g_\gamma$ by \[ \mathcal{L}^g_\gamma u(x)=\int -\frac{|\xi|^{\gamma}}{g(|\xi|)} \hat{u}(\xi)e^{ix\cdot \xi}d\xi, \] where $\gamma\in \mathbb{R}$ and $g:\mathbb{R}\to \mathbb{R}$ is radial, nondecreasing, and bounded below by 1. Note that if we define $G$ to be the Fourier multiplier with symbol $1/g$, then $L^g_\gamma=-G(-\Delta)^{\gamma/2}$. The goal here is to prove operator estimates for $\mathcal{L}^g_\gamma$, and we begin by stating the Mikhlin multiplier theorem, which will be referenced often in this section. \begin{theorem}[Mikhlin multiplier theorem] Let $M$ be an operator with symbol $m:\mathbb{R}^n\to \mathbb{R}^n$. If $|x|^k|\nabla^k m(x)|$ is bounded for all $0\leq k\leq n/2+1$, then $M$ is an $L^p(\mathbb{R}^n)$ multiplier for all $10$ and $|g^{(k)}(r)|\leq Cr^{-k}$ for all $1\leq k$. These observations inform the next proposition. \begin{proposition}\label{annie2} Let $10$, and assume $|g^{(k)}(r)|\leq Cr^{-k}$ for all $1\leq k\leq n/2+1$. Then $(1-\mathcal{L}^g_\gamma)^{-1}: B^{s_1-(\gamma-\varepsilon)}_{p,q} (\mathbb{R}^n)\rightarrow B^{s_1}_{p,q}(\mathbb{R}^n)$ for any small $\varepsilon>0$, with \[ \|(1-\mathcal{L}^g_\gamma)^{-1} f\|_{B^{s_1}_{p,q}} \leq C\|f\|_{B^{s_1-(\gamma-\varepsilon)}_{p,q}}. \] \end{proposition} \begin{proof} As in the previous proposition, this follows by showing that the symbol $\frac{(1+r^2)^{(\gamma-\varepsilon)/2}}{1+r^\gamma/g(r)}$ is a Mikhlin multiplier. First, we re-write this as $\frac{g(r)}{(1+r^2)^{\varepsilon/2}} \frac{(1+r^2)^{\gamma/2}}{g(r)+r^\gamma}$. That each of these terms individually satisfies the Mikhlin condition follows directly from the assumptions on $g$. \end{proof} We remark that the $\varepsilon$ loss between these two results is due to the necessity of controlling the growth of the $g(r)$ term. \section{Operator estimates for $e^{t\mathcal{L}}$}\label{operator est} As in the previous section, we define the Fourier multiplier $\mathcal{L}^g_\gamma$ by \[ \mathcal{L}^g_\gamma u(x) =\int -\frac{|\xi|^{\gamma}}{g(|\xi|)}\hat{u}(\xi)e^{ix\cdot \xi}d\xi, \] where $\gamma\in \mathbb{R}$ and $g:\mathbb{R}\to \mathbb{R}$ is nondecreasing and bounded below by 1. We define the operator $e^{t\mathcal{L}^g_\gamma}$ to be the Fourier multiplier with symbol $e^{-t|\xi|^\gamma/g(|\xi|)}$. The goal of this section is to establish operator bounds for $e^{t\mathcal{L}^g_\gamma}$ in the case where $\gamma>1$, and following the general outline of the same task for $e^{t\Delta}$, we need to first establish $L^p-L^q$ boundedness for the operator. Also note that, throughout the section, we will assume $01$. Assume $|g(r)|\leq C(1+r)^\delta$ for any $\delta>0$ and assume \begin{equation}\label{mik condition 65} |g^{(k)}(r)|\leq Cr^{-k} \end{equation} holds for $1\leq k\leq n/2+1$. Then \[ e^{t\mathcal{L}^g_\gamma}:L^p(\mathbb{R}^n)\to L^p(\mathbb{R}^n), \] and we have the bound \begin{equation}\label{mal2} \|e^{t\mathcal{L}^g_\gamma}f\|_{L^p}\leq C\|f\|_{L^p}. \end{equation} \end{proposition} \begin{proof} We will show that $e^{-r^\gamma/g(r)}$ satisfies the Mikhlin condition, and then the result follows by the Mikhlin multiplier theorem. First, we observe that the multiplier is clearly bounded. Then \[ \big|\frac{d}{dr}e^{-r^\gamma/g(r)} \big| \leq C\Big(\frac{r^{\gamma-1}}{g(r)}+\frac{r^\gamma g'(r)}{g(r)^2}\Big) e^{-r^\gamma/g(r)}\leq Cr^{\gamma-1}e^{-Cr^{\gamma-\delta}}\leq Cr^{-1}, \] where in the last inequality we used that $\gamma>1$ and we have chosen $\delta$ to be a small positive number. Similar calculations hold for the remaining derivatives. \end{proof} Now we consider the case where $p\neq q$. \begin{proposition}\label{lpsemi} Let $1\leq p< q\leq \infty$, and assume $\gamma>1$. Then \[ e^{t\mathcal{L}^g_\gamma}:L^p(\mathbb{R}^n)\to L^q(\mathbb{R}^n), \] and we have the bound \begin{equation}\label{mal2b} \|e^{t\mathcal{L}^g_\gamma}f\|_{L^q}\leq Ct^{-(n/p-n/q)/\gamma}\|f\|_{L^p}, \end{equation} provided \begin{equation}\label{mal condition} \sup_{t\in (0, 1)} \int_{0}^\infty r^{n-1} | \partial_r^{(n)}(e^{-r^\gamma/g(rt^{-1/\gamma})}) | dr<\infty. \end{equation} \end{proposition} \begin{proof} For notational convenience, we will suppress the subscript $\gamma$ and the superscript $g$ on the operator for the duration of the proof. Setting $e^{t\mathcal{L}}f=e^{t\mathcal{L}}\delta\ast f$, where the Fourier Transform of $e^{t\mathcal{L}}\delta(x)$ is equal to $e^{-t|\xi|^\gamma/g(|x|)}$, and applying Young's inequality, we obtain that \[ \|e^{t\mathcal{L}}f\|_{L^q}\leq \|e^{t\mathcal{L}}\delta\|_{L^r}\|f\|_{L^p}, \] where $1+1/q=1/r+1/p$. Formally, we have that \[ e^{t\mathcal{L}}\delta(\xi) =C\int_{\mathbb{R}^n}e^{-t|x|^\gamma/g(|x|)}e^{ix\cdot\xi}dx. \] Making the variable change $x\to t^{-1/\gamma}x$, we obtain \[ e^{t\mathcal{L}}\delta(\xi)=Ct^{-n/\gamma}\int_{\mathbb{R}^n} e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} e^{it^{-1/\gamma}x\cdot\xi}dx. \] Taking the $L^r(\mathbb{R}^n)$ norm gives \[ \|e^{t\mathcal{L}}\delta\|_{L^r}=Ct^{-n/\gamma}\Big(\int_{\mathbb{R}^n} \Big| \int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} e^{it^{-1/\gamma}x\cdot\xi}dx \Big|^rd\xi\Big)^{1/r}. \] Making the variable change $\xi\to t^{-1/\gamma}\xi$, this finally becomes \[ \|e^{t\mathcal{L}}\delta\|_{L^r}=Ct^{-n/\gamma+n/(r\gamma)} \Big(\int_{\mathbb{R}^n}\Big|\int_{\mathbb{R}^n}e^{-|x|^\gamma/ g(t^{-1/\gamma}|x|)} e^{ix\cdot\xi}dx \Big|^rd\xi\Big)^{1/r}. \] Since $1-1/r=1/p-1/q$, it only remains to obtain a $t$-independent bound for the integral. First, we have \begin{align*} &\int_{\mathbb{R}^n}\Big| \int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} e^{ix\cdot\xi}dx \Big|^rd\xi \\ &\leq \int_{|\xi|<1}\Big|\int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} e^{ix\cdot\xi}dx \Big|^rd\xi \\ &\quad + \int_{|\xi|>1}\Big|\int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} e^{ix\cdot\xi}dx \Big|^rd\xi \\ &\leq C\int_{|\xi|>1}\Big|\int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} dx \Big|^r \\ &\quad +\int_{|\xi|>1}\Big|\int_{\mathbb{R}^n}e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)} \frac{\partial_{x_1}\partial_{x_2}\dots \partial_{x_n}}{\xi_1\xi_2\dots \xi_n}e^{ix\cdot\xi}dx \Big|^rd\xi \\ &\leq C+\int_{|\xi|>1}\frac{1}{|\xi_1|^r|\xi_2|^r\dots |\xi_n|^r} \Big|\int_{\mathbb{R}^n}\partial_{x_1}\partial_{x_2}\dots \partial_{x_n}(e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)}) e^{ix\cdot\xi}dx \Big|^rd\xi\\ &\leq C+C\Big(\int_{\mathbb{R}^n} \Big| \partial_{x_1}\partial_{x_2} \dots \partial_{x_n}(e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)}) \Big| dx \Big)^r, \end{align*} where the last inequality follows from observing that since $p1$, and therefore the integral in each $\xi$ coordinate is finite. To compute the final integral, we convert to polar coordinates, and remark that since $e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)}$ is radial, we can disregard the radial portion of each coordinate derivative's expression in polar form, and we have \[ \int_{\mathbb{R}^n} \Big| \partial_{x_1}\partial_{x_2} \dots \partial_{x_n}(e^{-|x|^\gamma/g(t^{-1/\gamma}|x|)}) \Big| dx \leq C\int_{0}^\infty r^{n-1}\Big| \partial_r^{(n)} (e^{-r^\gamma/g(rt^{-1/\gamma})}) \Big| dr, \] which is \eqref{mal condition}, and thus completes the proposition. \end{proof} Our next task is to establish sufficiency criteria for \eqref{mal condition}. \begin{proposition}\label{rip city 5} If \begin{equation}\label{rip city}|g^{(k)}(r)|\leq Cr^{-k}, \end{equation} for all $0\leq k\leq n$, then \eqref{mal condition} holds. \end{proposition} \begin{proof} We begin with the special case where $g$ is a constant function, and without loss of generality assume $g(r)=1$. Then \eqref{mal condition} becomes \[ \int_{0}^\infty r^{n-1}\big| \partial_r^{(n)}(e^{-r^\gamma})\big|\, dr. \] Computing the derivatives through repeated use of the product rule will yield the addition of $2^{n-1}$ terms of the form $r^\alpha e^{-r^\gamma}$. The decay provided by the exponential term makes each of these integrable for large $r$. For the region where $r$ is small, the most singular term is of the form $r^{n-1}r^{n(\gamma-n)}e^{-r^\gamma}$, which is integrable for small $r$ because $\gamma>1$, and this finishes the argument for the special case where $g$ is constant. If $\gamma>1$ and $g$ is not constant, the proof is similar, but more combinatorially intense. To see this, we first consider the case where $n=1$. Then we have \begin{align*} &\int_{0}^\infty \big| \partial_r(e^{-r^\gamma/g(rt^{-1/\gamma})})\big| \, dr\\ &=\int_0^\infty \Big(\frac{Cr^{\gamma-1}}{g(rt^{-1/\gamma})} +C\frac{r^\gamma g'(rt^{-1/\gamma})t^{-1/\gamma}}{(g(rt^{-1/\gamma}))^2}\Big) e^{-r^\gamma/g(rt^{-1/\gamma})} dr. \end{align*} Using our assumption on $g$, we have \begin{align*} &\int_0^\infty \Big(\frac{Cr^{\gamma-1}}{g(rt^{-1/\gamma})} +C\frac{r^\gamma g'(rt^{-1/\gamma})t^{-1/\gamma}}{(g(rt^{-1/\gamma}))^2}\Big) e^{-r^\gamma/g(rt^{-1/\gamma})} dr.\\ &\leq C\int_0^\infty \Big(\frac{Cr^{\gamma-1}}{g(rt^{-1/\gamma})} +C\frac{r^\gamma t^{-1/\gamma}}{rt^{-1/\gamma} (g(rt^{-1/\gamma}))^2}\Big) e^{-r^\gamma/g(rt^{-1/\gamma})} dr. \end{align*} Finally, since $g$ is bounded below by one and is bounded above by assumption, we have \[ \int_0^\infty \Big(\frac{Cr^{\gamma-1}}{g(rt^{-1/\gamma})} +C\frac{r^\gamma t^{-1/\gamma}}{rt^{-1/\gamma} (g(rt^{-1/\gamma}))^2}\Big) e^{-r^\gamma/g(rt^{-1/\gamma})} dr \leq C\int_0^\infty r^{\gamma-1}e^{-Cr^\gamma}dr, \] which reduces the problem to the constant $g$ case. In general, when computing the derivatives, either the derivatives fall on the $r^\gamma$ term (and then since $g$ is bounded below, we are back in constant $g$ case) or the derivatives act on the $g(rt^{-1/\gamma})$ term. In that case, because of \eqref{rip city}, the derivatives of $g$ introduce exactly as much decay in $r$ as occurs when differentiating $r^\gamma$, and the $t$ dependent terms cancel out, so this case also reduces to a term from the constant $g$ case. \end{proof} Recall from the introduction that we ultimately seek results for $g$ of the form $g(r)=\ln(a+r)$, where $\ln(a)>1$. Since such a $g$ is unbounded, it does not satisfy the requirements of the previous proposition. The following result adapts that argument to this particular case. \begin{proposition} Let $g(r)\leq C(1+r)^\delta$ for any $\delta>0$, and assume that \[ |g^{(k)}(r)|\leq C|r|^{-k}, \] for $1\leq k\leq n$. Then \begin{equation}\label{rip city 3} \|e^{t\mathcal{L}^g_\gamma}f\|_{L^q} \leq Ct^{-(n/p-n/q)/(\gamma-\varepsilon)}\|f\|_{L^p}, \end{equation} for any small $\varepsilon>0$, provided $0-1$. Since $\alpha>-1$, we have \begin{align*} \int_0^\infty r^{\alpha}e^{-r^\gamma/g(rt^{-1/\gamma})} dr &\leq \int_0^1 r^{\alpha}e^{-r^\gamma/g(rt^{-1/\gamma})} dr +\int_1^\infty r^{\alpha}e^{-r^\gamma/g(rt^{-1/\gamma})} dr\\ &\leq C+\int_1^\infty r^{\alpha}e^{-r^\gamma/(1+rt^{-1/\gamma})^\delta} dr\\ &\leq C+\int_1^\infty r^{\alpha}e^{-Cr^{\gamma-\delta}t^{\delta/\gamma}} dr, \end{align*} where we used the assumption on $g$ and $\delta$ is small positive number to be specified later. Making the variable change $r\to rt^{\delta/(\gamma(\gamma-\delta))}$, we finally obtain \[ \int_1^\infty r^{\alpha}e^{-Cr^\gamma/g(rt^{-1/\gamma})} dr \leq C+t^{-\delta\alpha/\gamma(\gamma-\delta)} \int_1^\infty r^\alpha e^{-Cr^{\gamma-\delta}} dr \leq C(1+t^{-\delta\alpha/\gamma(\gamma-\delta)}), \] provided $\gamma-\delta>1$. Plugging back into \eqref{rip city 11}, we have \[ \|e^{t\mathcal{L}^g_\gamma}f\|_{L^q} \leq Ct^{-(n/p-n/q)/\gamma-\delta\alpha/\gamma(\gamma-\delta)}\|f\|_{L^p}. \] Choosing a sufficiently small $\delta$ finishes the proposition. \end{proof} Before moving on, we remark that, as in the previous section, adapting to the logarithmic $g$ comes at a cost of additional singularity in the time variable. Now that we have established $L^p-L^q$ boundedness for $e^{t\mathcal{L}^g_\gamma}$, we turn our attention to incorporating changes in regularity. We will rely heavily on the following, which is \cite[Proposition 7.2 Chapter 13]{T3}. \begin{proposition}\label{taylor prop} Let $e^{tA}$ be a holomorphic semigroup on a Banach space $X$. Then, for $t>0$, \[ \|Ae^{tA}f\|_X\leq \frac{C}{t}\|f\|_X, \] for $00$ and let $|g^{(k)}(r)|\leq C|r|^{-k}$ for all $1\leq k\leq n/2+1$. Then $e^{t\mathcal{L}^g_\gamma}: B^{s_1}_{p,q}(\mathbb{R}^n)\to B^{s_2}_{p,q}(\mathbb{R}^n)$ and \begin{equation} \|e^{t\mathcal{L}^g_\gamma} f\|_{B^{s_2}_{p,q}} \leq t^{-(s_2-s_1)/(\gamma-\varepsilon)}\|f\|_{B^{s_1}_{p,q}}, \end{equation} for any $\varepsilon>0$. \end{proposition} \begin{proof} As in the previous proposition, we let $s_2=\gamma-\varepsilon$ and set $s_1=0$, and the rest of the argument will follow as before provided we show that the Fourier multiplier with symbol \[ m(\xi)=\frac{|\xi|^{\gamma-\varepsilon}g(\xi)}{|\xi|^\gamma}, \] and with support in the annulus $|\xi|\geq 1/2$, is bounded on $L^p(\mathbb{R}^n)$. As this follows from the assumptions on $g$, the proof is complete. \end{proof} The following is direct combination of Proposition \ref{rip city 5} and Proposition \ref{sobolev}. \begin{proposition}\label{heat kernel bound} Let $10$, and let $|g^{(k)}(r)|\leq C|r|^{-k}$ for all $1\leq k\leq n/2+1$. Then $e^{t\mathcal{L}^g_\gamma}: B^{s_1}_{p_1,q}(\mathbb{R}^n)\to B^{s_2}_{p_2,q}(\mathbb{R}^n)$ and \begin{equation} \|e^{t\mathcal{L}^g_\gamma} f\|_{B^{s_2}_{p_2,q}} \leq t^{-(s_2-s_1+n/p_1-n/p_2)/(\gamma-\varepsilon)}\|f\|_{B^{s_1}_{p_1,q}}, \end{equation} for any small $\varepsilon>0$. \end{proposition} We remark that these results also apply to Sobolev spaces (see \cite[Section 2]{besovpaper2} for an example of a similar process applied to the standard heat kernel). \section{Higher regularity for the local existence result} \label{Higher regularity for the local existence result} As was mentioned in the introduction, the solutions to the generalized Leray-alpha equations constructed here are smooth for all $t>0$ at which the solution exists. In this section we prove that the solutions to Theorem \ref{old style short} have this additional regularity and quantify the blow-up that occurs in these higher regularity norms as $t\to 0$. We use an induction argument inspired by the results in \cite{katoinduction} for the Navier-Stokes equation. We remark that similar results can be proven for the other theorems in this paper, but require different (and in some cases much more involved) arguments. \begin{proposition}\label{higher regularity theorem} Let $u_0\in B^{s_1}_{p,q}(\mathbb{R}^n)$ be divergence-free. Let $u$ be a solution to the generalized Leray-alpha equation \eqref{leray1} given by Theorem \ref{old style short}. Then for all $r\geq s_1$ we have that $u\in \dot{C}^T_{(r-s_1)/\gamma_1;r,p,q}$. \end{proposition} Before starting the proof, recall from Theorem \ref{old style short} that $s_1>0$ and that \begin{equation}\label{durant} \begin{gathered} \gamma_1>1, \quad \gamma_2>0, \quad s_2>\gamma_2,\\ s_2-s_1<\min\{\gamma_1/2, 1\},\\ \gamma_1\geq s_2-s_1+n/p+1. \end{gathered} \end{equation} \begin{proof} We start with the solution $u$ given by Theorem \ref{old style short}. Then let $\delta>0$ be arbitrary and let $v=t^\delta u$. We note that $v(0)=v_0=0$. Then \begin{align*} \partial_t v &=\delta t^{\delta-1} u+t^\delta \partial_t u \\ &=\delta t^{-1} v+t^\delta P(\mathcal{L}_1 u-(1-\mathcal{L}_2)^{-1}\operatorname{div}(u\otimes (1-\mathcal{L}_2)u)) \\ &=\delta t^{-1} v+ \mathcal{L}_1 v-t^{-\delta}P(1-\mathcal{L}_2)^{-1}\operatorname{div}(v\otimes (1-\mathcal{L}_2)v)). \end{align*} Applying Duhamel's principle, we obtain \begin{align*} v&=e^{t\mathcal{L}_1}v_0+\delta\int_0^t e^{(t-s)\mathcal{L}_1}s^{-1}v(s)ds +\int_0^t e^{(t-s)\mathcal{L}_1}s^{-\delta}W(v(s))ds\\ &=\delta\int_0^t e^{(t-s)\mathcal{L}_1}s^{-1}v(s)ds +\int_0^t e^{(t-s)\mathcal{L}_1}s^{-\delta}W(v(s))ds, \end{align*} where we recall $W(f,g)=-P(1-\mathcal{L}_2)^{-1}\operatorname{div}(f(s)\otimes (1-\mathcal{L}_2)g(s))$ (and for notational convenience set $W(f,f)=W(f)$) and in the last line used that $v_0=0$. Using $v=t^\delta u$, we obtain \[ u=\delta t^{-\delta}\int_0^t e^{(t-s)\mathcal{L}_1}s^{\delta-1}u(s)ds +t^{-\delta}\int_0^t e^{(t-s)\mathcal{L}_1}s^\delta W(u(s))ds. \] The key idea here is that we can choose $\delta$ to be large enough to cancel any singularities that occur at $s=0$. Now we are ready to set up the induction. We have by Theorem \ref{old style short} that the local solution $u$ is in $\dot{C}^T_{(s_2-s_1)/\gamma_1;s_2,p,q}$, where $s_2>\gamma_2$. For induction, we assume this solution $u$ is also in $\dot{C}^T_{(k-s_1)/\gamma_1;k,p,q}$ for some $k\geq s_2$, and seek to show that $u$ is in $\dot{C}^T_{a^*;k+h,p,q}$, where $a^*=(k+h-s_1)/\gamma_1$ and $h$ is a fixed number between $0$ and $1$ which will be chosen later. An application of Proposition \ref{heat kernel bound} gives \begin{equation}\label{hugo} \begin{aligned} &\|u\|_{B^{k+h}_{p,q}}\\ &\leq t^{-\delta}\Big(\delta\int_0^t \|e^{(t-s)\mathcal{L}_1} s^{\delta-1}u(s)\|_{B^{k+h}_{p,q}}ds + \int_0^t \|e^{(t-s)\mathcal{L}_1} s^{\delta}W(u(s))\|_{B^{k+h}_{p,q}}ds\Big)\\ &\leq Ct^{-\delta}\delta \int_0^t |t-s|^{-h/\gamma_1}s^{\delta-1} \|u(s)\|_{B^{k}_{p,q}} \\ &\quad + t^{-\delta}\int_0^t |t-s|^{-b_1/\gamma_1}s^{\delta} \|W(u(s))\|_{B^{k-1}_{\tilde{p},q}} ds, \end{aligned} \end{equation} where $b_1=h+1+n/\tilde{p}-n/p$. For the first term in the right hand side of \eqref{hugo}, we have \begin{equation}\label{hugo2} \begin{aligned} &t^{-\delta}\int_0^t |t-s|^{-h/\gamma_1}s^{\delta-1}\|u(s)\|_{B^{k}_{p,q}}ds\\ &\leq t^{-\delta}\|u\|_{(k-s_1)/\gamma_1;k,p,q} \int_0^t |t-s|^{-h/\gamma_1}s^{\delta-1-(k-s_1)/\gamma_1}ds\\ &\leq C\|u\|_{(k-s_1)/\gamma_1;k,p,q}t^{-\delta}t^{-h/\gamma_1} t^{\delta-1-(k-s_1)/\gamma_1+1}\\ &\leq Ct^{-(k+h-s_1)/\gamma_1}\|u\|_{(k-s_1)/\gamma_1;k,p,q} \end{aligned} \end{equation} This calculation implicitly assumes that the exponents of $|t-s|$ and $s$ in the integral are both strictly greater than negative $1$. For $|t-s|$, this holds provided $h/\gamma_1<1$. For $s$, it works for a sufficiently large choice of $\delta$. We note that without modifying the PDE to include these $t^\delta$ terms, we would need $(k-s_1)/\gamma_1$ to be less than $1$, which does not hold for large $k$. For the second piece, we start by bounding $\|W(u)\|_{B^{k-1}_{\tilde{p},q}}$. Using Proposition \ref{annie2}, Proposition \ref{product est 1}, and finally Proposition \ref{annie} and Proposition \ref{sob embedding}, we obtain \begin{equation}\label{mal38} \begin{aligned} \|W(u(s))\|_{B^{k-1}_{\tilde{p},q}} &\leq \|u\otimes (1-\mathcal{L}_2)u\|_{B^{k-\gamma_2}_{\tilde{p},q}}\\ &\leq \|u\|_{L^{p_1}}\|(1-\mathcal{L}_2)u\|_{B^{k-\gamma_2}_{p,q}} +\|u\|_{B^{k-\gamma_2}_{p, q}}\|(1-\mathcal{L}_2)u\|_{L^{p_3}} \\ &\leq \|u\|_{B^{r_1^+}_{p,q}}\|u\|_{B^{k}_{p,q}} +\|u\|_{B^{k-\gamma_2}_{p,q}}\|u\|_{B^{\gamma_2+r_2^+}_{p,q}} \end{aligned} \end{equation} where \begin{equation}\label{aldi65} \begin{gathered} 1/\tilde{p}=1/p_1+1/p=1/p+1/p_2,\\ p_i=\frac{np}{n-r_ip}, \quad i=1,2 \end{gathered} \end{equation} and \begin{equation}\label{aldi652} \begin{gathered} r_i 0$, so the constraint will be satisfied if we choose $h$ to be small enough so that $r^+_1>h.$ Note that $r_1<\min\{s_2, n/p\}$, so this choice only depends on $s_2, n$, and $p$. Also from \eqref{durant}, we have \[ 1-(1+n/p-s_1+\varepsilon)/\gamma_1=1-(1+n/p-s_1+s_2) /\gamma_1+(s_2-\varepsilon) /\gamma_1\geq s_2/\gamma. \] Applying this to \eqref{induct2}, we finally get \begin{equation}\label{ward} I_1\leq Ct^{-(h+k-s_1)/\gamma_1+(s_2-\varepsilon)/\gamma_1}. \end{equation} A similar calculation for $I_2$ yields \begin{equation} \label{durant2} \begin{aligned} I_2&\leq C\|u\|_{a_3;k-\gamma_2,p,q}\|u\|_{a_4;\gamma_2+r^+_2,p,q} t^{-(h+k-s_1)/\gamma_1+s_2/\gamma_1}\\ &\leq C t^{-(h+k-s_1)/\gamma_1+(s_2-\varepsilon)/\gamma_1}, \end{aligned} \end{equation} where we recall that $r^+_2h$, which now means the choice of $h$ also depends on $\gamma_2$. Using \eqref{ward} and \eqref{durant2} in \eqref{hugo3}, we have \begin{equation}\label{durant3} t^{-\delta}\int_0^t |t-s|^{-b_1/\gamma_1}s^{\delta} \|W(u)\|_{B^{k-1}_{\tilde{p},q}} ds\leq C t^{-(h+k-s_1) /\gamma_1+(s_2-\varepsilon)/\gamma_1}, \end{equation} and using \eqref{hugo2} and \eqref{durant3} in \eqref{hugo} gives \[ \|u(t)\|_{B^{k+h}_{p,q}}\leq C\delta t^{-(k+h-s_1)/\gamma_1} \|u\|_{(k-s_1)/\gamma_1;k,p,q}+C t^{-(h+k-s_1)/\gamma_1+(s_2-\varepsilon)/\gamma_1}, \] Multiplying both sides by $t^{(k+h-s_1)/\gamma_1}$ and taking the supremem over $t$, we finally get \[ \|u\|_{(k+h-s_1)/\gamma_1;k+h, p, q}\leq C\delta \|u\|_{(k-s_1)/\gamma_1;k,p,q}+C \sup_{t\leq T} t^{s_2/\gamma_1}, \] which completes the induction argument. We remark that $\delta$ is chosen after beginning the induction step (and thus can be absorbed into the constant), while the appropriate value of $h$ is fixed by the known parameters $n, p, s_1, s_2$, and $\gamma_2$. \end{proof} \begin{thebibliography}{10} \bibitem{barbato} D.~Barbato, F.~Morandin, and M.~Romito, \emph{Global regularity for a slightly supercritical hyperdissipative {N}avier-{S}tokes system}, arXiv:1407.6734. \bibitem{chae} D.~Chae; \emph{Local existence and blow-up criterion for the {E}uler equations in the {B}esov spaces}, Asymptotic {A}nalysis \textbf{38} (2004), 339--358. \bibitem{chemin} J. Y. Chemin; \emph{About the navier-stokes system}, Publications du Laboratoire d'analyse numerique (1996). \bibitem{galplan} I.~Gallagher, F.~Planchon; \emph{On global infinite energy solutions to the {N}avier-{S}tokes equations in two dimensions}, Arch. Ration. Mech. Anal. \textbf{161} (2002), no.~4, 307--337. \bibitem{holmleray} J.~D. Gibbon, D.~D. Holm; \emph{Estimates for the {LANS}-{$\alpha$}, {L}eray-{$\alpha$} and {B}ardina models in terms of a {N}avier-{S}tokes {R}eynolds number}, Indiana Univ. Math. J. \textbf{57} (2008), no.~6, 2761--2773. \bibitem{hmidi} T.~Hmidi; \emph{On a maximum principle and its applications to logarithmically critical {B}oussinesq system}, Analysis and PDE \textbf{4} (2001), 247--284. \bibitem{Kato} T.~Kato; \emph{Strong ${L}^p$-solutions of the {N}avier {S}tokes {E}quation in $\mathbb{R}^m$, with {A}pplications to {W}eak {S}olutions}, Mathematische {Z}eitschrift \textbf{187} (1984), 471--480. \bibitem{katoinduction} T.~Kato; \emph{The {N}avier-{S}tokes equation for an incompressible fluid in {${\bf R}^2$} with a measure as the initial vorticity}, Differential Integral Equations \textbf{7} (1994), no.~3-4, 949--966. \bibitem{KP} T.~Kato, G.~Ponce; \emph{The {N}avier-{S}tokes equation with weak initial data}, Internat. Math. Res. Notices (1994), no.~10, 10 pages (electronic). \bibitem{KT} H.~Koch, D.~Tataru; \emph{Well-posedness for the {N}avier-{S}tokes equations}, Adv. Math. \textbf{157} (2001), no.~1, 22--35. \bibitem{lady356} O.~A. Ladyzhenskaya; \emph{The mathematical theory of viscous incompressible flow}, Second English edition, revised and enlarged. Translated from the Russian by Richard A. Silverman and John Chu. Mathematics and its Applications, Vol. 2, Gordon and Breach Science Publishers, New York, 1969. \bibitem{nsbook} P.~G. Lemarie-Rieusset; \emph{Recent developments in the navier-stokes problem}, Chapman and Hall/CRC, 2002. \bibitem{olsontitialphalike} E.~Olson, E.~S. Titi; \emph{Viscosity versus vorticity stretching: global well-posedness for a family of {N}avier--{S}tokes-alpha-like models}, Nonlinear Analysis. Theory, Methods \& Applications. An International Multidisciplinary Journal. Series A: Theory and Methods \textbf{66} (2007), no.~11, 2427--2458. \bibitem{besovpaper2} N.~Pennington; \emph{Local and global existence for the {L}agrangian {A}veraged {N}avier-{S}tokes equation in {B}esov spaces}, Electron. J. Diff. Equ. \textbf{2012} (2012), no.~89, 1--19. \bibitem{sobpaper} N.~Pennington; \emph{Lagrangian {A}veraged {N}avier-{S}tokes equation with rough data in {S}obolev spaces}, Journal of Mathematical Analysis and Applications \textbf{403} (2013), 72--88. \bibitem{taolog} T.~Tao; \emph{Global regularity for a logarithmically supercritical hyperdissipative {N}avier-{S}tokes equation}, Analysis \& PDE \textbf{2} (2009), no.~3, 361--366. \bibitem{T3} M.~Taylor; \emph{Partial {D}ifferential {E}quations}, Springer-{V}erlag {N}ew {Y}ork, {I}nc., 1996. \bibitem{TT} M.~Taylor; \emph{Tools for {P}{D}{E}}, {M}athematical {S}urveys and {M}onographs, {V}ol. 81, {A}merican {M}athematical {S}ociety, Providence {R}{I}, 2000. \bibitem{genmdh} J.~Wu; \emph{The generalized {M}{H}{D} equations}, Journal of {D}ifferential {E}quations \textbf{195} (2003), 284--312. \bibitem{Wu} J.~Wu; \emph{Lower {B}ounds for an {I}ntegral {I}nvolving {F}ractional {L}aplacians and {G}eneralized {N}avier-{S}tokes {E}quations in {B}esov {S}paces}, Communications in {M}athematical {P}hysics \textbf{263} (2005), 803--831. \bibitem{globalmhd} J.~Wu; \emph{Global regularity for a class of generalized magnetohydrodynamic equations}, Journal of {M}athematical {F}luid {M}echanics \textbf{13} (2011), 295--305. \bibitem{kazuo} K.~Yamazaki; \emph{On the global regularity of generalized {L}eray-alpha type models}, Nonlinear Analysis. Theory, Methods \& Applications. An International Multidisciplinary Journal. Series A: Theory and Methods \textbf{75} (2012), no.~2, 503--515. \bibitem{kazuomhd} K.~Yamazaki; \emph{Global regularity of logarithmically supercritical {M}{H}{D} system with zero diffusivity}, Applied Mathematics Letters \textbf{29} (2014), 46--51. \end{thebibliography} \end{document}