diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzcpfx" "b/data_all_eng_slimpj/shuffled/split2/finalzzcpfx" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzcpfx" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nIn this paper, we study the following kinetic operator \n\\begin{equation}\\label{yo2}\nP=\\partial_t+v \\cdot \\partial_x + a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}, \\ t \\in \\mathbb{R}, \\ x,v \\in \\mathbb{R}^{n},\n\\end{equation}\nwhere $0<\\sigma<1$ is a constant parameter, $x \\cdot y$ stands for the standard dot-product on $\\mathbb{R}^n$ and $a$ denotes a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function satisfying \n\\begin{equation}\\label{eq0.5}\n\\exists a_0>0, \\forall (t,x,v) \\in \\mathbb{R}^{2n+1}, \\ a(t,x,v) \\geq a_0>0.\n\\end{equation}\nHere the notation $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ stands for the space of $C^{\\infty}(\\mathbb{R}^{2n+1})$ functions whose derivatives of any order are bounded on $\\mathbb{R}^{2n+1}$ and \n$(-\\tilde{\\Delta}_v)^{\\sigma}$ is the Fourier multiplier with symbol\n\\begin{equation}\\label{eq0}\nF(\\eta)=|\\eta|^{2\\sigma}w(\\eta)+|\\eta|^2\\big(1-w(\\eta)\\big), \\ \\eta \\in \\mathbb{R}^n,\n\\end{equation}\nwith $|\\cdot|$ being the Euclidean norm, $w$ a $C^{\\infty}(\\mathbb{R}^n)$ function satisfying $0 \\leq w \\leq 1$, $w(\\eta)=1$ if $|\\eta| \\geq 2$, $w(\\eta)=0$ if $|\\eta| \\leq 1$; and $D_t=(2\\pi i)^{-1}\\partial_t$, $D_x=(2\\pi i)^{-1}\\partial_x$, $D_v=(2\\pi i)^{-1}\\partial_v$.\n\n\nWhen $\\sigma=1$, this operator reduces to the so-called Vlasov-Fokker-Planck operator, whereas when $0<\\sigma<1$, it stands for a simplified linear model of the spatially inhomogeneous Boltzmann equation without angular cutoff (see the end of this introduction together with section~\\ref{kkboltz} in appendix). This is our motivation for studying the regularizing properties of this linear model and establishing hypoelliptic estimates with optimal loss of derivatives with respect to the exponent $0<\\sigma<1$ of the fractional Laplacian $(-\\tilde{\\Delta}_v)^{\\sigma}$. This linear model has the familiar structure \n$$\\textrm{Transport part in the }(t,x) \\textrm{ variables}\\ + \\textrm{ Elliptic part in the }v \\textrm{ variable}$$ \nand it is easy to get the regularity in the $v$ variable. The non-commutation of the transport part (the skew-adjoint part) with the self-adjoint elliptic part $(-\\tilde{\\Delta}_v)^{\\sigma}$ will produce the regularizing effect in the $x$ variable. \n\n\nRegarding this linear model, the existence and the $C^{\\infty}$ regularity for the solutions of the Cauchy problem to linear and semi-linear equations associated with the operator (\\ref{yo2}) were proved in \\cite{MoXu}. H.~Chen, W.-X.~Li and C.-J.~Xu have also recently studied its Gevrey hypoellipticity. More specifically, they established in~\\cite{chen} (Proposition~2.1) the following hypoelliptic estimate. Let $P$ be the operator defined in (\\ref{yo2}) and $K$ a compact subset of $\\mathbb{R}^{2n+1}$. For any $s \\geq 0$, there exists a positive constant $C_{K,s}>0$ such that for any $u \\in C_0^{\\infty}(K)$,\n\\begin{equation}\\label{yo4} \n\\|u\\|_{s+\\delta} \\leq C_{K,s}\\big(\\|Pu\\|_{s}+\\|u\\|_{s}\\big),\n\\end{equation}\nwith $\\|\\cdot\\|_{s}$ standing for the $H^s(\\mathbb{R}^{2n+1})$ Sobolev norm and \n\\begin{equation}\\label{yo4.5}\n\\delta=\\textrm{max}\\Big(\\frac{\\sigma}{4},\\frac{\\sigma}{2}-\\frac{1}{6}\\Big)>0.\n\\end{equation}\nThe notation $C_0^{\\infty}(K)$ stands for the set of $C_0^{\\infty}(\\mathbb{R}^{2n+1})$ functions with support in $K$. This hypoelliptic estimate with loss of \n$$\\mbox{\\rm max}(2\\sigma,1)-\\delta>0,$$ \nderivatives is then a key instrumental ingredient for their investigation of the Gevrey hypoellipticity of the operator $P$. \nHowever, this hypoelliptic estimate (\\ref{yo4}) is not optimal. \nIn the present work, we are interested in establishing hypoelliptic estimates with optimal loss of derivatives with respect to the exponent $0<\\sigma<1$ of the fractional Laplacian $(-\\tilde{\\Delta}_v)^{\\sigma}$. More specifically, we shall show by using different microlocal techniques that the operator $P$ is hypoelliptic with a loss of \n$$\\frac{\\mbox{\\rm max}(4\\sigma^2,1)}{(2\\sigma+1)}>0,$$ \nderivatives, that is, that the hypoelliptic estimates (\\ref{yo4}) hold with the new positive gain \n\\begin{equation}\\label{yo6}\n\\delta=\\frac{2\\sigma}{2\\sigma+1}>0,\n\\end{equation}\nwhich improves for any $0<\\sigma<1$ the gain provided by (\\ref{yo4.5}),\n$$\\frac{2\\sigma}{2\\sigma+1}>\\textrm{max}\\Big(\\frac{\\sigma}{4},\\frac{\\sigma}{2}-\\frac{1}{6}\\Big).$$\nOur main result reads as follows.\n\n\n\\bigskip\n\n\n\\begin{theorem}\\label{TTH1}\nLet $P$ be the operator defined in \\emph{(\\ref{yo2})}, $K$ be a compact subset of $\\mathbb{R}^{2n+1}$ and $s \\in \\mathbb{R}$. Then, there exists a positive constant $C_{K,s}>0$ such that for all $u \\in C_0^{\\infty}(K)$,\n\\begin{equation}\\label{yo5}\n\\big\\|(1+|D_t|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_x|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_{v}|^{2\\sigma})u\\big\\|_{s} \\leq C_{K,s} \\big(\\|Pu\\|_{s}+\\|u\\|_{s}\\big),\n\\end{equation}\nwith $\\|\\cdot\\|_{s}$ being the $H^s(\\mathbb{R}^{2n+1})$ Sobolev norm.\n\\end{theorem}\n\n\\bigskip\n\nThe hypoelliptic estimates (\\ref{yo5}) are optimal in term of the exponents of derivative terms appearing in their left-hand-side, namely, $2\\sigma\/(2\\sigma+1)$ for the regularity in the time and space variables and $2\\sigma$ for the regularity in the velocity variable. The exponent $2\\sigma$ for the regularity in the velocity variable has indeed the same growth as the diffusive part of the kinetic operator (\\ref{yo2}). Regarding the optimality of the exponent $2\\sigma\/(2\\sigma+1)$ for the regularity in the time and space variables, we first notice that Theorem~\\ref{TTH1} is a natural extension for the values of the parameter $0<\\sigma<1$ of the well-known optimal hypoelliptic estimates with loss of $4\/3$ derivatives known for the Vlasov-Fokker-Planck operator, case $\\sigma=1$, (see~\\cite{bouchut,chen1,perthame}),\n$$\\big\\|(1+|D_t|^{2\/3}+|D_x|^{2\/3}+|D_{v}|^{2})u\\big\\|_{s} \\leq C_{K,s} \\big(\\|Pu\\|_{s}+\\|u\\|_{s}\\big).$$\nSee also~\\cite{landau} for general microlocal methods for proving optimal hypoelliptic estimates with loss of $4\/3$ derivatives for certain classes of kinetic equations. \n\n\nWe deduce the optimality of the exponent $2\\sigma\/(2\\sigma+1)$ for the regularity in the time and space variables by using a simple scaling argument. Indeed, let us consider the specific case when the function $a$ appearing in the definition of the kinetic operator $P$ is constant and assume that the hypoelliptic estimates (\\ref{yo4}) hold for a positive gain $\\delta>0$. It follows that the estimate \n$$\\|(|D_t|^{\\delta}+|D_x|^{\\delta}+|D_{v}|^{\\delta})u\\|_{L^2} \\leq C\\big(\\|iD_tu+iv \\cdot D_xu + |D_v|^{2\\sigma}u\\|_{L^2}+\\|u\\|_{L^2}\\big),$$\nholds for any $u \\in C_0^{\\infty}(\\mathbb{R}^{2n+1})$ with support in the closed unit Euclidean ball $B_1=\\overline{B(0,1)}$ in $\\mathbb{R}^{2n+1}$. The symplectic invariance of the Weyl quantization (Theorem~18.5.9 in~\\cite{hormander}) shows that for any $\\lambda \\geq 1$, \n$$T_{\\lambda}^{-1}\\big(iD_t+iv \\cdot D_x + |D_v|^{2\\sigma}\\big)T_{\\lambda}=\\lambda^{\\frac{2\\sigma}{2\\sigma+1}}\\big(iD_t+iv \\cdot D_x + |D_v|^{2\\sigma}\\big)$$\nand\n$$T_{\\lambda}^{-1}\\big(|D_t|^{\\delta}+|D_x|^{\\delta}+|D_{v}|^{\\delta}\\big)T_{\\lambda}=\\lambda^{\\frac{2\\sigma\\delta}{2\\sigma+1}}|D_t|^{\\delta}+\\lambda^{\\delta}|D_x|^{\\delta}+\\lambda^{\\frac{\\delta}{2\\sigma+1}}|D_{v}|^{\\delta},$$\nwhere $T_{\\lambda}$ stands for the following unitary transformation on $L^{2}(\\mathbb{R}^{2n+1})$, \n$$T_{\\lambda}u(t,x,v)=\\lambda^{\\frac{2\\sigma+n}{2(2\\sigma+1)}+\\frac{n}{2}}u(\\lambda^{\\frac{2\\sigma}{2\\sigma+1}}t,\\lambda x,\\lambda^{\\frac{1}{2\\sigma+1}}v).$$\nRecalling that $\\lambda \\geq 1$, we notice that the $C_0^{\\infty}(\\mathbb{R})$ function $T_{\\lambda}u$ is supported in $B_1$ if the $C_0^{\\infty}(\\mathbb{R})$ function $u$ is supported in $B_1$.\nBy applying the previous a priori estimate to functions $T_{\\lambda}u$, we obtain that \n$$\\|(|D_t|^{\\delta}+|D_x|^{\\delta}+|D_{v}|^{\\delta})T_{\\lambda}u\\|_{L^2} \\leq C\\big(\\|(iD_t+iv \\cdot D_x + |D_v|^{2\\sigma})T_{\\lambda}u\\|_{L^2}+\\|T_{\\lambda}u\\|_{L^2}\\big).$$\nBy using that $T_{\\lambda}^{-1}$ is a unitary transformation on $L^{2}(\\mathbb{R}^{2n+1})$, we deduce that for any $\\lambda \\geq 1$, \n\\begin{multline*}\n\\|(\\lambda^{\\frac{2\\sigma\\delta}{2\\sigma+1}}|D_t|^{\\delta}+\\lambda^{\\delta}|D_x|^{\\delta}+\\lambda^{\\frac{\\delta}{2\\sigma+1}}|D_{v}|^{\\delta})u\\|_{L^2} \\\\\\ \\leq C\\big(\\lambda^{\\frac{2\\sigma}{2\\sigma+1}}\\|(iD_t+iv \\cdot D_x + |D_v|^{2\\sigma})u\\|_{L^2}+\\|u\\|_{L^2}\\big).\n\\end{multline*}\nThis estimate may hold for any $\\lambda \\geq 1$ only if \n$$\\delta \\leq \\frac{2\\sigma}{2\\sigma+1}.$$\nThis scaling argument shows that the positive gain $2\\sigma\/(2\\sigma+1)>0$ in the hypoelliptic estimates (\\ref{yo4}) is the optimal possible one.\n\n\nAs a consequence of these optimal hypoelliptic estimates, we obtain the following result where we write\n$$f \\in H^s_{\\textrm{loc}, (t_0,x_0)}(\\mathbb{R}^{2n+1}_{t,x,v}),$$\nif there exists an open neighborhood $U$ of the point $(t_0,x_0)$ in $\\mathbb{R}^{n+1}$ such that\n$\\phi(t,x) f\\in H^s(\\mathbb{R}_{t,x,v}^{2n+1})$\nfor any $\\phi \\in C_0^\\infty(U)$. \n\n\n\\bigskip\n\n\\begin{Corollary}\\label{ev1}\nLet $P$ be the operator defined in \\emph{(\\ref{yo2})} and $N \\in \\mathbb{N}$. \nIf $u \\in H_{-N}(\\mathbb{R}_{t,x,v}^{2n+1})$ and $Pu \\in H^s_{\\emph{\\textrm{loc}}, (t_0,x_0)}(\\mathbb{R}^{2n+1}_{t,x,v})$ with $s \\ge 0$, then there exists \nan integer $k \\geq 1$ such that \n\\[\n\\frac{u}{\\langle v \\rangle^k} \\in H^{s+\\frac{2\\sigma}{2\\sigma+1}}_{\\emph{\\textrm{loc}}, (t_0,x_0)}(\\mathbb{R}^{2n+1}_{t,x,v}),\n\\]\nwhere $\\langle v \\rangle=(1+|v|^2)^{1\/2}$.\nIn particular, if $u \\in H_{-N}(\\mathbb{R}^{2n+1})$ and $Pu \\in H^{\\infty}(\\mathbb{R}^{2n+1})$ then $u \\in C^{\\infty}(\\mathbb{R}^{2n+1})$.\n\\end{Corollary}\n\n\\bigskip\n\n\n\n\nCorollary~\\ref{ev1} allows to recover the $C^{\\infty}$ hypoellipticity proved in \\cite{MoXu} (Theorem~1.2) with now optimal loss of derivatives. Notice that the equation (\\ref{yo2}) is not a classical pseudodifferential equation. Indeed, the coefficient $v$ in (\\ref{yo2}) is unbounded and the fractional Laplacian $(-\\tilde{\\Delta}_v)^{\\sigma}$ is a classical pseudo-differential operator in the velocity variable $v$ but not in all the variables $t,x,v$. This accounts for parts of the difficulties encountered when studying this kinetic operator in particular when using cutoff functions in the velocity variable. This also accounts for the weight $\\langle v \\rangle^{-k}$ appearing in the statement of Corollary~\\ref{ev1}. Notice that the proof of this result is constructive and that one may derive an explicit (possibly not sharp) bound on the integer $k \\geq 1$.\n\n\n\nThe proof of Theorem~\\ref{TTH1} is relying on some microlocal techniques developed by N.~Lerner for proving energy estimates while using the Wick quantization \\cite{cubo}. Let us mention that some of these techniques were already used in the work~\\cite{lms}. The strategy for proving Theorem~\\ref{TTH1} is the following. We want to consider this equation as an evolution equation along the characteristic curves of $\\partial_t+v \\cdot \\partial_x$; for that purpose, we straighten this vector field and get the normal form\n$$iD_t+a(t,D_{x_2},D_{x_1}+tD_{x_2})F(x_2-tx_1).$$\nThis normal form suggests to derive some a priori estimates for the one-dimensional first-order differential operator \n$$iD_t+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1),$$\ndepending on the parameters $x_1,x_2,\\xi_1,\\xi_2 \\in \\mathbb{R}^n$. We then deduce from those a priori estimates some a priori estimates for the operator \n$$iD_t+\\big[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)\\big]^{\\textrm{Wick}},$$\ndefined by using the Wick quantization. As a last step, we need to control some remainder terms in order to come back to the standard quantization and derive a priori estimates for the original operator\n$$iD_t+a(t,D_{x_2},D_{x_1}+tD_{x_2})F(x_2-tx_1).$$\nFor the sake of completeness and to keep the paper essentially self-contained, the definition and all the main features of the Wick quantization are recalled in appendix (Section~\\ref{appendix}). Next section is devoted to the proof of a key hypoelliptic estimate (Proposition~\\ref{th1}) which is the core of the present work. This key estimate is then the main ingredient in Section~\\ref{section3} for proving Theorem~\\ref{TTH1}. Corollary~\\ref{ev1} is established in Section~\\ref{section4}. \n\n\n\n\n\nBefore ending this introduction, we give some references and comments about the hypoelliptic properties of the non-cutoff Boltzmann equation. Following the rigorous derivation of the lower bound for the non-cutoff Boltzmann collision operator in~\\cite{alexandre0}, there have been many works on the regularity for the solutions of the Boltzmann equation in both \nspatially homogeneous (see \\cite{alexandre1, al-saf-1, desv1,HMUY, MU,Mo2} and references therein) and\ninhomogeneous cases (see \\cite{amuxy3, amuxy4, amuxy5-3}).\nIn all those works, it was highlighted that the Boltzmann collision operator behaves essentially as a fractional Laplacian $(-\\tilde{\\Delta}_v)^{\\sigma}$ \nunder the angular singularity assumption on the collision cross-section (see \\cite{alexandre0, amuxy3}). We refer the reader to Section~\\ref{kkboltz} in appendix for comprehensive explanations about the relevance of this model. In the spatially homogeneous case, this diffusive structure implies a $C^{\\infty}$ smoothing effect for the weak solutions to the Cauchy problem constructed in~\\cite{Villani} (See \\cite{amuxy7, HMUY}). Furthermore, as in the case of the heat equation, the spatially homogeneous Boltzmann equation without angular cutoff enjoys a\nsmoothing effect in the Gevrey class of order $\\sigma$ (see \\cite{ L-X, MU, Mo2}). Related to this Gevrey smoothing effect for the spatially homogeneous Boltzmann equation, an \nultra-analytic smoothing effect was proved in~\\cite{Mo1} for both non-linear homogeneous Landau equations and inhomogeneous linear Landau equations.\nRegarding the study of the Boltzmann equation in Gevrey spaces, we also refer the reader to the seminal work \\cite{Ukai} which establishes the existence and uniqueness in Gevrey classes of a local solution to the Cauchy problem for the Boltzmann equation in both spatially homogeneous and inhomogeneous cases.\nConsidering now the spatially inhomogeneous Boltzmann equation without angular cutoff, \nthe $C^\\infty$ hypoellipticity was established in \\cite{amuxy3,amuxy4,amuxy5-3} by using the coercivity estimate for proving the regularity with respect to the velocity variable $v$ (see~\\cite{alexandre0,amuxy3}) and a version of the uncertainty principle for proving the regularity in the time and space variables $t,x$ via bootstrap arguments (see~\\cite{amuxy2}). However, it should be noted that those works do not provide any optimal hypoelliptic estimates for the spatially inhomogeneous Boltzmann equation without angular cutoff. As an attempt to understand further the smoothing effect induced by the Boltzmann collision operator and to relate exactly the structure of the angular singularity in the collision cross-section to this regularizing effect, the present work studies the hypoelliptic properties of the simplified linear model (\\ref{yo2}) and aims at giving insights on the hypoelliptic properties what may be expected for the general spatially inhomogeneous Boltzmann equation without angular cutoff. \n\n\n\n\n\n\n\n\n\\section{A key hypoelliptic estimate}\\label{section2}\n\nTheorem~\\ref{TTH1} will be derived from the following key hypoelliptic estimate: \n\n\n\n\\bigskip\n\n\n\\begin{proposition}\\label{th1}\nLet $P$ be the operator defined in~\\emph{(\\ref{yo2})} and $T>0$ be a positive constant. Then, there exists a positive constant $C_T>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x,v}^{2n+1})$ satisfying\n\\begin{equation}\\label{hypo0}\n\\emph{\\textrm{supp }} u(\\cdot,x,v) \\subset [-T,T], \\ (x,v) \\in \\mathbb{R}^{2n},\n\\end{equation}\nwe have\n\\begin{equation}\\label{hypo}\n\\big\\|(1+|D_x|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_{v}|^{2\\sigma})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C_T \\big(\\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}\\big).\n\\end{equation}\n\\end{proposition}\n\n\\bigskip\n\n\\noindent\n\nIn the following, we use standard notations for symbol classes, see~\\cite{hormander} (Chapter~18) or \\cite{birkhauser}.\nThe symbol class $S(m,\\Gamma)$ associated to the order function $m$ and metric\n$$\\Gamma=\\frac{dx^2}{\\varphi(x,\\xi)^2}+\\frac{d\\xi^2}{\\Phi(x,\\xi)^2},$$\nwith $\\varphi$, $\\Phi$ given positive functions,\nstands for the set of functions $a\\in C^{\\infty}(\\mathbb{R}_{x,\\xi}^{2n},\\mathbb{C})$ satisfying for all $\\alpha \\in \\mathbb{N}^{n}$ and $\\beta \\in \\mathbb{N}^n$, \n\\begin{equation}\\label{ro1}\n\\exists C_{\\alpha, \\beta}>0, \\forall (x,\\xi) \\in \\mathbb{R}^{2n}, \\ |\\partial_{x}^{\\alpha}\\partial_{\\xi}^{\\beta} a(x,\\xi)| \\leq C_{\\alpha, \\beta} m(x,\\xi) \\varphi(x,\\xi)^{-|\\alpha|}\\Phi(x,\\xi)^{-|\\beta|}.\n\\end{equation}\n\n\n\n\\subsection{Some symbol reductions} We begin by few symbol reductions in order to reduce the symbol of the operator to a convenient normal form. For convenience only, we shall use the Weyl quantization rather than the standard one.\nNotice that it will be sufficient to prove the hypoelliptic estimate (\\ref{hypo}) for the operator $p^w$ defined by the Weyl quantization of the symbol \n\\begin{equation}\\label{eq1}\np(t,x,v;\\tau,\\xi,\\eta)=2\\pi i\\tau+2\\pi iv\\cdot \\xi+a(t,x,v)F(2\\pi\\eta),\n\\end{equation}\nwith $\\tau$, $\\xi$, $\\eta$ respectively standing for the dual variables of the variables $t$, $x$, $v$ and $F$ being the function defined in (\\ref{eq0}),\nwhile using the following normalization for the Weyl quantization\n$$(a^wu)(x)=\\int_{\\mathbb{R}^{2n}}e^{2i\\pi(x-y)\\cdot \\xi}a\\big(\\frac{x+y}{2},\\xi\\big)u(y)dyd\\xi.$$\nIndeed, symbolic calculus shows that the operator\n$$R=P-p^w+\\frac{1}{2i}\\nabla_va(t,x,v) \\cdot (\\nabla F)(2\\pi D_v),$$\nis bounded on $L^2$. This is a direct consequence of the $L^2$ continuity theorem in the class $S_{00}^0$ (case $m=\\varphi=\\Phi=1$ in (\\ref{ro1})) after noticing that the Weyl symbol of the operator $R$ together with all its derivatives of any order are bounded on $\\mathbb{R}^{4n+2}$, since all the derivatives of order greater or equal to 2 of the symbol $F$ are bounded given the choice of the positive parameter $0 < \\sigma <1$. We refer the reader to formula (2.1.26) in \\cite{birkhauser} for explicit constants in the composition formula with the normalization of the Weyl quantization chosen here. It then remains to notice that for any $\\varepsilon>0$, there is a positive constant $C_{\\varepsilon}>0$ such that \n\\begin{multline*}\n\\big\\|\\nabla_va(t,x,v) \\cdot (\\nabla F)(2\\pi D_v)u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\lesssim \\big\\|(\\nabla F)(2\\pi D_v)u\\big\\|_{L^2(\\mathbb{R}^{2n+1})}\\\\ \\lesssim \\|(1+|D_v|^{2\\sigma-1})u\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq \\varepsilon \\|(1+|D_v|^{2\\sigma})u\\|_{L^2(\\mathbb{R}^{2n+1})} +C_{\\varepsilon}\\|u\\|_{L^2(\\mathbb{R}^{2n+1})},\n\\end{multline*} \nsince the function $\\nabla_va$ is bounded on $\\mathbb{R}^{2n+1}$. When studying the operator $p^w$, it is convenient to work on the Fourier side in the variables $x,v$. This is equivalent to study the operator defined by the Weyl quantization of the symbol\n$$2\\pi i\\tau-2\\pi i x \\cdot \\eta +a(t,-\\xi,-\\eta)F(2\\pi v).$$\nAfter relabeling the variables and simplifying the notations, we are reduced to study the operator $P=p^w$ defined by the Weyl quantization of the symbol\n\\begin{equation}\\label{eq2}\np(t,x,v;\\tau,\\xi,\\eta)= i\\tau- iy \\cdot \\xi+a(t,\\xi,\\eta)F(x),\n\\end{equation}\nwhere $a$ stands for a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function satisfying \n\\begin{equation}\\label{eq00.5}\n\\exists a_0>0, \\ \\forall (t,\\xi,\\eta) \\in \\mathbb{R}^{2n+1}, \\ a(t,\\xi,\\eta) \\geq a_0>0,\n\\end{equation}\nand $F$ is the function\n\\begin{equation}\\label{eq00}\nF(x)=|x|^{2\\sigma}w(x)+|x|^2\\big(1-w(x)\\big),\n\\end{equation}\nwith $w$ a new function having experienced a small homothetic transformation compared to the function appearing in (\\ref{eq0}).\nFor the sake of simplicity, we shall keep the definition given in (\\ref{eq0}) and assume that $w \\in C^{\\infty}(\\mathbb{R}^n)$, $0 \\leq w \\leq 1$, $w(\\eta)=1$ if $|\\eta| \\geq 2$, and $w(\\eta)=0$ if $|\\eta| \\leq 1$. \nUsing these notations, Proposition~\\ref{th1} is equivalent to the proof of the following a priori estimate. For any $T>0$, there exists a positive constant $C_T>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x,y}^{2n+1})$ satisfying\n\\begin{equation}\\label{eq3}\n\\textrm{supp } u(\\cdot,x,y) \\subset [-T,T], \\ (x,y) \\in \\mathbb{R}^{2n},\n\\end{equation}\nwe have\n\\begin{equation}\\label{hypo2}\n\\big\\|(1+|x|^{2\\sigma}+|y|^{\\frac{2\\sigma}{2\\sigma+1}})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C_T \\big(\\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}\\big).\n\\end{equation}\nIn order to do so, we consider the new variables \n$$(x_1,x_2)=(y,x+ty),$$ \nwith $t \\in \\mathbb{R}$ fixed. We define $A(x,y)=(y,x+ty)$. Associated to this linear change of variables are the real linear symplectic transformation\n$$(x_1,x_2;\\xi_1,\\xi_2)=\\chi(x,y;\\xi,\\eta)=(A(x,y);(A^{-1})^T(\\xi,\\eta))=(y,x+ty;\\eta-t\\xi,\\xi),$$\nand the two unitary operators on $L^2(\\mathbb{R}^{2n})$\n$$(M_tu)(x_1,x_2)=u(x_2-tx_1,x_1); \\ (M_t^{-1}u)(x,y)=u(y,x+ty).$$ \nKeeping on considering the $t$-variable as a parameter and defining the symbol\n$$b_t(x,y;\\xi,\\eta)=- iy \\cdot \\xi+a(t,\\xi,\\eta)F(x),$$\nwe have\n$$(b_t \\circ \\chi^{-1})(x_1,x_2;\\xi_1,\\xi_2)=-ix_1 \\cdot \\xi_2+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1),$$\nand the symplectic invariance of the Weyl quantization (Theorem~18.5.9 in~\\cite{hormander}) or a simple direct calculation\n$$M_t^{-1}a^wM_t=(a \\circ \\chi)^w,$$ \nshow that \n$$M_t b_t^w(x,y,D_x,D_y)M_t^{-1}=-ix_1 \\cdot D_{x_2}+\\big[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)\\big]^w.$$\nWe then consider the two unitary operators acting on $L^2(\\mathbb{R}_{t,x,y}^{2n+1})$,\n$$(Mu)(t,x_1,x_2)=u(t,x_2-tx_1,x_1), \\ (M^{-1}u)(t,x,y)=u(t,y,x+ty),$$ \nand notice that \n$$(MiD_t M^{-1}u)(t,x_1,x_2)=iD_t u(t,x_1,x_2)+ix_1 \\cdot D_{x_2}u(t,x_1,x_2).$$\nIt follows that \n$$iD_t+\\big[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)\\big]^w=MPM^{-1}.$$\nNotice also that \n$$1+|x_2-tx_1|^{2\\sigma}+|x_1|^{\\frac{2\\sigma}{2\\sigma+1}}=M(1+|x|^{2\\sigma}+|y|^{\\frac{2\\sigma}{2\\sigma+1}})M^{-1}.$$\nWe can therefore reduce the proof of Proposition~\\ref{th1} to the proof of the following a priori estimate. For any $T>0$, there exists a positive constant $C_T>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n\\begin{equation}\\label{eq4}\n\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},\n\\end{equation}\nwe have\n\\begin{equation}\\label{eq5}\n\\big\\|(1+|x_2-tx_1|^{2\\sigma}+|x_1|^{\\frac{2\\sigma}{2\\sigma+1}})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C_T \\big(\\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}\\big),\n\\end{equation}\nfor the operator \n\\begin{equation}\\label{eq6}\nP=iD_t+\\big[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)\\big]^w.\n\\end{equation}\n\n\\bigskip\n\n\n\\subsection{Energy estimates via the Wick quantization}\nIn order to establish the a priori estimate (\\ref{eq5}), we shall use some techniques developed in~\\cite{cubo} for proving energy estimates via the Wick quantization. We recall in appendix (see Section~\\ref{appendix}) the definition and all the main features of the Wick quantization which will be used here. \n\n\nA first step in adapting this approach is to study the first-order differential operator on $L^2(\\mathbb{R}_t)$,\n$$\\tilde{P}=iD_t+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1),$$\nwhere the variables $x_1,x_2,\\xi_1,\\xi_2$ are considered as parameters, and prove some a priori estimates with respect to those parameters. \nWe begin by noticing from (\\ref{eq00.5}) that for any $u \\in \\mathscr{S}(\\mathbb{R}_t)$, \n$$\\textrm{Re}(\\tilde{P}u,u)_{L^2(\\mathbb{R}_t)}=\\int_{\\mathbb{R}}a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)|u(t)|^2dt \n\\geq a_0 \\big\\|F(x_2-tx_1)^{1\/2}u\\big\\|_{L^2(\\mathbb{R}_t)}^2,$$\nsince the operator $iD_t$ is skew-adjoint.\nIt follows from the Cauchy-Schwarz inequality that\n\\begin{equation}\\label{eq7}\na_0 \\big\\|F(x_2-tx_1)^{1\/2}u\\big\\|_{L^2(\\mathbb{R}_t)}^2 \\leq \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|u\\|_{L^2(\\mathbb{R}_t)}.\n\\end{equation}\nBy denoting $H={\\mathrm{1~\\hspace{-1.4ex}l}}_{\\mathbb{R}_+}$ the Heaviside function, we may write for any $T \\in \\mathbb{R}$,\n\\begin{multline}\\label{eq8}\n 2\\textrm{Re}(\\tilde{P}u,-H(t-T)u)_{L^2(\\mathbb{R}_t)} = 2\\textrm{Re}(D_tu,iH(t-T)u)_{L^2(\\mathbb{R}_t)} \\\\\n -2\\int_{\\mathbb{R}}H(t-T)a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)|u(t)|^2dt,\n\\end{multline}\nand obtain by a simple integration by parts that \n$$2\\textrm{Re}(D_tu,iH(t-T)u)_{L^2(\\mathbb{R}_t)}=([D_t,iH(t-T)]u,u)_{L^2(\\mathbb{R}_t)}=\\frac{1}{2\\pi}|u(T)|^2.$$\nRecalling that $a \\in L^{\\infty}(\\mathbb{R}^{2n+1})$, one may find a positive constant $C_0>0$ such that \n$$\\Big|2\\int_{\\mathbb{R}}H(t-T)a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)|u(t)|^2dt\\Big| \\leq C_0\\big\\|F(x_2-tx_1)^{1\/2}u\\big\\|_{L^2(\\mathbb{R}_t)}^2.$$\nIt follows from (\\ref{eq7}), (\\ref{eq8}) and the Cauchy-Schwarz inequality that there exists a constant $C_1>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$,\n\\begin{equation}\\label{eq9}\n\\|u\\|_{L^{\\infty}(\\mathbb{R}_t)}^2 \\leq C_1\\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|u\\|_{L^2(\\mathbb{R}_t)}.\n\\end{equation}\nFollowing~\\cite{cubo}, we split-up the $L^2(\\mathbb{R}_t)$-norm of $u$,\n\\begin{multline}\\label{eq10}\n |x_1|^{\\frac{2\\sigma}{2\\sigma+1}}\\|u\\|_{L^2(\\mathbb{R}_t)}^2\n= \\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} > |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}}|u(t)|^2dt \\\\\n+\\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} \\leq |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}} |u(t)|^2dt ,\n\\end{multline}\nand estimate from above the first integral as\n\\begin{multline*}\n\\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} > |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}}|u(t)|^2dt \\\\\n\\leq |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} m\\big(\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} > |x_2-tx_1|^{2\\sigma}\\}\\big) \\|u\\|_{L^{\\infty}(\\mathbb{R}_t)}^2 \\leq 2\\|u\\|_{L^{\\infty}(\\mathbb{R}_t)}^2,\n\\end{multline*}\nwith $m$ standing for the Lebesgue measure on $\\mathbb{R}$.\nIt follows from (\\ref{eq9}) that this first integral can be estimated from above as\n\\begin{equation}\\label{eq10.5}\n\\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} > |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}}|u(t)|^2dt \\leq 2C_1\\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|u\\|_{L^2(\\mathbb{R}_t)}.\n\\end{equation}\nWhile estimating from above the second integral in the right-hand-side of (\\ref{eq10}), we may first write that \n\\begin{multline*}\n\\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} \\leq |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}} |u(t)|^2dt\\\\ \\leq \\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} \\leq |x_2-tx_1|^{2\\sigma}\\}}|x_2-tx_1|^{2\\sigma} |u(t)|^2dt\n\\leq \\int_{\\mathbb{R}}|x_2-tx_1|^{2\\sigma} |u(t)|^2dt,\n\\end{multline*}\nand then notice from (\\ref{eq00}) that there exists a positive constant $C_2>0$ such that for all $(t,x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n+1}$,\n$$|x_2-tx_1|^{2\\sigma} \\leq F(x_2-tx_1)+C_2.$$\nIt follows from (\\ref{eq7}) that \n\\begin{multline}\\label{eq11}\n\\int_{\\{t \\in \\mathbb{R} :\\ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}} \\leq |x_2-tx_1|^{2\\sigma}\\}}|x_1|^{\\frac{2\\sigma}{2\\sigma+1}} |u(t)|^2dt \\\\ \\leq \\|F(x_2-tx_1)^{1\/2}u\\|_{L^2(\\mathbb{R}_t)}^2+C_2\\|u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq\na_0^{-1} \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|u\\|_{L^2(\\mathbb{R}_t)}+C_2\\|u\\|_{L^2(\\mathbb{R}_t)}^2.\n\\end{multline}\nWe deduce from (\\ref{eq10}), (\\ref{eq10.5}) and (\\ref{eq11}) that there exist some positive constants $C_3>0$ and $C_4>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$,\n$$ |x_1|^{\\frac{2\\sigma}{2\\sigma+1}}\\|u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq C_3 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|u\\|_{L^2(\\mathbb{R}_t)}+C_3\\|u\\|_{L^2(\\mathbb{R}_t)}^2,$$\nthat is\n$$|x_1|^{\\frac{2\\sigma}{2\\sigma+1}}\\|u\\|_{L^2(\\mathbb{R}_t)} \\leq C_3 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}+C_3\\|u\\|_{L^2(\\mathbb{R}_t)},$$\nwhich implies that \n\\begin{equation}\\label{eq12}\n \\|\\langle x_1\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq C_4 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2+C_4\\|u\\|_{L^2(\\mathbb{R}_t)}^2,\n\\end{equation}\nwith $\\langle x \\rangle=(1+|x|^2)^{1\/2}$.\nWe shall now prove that there exists a positive constant $C_5>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$,\n\\begin{equation}\\label{eq13}\n\\|\\langle x_1\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2+ \\||x_2-tx_1|^{2\\sigma}u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq C_5 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2+C_5\\|u\\|_{L^2(\\mathbb{R}_t)}^2.\n\\end{equation}\nIn order to do so, let $\\varepsilon_0 \\in \\{\\pm1\\}$ and write the components of the variables $x_1,x_2$ as\n$$x_1=(x_{1,1},x_{1,2},...,x_{1,n}) \\textrm{ and } x_2=(x_{2,1},x_{2,2},...,x_{2,n}).$$ \nLet $j \\in \\{1,...,n\\}$. We shall first study the case when the real parameter \n\\begin{equation}\\label{oyo1}\nx_{1,j} \\neq 0.\n\\end{equation}\nWhile expanding the following $L^2(\\mathbb{R}_t)$ dot-product where $H$ still denotes the Heaviside function\n\\begin{align*}\n& \\ 2\\textrm{Re}\\big(\\tilde{P}u,|x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\big)_{L^2(\\mathbb{R}_t)} \\\\\n=& \\ 2\\textrm{Re}\\big(D_tu,-i |x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\big)_{L^2(\\mathbb{R}_t)} \\\\ \n+ & \\ 2 \\int_{\\mathbb{R}}a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)|x_{2,j}-t x_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt,\n\\end{align*}\nwe notice that \n\\begin{multline*}\n 2\\textrm{Re}\\big(D_tu,-i |x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\big)_{L^2(\\mathbb{R}_t)}\\\\\n= \\big([D_t,-i |x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})]u,u\\big)_{L^2(\\mathbb{R}_t)}. \n\\end{multline*}\nA direct computation gives that \n\\begin{align*}\n& \\ \\big([D_t,-i |x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})]u,u\\big)_{L^2(\\mathbb{R}_t)}\\\\\n=& \\ \\frac{\\varepsilon_0 \\sigma}{\\pi}\\int_{\\mathbb{R}} x_{1,j} |x_{2,j}-tx_{1,j}|^{2\\sigma-1}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt\\\\\n& \\ + \\varepsilon_0\\frac{2^{2\\sigma-1} }{\\pi} \\frac{x_{1,j}}{|x_{1,j}|} \\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}|u(T)|^2,\n\\end{align*}\nwith\n$$T=x_{2,j}x_{1,j}^{-1}-2\\varepsilon_0 x_{1,j}^{-1}\\langle x_1\\rangle^{\\frac{1}{2\\sigma+1}}$$\nand we deduce from (\\ref{eq9}) that \n\\begin{align*}\n& \\ \\big|\\big([D_t,-i |x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})]u,u\\big)_{L^2(\\mathbb{R}_t)}\\big|\\\\\n\\leq & \\ \\frac{2^{2\\sigma-1}}{\\pi}C_1 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}\\\\\n+& \\ \\frac{\\sigma}{\\pi}\\int_{\\mathbb{R}} |x_{1,j}| |x_{2,j}-tx_{1,j}|^{2\\sigma-1}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt.\n\\end{align*}\nSince from (\\ref{eq00.5}), \n\\begin{multline*}\n2a_0\\|F(x_2-tx_1)^{1\/2}|x_{2,j}-t x_{1,j}|^{\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}^2 \\\\ \\leq 2 \\int_{\\mathbb{R}}a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)|x_{2,j}-t x_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt,\n\\end{multline*}\nwe deduce from the Cauchy-Schwarz inequality; and all the previous identities and estimates obtained after (\\ref{oyo1}) that \n\\begin{align*}\n& \\ 2a_0\\|F(x_2-tx_1)^{1\/2}|x_{2,j}-t x_{1,j}|^{\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}^2\\\\\n\\leq & \\ \\frac{2^{2\\sigma-1}}{\\pi}C_1 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}\\\\\n+ & \\ \\frac{\\sigma}{\\pi}\\int_{\\mathbb{R}} |x_{1,j}| |x_{2,j}-tx_{1,j}|^{2\\sigma-1}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt\\\\\n+& \\ 2\\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}\\||x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}.\n\\end{align*}\nSince from (\\ref{eq00}), we have\n$$F(x_2-tx_1)=|x_2-t x_1|^{2\\sigma} \\geq |x_{2,j}-t x_{1,j}|^{2\\sigma},$$\non the support of the function $H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})$, we then deduce from the previous estimate that there exists a positive constant $C_6>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, $x_{1,j} \\neq 0$, \n\\begin{multline}\\label{eq14}\n C_6^{-1}\\||x_{2,j}-t x_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}^2\n\\leq \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2\\\\\n+ \\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2\n+ \\int_{\\mathbb{R}} |x_{1,j}| |x_{2,j}-tx_{1,j}|^{2\\sigma-1}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt.\n\\end{multline}\nBy using that \n$$|x_{2,j}-tx_{1,j}|^{-1} \\leq \\frac{1}{2} \\langle x_1 \\rangle^{-\\frac{1}{2\\sigma+1}},$$\non the support of the function $H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})$, one can estimate from above the following integral as\n\\begin{align*}\n& \\ \\int_{\\mathbb{R}} |x_{1,j}| |x_{2,j}-tx_{1,j}|^{2\\sigma-1}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt \\\\ \n\\leq & \\ 2^{-1} \\int_{\\mathbb{R}} \\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}|x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})|u(t)|^2dt\\\\\n\\leq & \\ 2^{-1} \\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}\\||x_{2,j}-tx_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}.\n\\end{align*}\nAccording to (\\ref{eq14}), this implies that there exists a positive constant $C_7>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in\\mathbb{R}^{4n}$, $x_{1,j} \\neq 0$, \n\\begin{multline}\\label{eq15}\n C_7^{-1}\\||x_{2,j}-t x_{1,j}|^{2\\sigma}H(\\varepsilon_0(x_{2,j}-tx_{1,j})-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})u\\|_{L^2(\\mathbb{R}_t)}^2\\\\\n\\leq \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2\n+ \\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2.\n\\end{multline}\nFinally, since\n\\begin{multline*}\n|x_{2,j}-t x_{1,j}|^{2\\sigma} \\leq |x_{2,j}-t x_{1,j}|^{2\\sigma}H(x_{2,j}-tx_{1,j}-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})\\\\\n+ |x_{2,j}-t x_{1,j}|^{2\\sigma}H(-x_{2,j}+tx_{1,j}-2\\langle x_1 \\rangle^{\\frac{1}{2\\sigma+1}})+2^{2\\sigma}\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}},\n\\end{multline*}\nwe deduce from (\\ref{eq15}) that there exists a positive constant $C_8>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, $x_{1,j} \\neq 0$, \n\\begin{equation}\\label{eq16}\n\\||x_{2,j}-t x_{1,j}|^{2\\sigma}u\\|_{L^2(\\mathbb{R}_t)}^2\\\\\n\\leq C_8 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2 + C_8\\|\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2,\n\\end{equation}\nwhich together with (\\ref{eq12}) proves the a priori estimate\n\\begin{equation}\\label{eq13.11}\n\\|\\langle x_1\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}_t)}^2+ \\||x_{2,j}-tx_{1,j}|^{2\\sigma}u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq C_9 \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2+C_9\\|u\\|_{L^2(\\mathbb{R}_t)}^2,\n\\end{equation}\nwith $C_9>0$ a positive constant; in the case when $x_{1,j} \\neq 0$. Assume now that $x_{1,j}=0$. In this case, a direct computation using (\\ref{eq00.5}) shows that \n\\begin{multline*}\n \\textrm{Re}(\\tilde{P}u,|x_{2,j}|^{2\\sigma}u)_{L^2(\\mathbb{R}_t)}=\\int_{\\mathbb{R}}a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-t x_1)|x_{2,j}|^{2\\sigma}|u(t)|^2dt \\\\\n\\geq a_0\\|F(x_2-t x_1)^{1\/2}|x_{2,j}|^{\\sigma}u\\|_{L^2(\\mathbb{R}_t)}^2,\n\\end{multline*}\nbecause $iD_t$ is a skew-adjoint operator.\nSince \n$$F(x_2-tx_1)=|x_2-t x_1|^{2\\sigma} \\geq |x_{2,j}|^{2\\sigma},$$ \nwhen $|x_{2,j}| \\geq 2$, we first deduce from the Cauchy-Schwarz inequality that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, $x_{1,j}=0$, $|x_{2,j}| \\geq 2$,\n$$ a_0\\||x_{2,j}|^{2\\sigma}u\\|_{L^2(\\mathbb{R}_t)} \\leq \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)},$$\nwhich also implies that \n$$ a_0^2\\||x_{2,j}|^{2\\sigma}u\\|_{L^2(\\mathbb{R}_t)}^2 \\leq \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2+2^{4\\sigma}a_0^2\\|u\\|_{L^2(\\mathbb{R}_t)}^2.$$\nNotice that this second estimate also holds when $|x_{2,j}| \\leq 2$. One can then deduce from another use of (\\ref{eq12}) that the estimate (\\ref{eq13.11}) also holds when $x_{1,j}=0$. This proves the following lemma.\n\n\\bigskip\n\n\\begin{lemma}\\label{prop0}\nConsider the first-order differential operator \n$$\\tilde{P}=iD_t+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1),$$\nwith parameters $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, and $a$ and $F$ standing for the functions defined in \\emph{(\\ref{eq00.5})} and \\emph{(\\ref{eq00})}. Then, there exists a positive constant $C>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_t)$ and $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$,\n\\begin{equation}\\label{eq17}\n\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})u\\|_{L^2(\\mathbb{R}_t)} \\leq C\\big(\\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}+\\|u\\|_{L^2(\\mathbb{R}_t)}\\big).\n\\end{equation}\n\\end{lemma}\n\n\n\\bigskip\n\n\\subsection{From a priori estimates for the one-dimensional operator with parameters to a priori estimates in Wick quantization}\nBy applying Lemma~\\ref{prop0}\n$$\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})u\\|_{L^2(\\mathbb{R}_t)}^2 \\lesssim \\|\\tilde{P}u\\|_{L^2(\\mathbb{R}_t)}^2+\\|u\\|_{L^2(\\mathbb{R}_t)}^2,$$\nto a function $\\Phi(t,x_1,x_2,\\xi_1,\\xi_2) \\in \\mathscr{S}(\\mathbb{R}^{4n+1})$ and integrating this a priori estimate with respect to the variables $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, we obtain that we may find a positive constant $C_1>0$ such that for all $\\Phi \\in \\mathscr{S}(\\mathbb{R}^{4n+1})$,\n\\begin{equation}\\label{eq01}\n\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2 \\leq C_1\\big(\\|\\tilde{P}\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2\\big).\n\\end{equation}\nLet $T>0$ be a positive constant. For any $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe shall consider its wave-packets transform in the variables $x_1,x_2$ with parameter $0<\\lambda \\leq 1$ defined as\n$$(W_{\\lambda}u)(t,Y)=\\big(u(t,\\cdot),\\varphi_{Y}^{\\lambda}\\big)_{L^2(\\mathbb{R}_{x_1,x_2}^{2n})} \\in \\mathscr{S}(\\mathbb{R}^{4n+1}), \\ Y=(y_1,y_2;\\eta_1,\\eta_2) \\in \\mathbb{R}^{4n},$$\nwith \n$$\\varphi_{Y}^{\\lambda}(x_1,x_2)=(2\\lambda)^{\\frac{n}{2}}e^{-\\pi \\lambda (|x_1-y_1|^2+|x_2-y_2|^2)}e^{2i \\pi((x_1-y_1)\\cdot \\eta_1+(x_2-y_2)\\cdot \\eta_2)}.$$\nWe apply the a priori estimate \n$$\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2 \\leq C_1\\big(\\|\\tilde{P}\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|\\Phi\\|_{L^2(\\mathbb{R}^{4n+1})}^2\\big),$$\nto the function $$\\Phi(t,X)=(W_{\\lambda}u)(t,X),$$ \nwith $X=(x_1,x_2;\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$. By using the fact that the wave-packets transform is an isometric mapping from $L^2(\\mathbb{R}_{x_1,x_2}^{2n})$ to $L^2(\\mathbb{R}_{x_1,x_2,\\xi_1,\\xi_2}^{4n})$, see appendix (Section~\\ref{appendix}), we obtain that \n\\begin{multline}\\label{eq02}\n\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2 \\\\ \\leq C_1\\big(\\|\\tilde{P}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}^2\\big).\n\\end{multline}\nWe recall from the appendix (Section~\\ref{appendix}) that the Wick quantization with parameter $0<\\lambda \\leq 1$ of a symbol $a$ is formally given by\n\\begin{equation}\\label{eq02.5}\na^{\\textrm{Wick}(\\lambda)}=W_{\\lambda}^*a^{\\mu}W_{\\lambda},\\ 1^{\\textrm{Wick}(\\lambda)}=W_{\\lambda}^*W_{\\lambda}=\\textrm{Id}_{L^2},\n\\end{equation}\nwhere $a^{\\mu}$ stands for the multiplication operator by the function $a$ on $L^2$, and that the operator \n$$\\pi_{\\lambda}=W_{\\lambda}W_{\\lambda}^*,$$ \nis an orthogonal projection on a closed proper subspace of $L^2$. It follows that \n\\begin{align*}\n& \\ \\|\\pi_{\\lambda}(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}\n+ \\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2 \\\\\n\\leq & \\ \\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2.\n\\end{align*}\nMoreover, since the wave-packets transform is an isometric mapping from $L^2(\\mathbb{R}_{x_1,x_2}^{2n})$ to $L^2(\\mathbb{R}_{x_1,x_2,\\xi_1,\\xi_2}^{4n})$, we may write that \n\\begin{align*}\n& \\ \\|\\pi_{\\lambda}(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|W_{\\lambda}W_{\\lambda}^*(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{align*}\nIt follows from (\\ref{eq02}) that \n\\begin{align*}\n& \\ \\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}^2 \\\\\n& \\ +\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2\\\\\n\\leq & \\ 2C_1\\big(\\|\\pi_{\\lambda}\\tilde{P}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|(1-\\pi_{\\lambda})\\tilde{P}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}^2\\big).\n\\end{align*}\nBy using similar arguments as previously, namely (\\ref{eq02.5}) and the fact that the wave-packets transform is an isometric mapping from $L^2(\\mathbb{R}_{x_1,x_2}^{2n})$ to $L^2(\\mathbb{R}_{x_1,x_2,\\xi_1,\\xi_2}^{4n})$, together with recalling that we only consider here the Wick quantization the variables $x_1,x_2$, and not in the $t$-variable, we obtain that \n\\begin{align*}\n& \\ \\|\\pi_{\\lambda}\\tilde{P}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|W_{\\lambda}W_{\\lambda}^*\\tilde{P}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|W_{\\lambda}W_{\\lambda}^*[iD_t+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|W_{\\lambda}(iD_t+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)})u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n= & \\ \\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{align*}\nOn the other hand, notice that \n\\begin{multline*}\n(1-\\pi_{\\lambda})\\tilde{P}W_{\\lambda}=(1-\\pi_{\\lambda})[iD_t+a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]W_{\\lambda}\\\\\n=(1-\\pi_{\\lambda})a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)W_{\\lambda},\n\\end{multline*}\nbecause \n$$(1-\\pi_{\\lambda})iD_tW_{\\lambda}=(1-W_{\\lambda}W_{\\lambda}^*)W_{\\lambda}iD_t=W_{\\lambda}(1-W_{\\lambda}^*W_{\\lambda})iD_t=0,$$\nsince $D_t$ commutes with the wave-packets transform in the variables $x_1,x_2$, and that $W_{\\lambda}^*W_{\\lambda}=\\textrm{Id}_{L^2}$.\nWe may then write that \n\\begin{multline*}\n(1-\\pi_{\\lambda})a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)W_{\\lambda}=a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)W_{\\lambda}\\\\\n-[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]W_{\\lambda}-a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)\\pi_{\\lambda}W_{\\lambda}.\n\\end{multline*}\nSince from (\\ref{eq02.5}), \n$$\\pi_{\\lambda}W_{\\lambda}=W_{\\lambda}W_{\\lambda}^*W_{\\lambda}=W_{\\lambda},$$\nit follows that \n$$(1-\\pi_{\\lambda})a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)W_{\\lambda}=-[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]W_{\\lambda}.$$\nWe then deduce that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\n (2C_1)^{-1}\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}^2 \\\\\n + (2C_1)^{-1}\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2\\\\\n\\leq \\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}^2\\\\\n + \\|[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}^2.\n\\end{multline*}\nWe now need to study the commutator term\n\\begin{multline*}\n[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]=[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)]F(x_2-tx_1)\\\\ +a(t,\\xi_2,\\xi_1+t\\xi_2)[\\pi_{\\lambda},F(x_2-tx_1)].\n\\end{multline*}\nIn order to do so, we recall from (\\ref{yo1}) in appendix that the kernel of the orthogonal projection $\\pi_{\\lambda}$ is given by \n\\begin{equation}\\label{EQ1}\ne^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}e^{i\\pi (x-y)\\cdot(\\xi+ \\eta)},\n\\end{equation} \nwith \n\\begin{equation}\\label{EQ2}\n\\Gamma_{\\lambda}(X)=\\lambda|x|^2+\\frac{|\\xi|^2}{\\lambda}, \\ X=(x,\\xi) \\in \\mathbb{R}^{4n}; \\ x=(x_1,x_2),\\ \\xi=(\\xi_1,\\xi_2) \\in \\mathbb{R}^{2n}.\n\\end{equation}\n\n\n\\bigskip\n\n\n\\begin{lemma}\\label{LEM1}\nLet $T>0$ be a positive constant. Then, there exists a positive constant $C>0$ such that for all $0<\\lambda \\leq 1$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\emph{\\textrm{supp }} u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\n\\|[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)]F(x_2-tx_1)W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\\\\n\\leq C\\lambda^{1\/2}\\|\\langle x_2-tx_1\\rangle^{2\\sigma}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}+C\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline*}\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{LEM1}}. Notice from (\\ref{EQ1}) that the kernel of the commutator \n$$[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)],$$ \nis given by\n$$K_{t,\\lambda}(X,Y)=e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}e^{i\\pi (x-y)\\cdot(\\xi+ \\eta)}\\big(a(t,\\eta_2,\\eta_1+t\\eta_2)-a(t,\\xi_2,\\xi_1+t\\xi_2)\\big).$$\nRecalling that $a$ is a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function and therefore a Lipschitz function, we may therefore find a positive constant $C_2>0$ such that for all $t \\in [-T,T]$, $0<\\lambda \\leq 1$ and $X,Y \\in \\mathbb{R}^{4n}$,\n\\begin{equation}\\label{E0}\n|K_{t,\\lambda}(X,Y)| \\leq C_2e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}|\\eta-\\xi|.\n\\end{equation}\nNotice that \n\\begin{multline*}\n\\int_{\\mathbb{R}^{4n}}|K_{t,\\lambda}(X,Y)|dY \\leq C_2\\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}|\\eta-\\xi|dY=C_2\\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(Y)}|\\eta|dY\\\\\n=C_2\\lambda^{1\/2}\\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}|Y|^2}|\\eta|dY=C_3\\lambda^{1\/2},\n\\end{multline*}\nwith $C_3>0$ a positive constant. \nBy symmetry of the estimate (\\ref{E0}), we also have \n$$\\int_{\\mathbb{R}^{4n}}|K_{t,\\lambda}(X,Y)|dX \\leq C_3\\lambda^{1\/2}.$$\nSchur test for integral operators together with (\\ref{eq00}) show that there exists a positive constant $C_4>0$ such that for all $0<\\lambda \\leq 1$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\n\\|[\\pi_{\\lambda},a(t,\\xi_2,\\xi_1+t\\xi_2)]F(x_2-tx_1)W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\leq C_3\\lambda^{1\/2}\\|F(x_2-tx_1)W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}\\\\\n\\leq C_4\\lambda^{1\/2}\\|\\langle x_2-tx_1\\rangle^{2\\sigma}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}+C_4\\|u\\|_{L^2(\\mathbb{R}^{2n+1})},\n\\end{multline*}\nsince we recall that for any $0<\\lambda \\leq 1$,\n$$\\|W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}=\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}. \\ \\Box$$\n\n\n\\bigskip\n\n\n\\begin{lemma}\\label{LEM2}\nLet $T>0$ be a positive constant. Then, there exists a positive constant $C>0$ such that for all $0<\\lambda \\leq 1$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\emph{\\textrm{supp }} u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\n\\|a(t,\\xi_2,\\xi_1+t\\xi_2)[\\pi_{\\lambda},F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\\\\n\\leq C \\lambda^{-\\frac{1+(2\\sigma-1)_+}{2}} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})},\n\\end{multline*}\nwith $(2\\sigma-1)_+=\\emph{\\textrm{max}}(2\\sigma-1,0)$.\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{LEM2}}. We first notice that \n$$\\|a(t,\\xi_2,\\xi_1+t\\xi_2)[\\pi_{\\lambda},F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\lesssim \\|[\\pi_{\\lambda},F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})},$$\nsince $a$ is a $L^{\\infty}(\\mathbb{R}^{2n+1})$ function. Arguing as in previous lemma, we notice that the kernel of the commutator $[\\pi_{\\lambda},F(x_2-tx_1)]$ \nis given by\n$$K_{t,\\lambda}(X,Y)=e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}e^{i\\pi (x-y)\\cdot(\\xi+ \\eta)}\\big(F(y_2-ty_1)-F(x_2-tx_1)\\big).$$\nRecall from (\\ref{eq00}) that there exists a positive constant $C_5>0$ such that for all $x \\in \\mathbb{R}^n$,\n\\begin{equation}\\label{eq271}\n|\\nabla F(x)| \\leq C_5 \\langle x \\rangle^{(2\\sigma-1)_+}\n\\end{equation}\nwith $(2\\sigma-1)_+=\\textrm{max}(2\\sigma -1,0)$. Writing \n\\begin{multline*}\nF(y_2-ty_1)-F(x_2-tx_1)\\\\ =\\big(y_2-x_2-t(y_1-x_1)\\big) \\cdot \\int_0^1\\nabla F\\big((1-\\theta)(x_2-y_2-t(x_1-y_1))+y_2-ty_1)\\big)d\\theta,\n\\end{multline*}\nwe have for all $(t,x_1,x_2,y_1,y_2) \\in [-T,T] \\times \\mathbb{R}^{4n}$,\n\\begin{align*}\n& \\ |F(y_2-ty_1)-F(x_2-tx_1)|\\\\\n \\lesssim & \\ |x-y| \\int_0^1\\langle (1-\\theta)(x_2-y_2-t(x_1-y_1))+y_2-ty_1)\\rangle^{(2\\sigma-1)_+} d\\theta\\\\\n\\lesssim & \\ |x-y| \\int_0^1 \\langle (1-\\theta)(x_2-y_2-t(x_1-y_1))\\rangle^{(2\\sigma-1)_+} \\langle y_2-ty_1\\rangle^{(2\\sigma-1)_+} d\\theta\\\\\n\\lesssim & \\ |x-y| \\langle y_2-ty_1\\rangle^{(2\\sigma-1)_+} \\langle x_2-y_2-t(x_1-y_1)\\rangle^{(2\\sigma-1)_+}.\n\\end{align*}\nSetting\n$$\\tilde{K}_{t,\\lambda}(X,Y)=\\langle y_2-ty_1\\rangle^{-(2\\sigma-1)_+}K_{t,\\lambda}(X,Y),$$\nwe have \n\\begin{equation}\\label{E2}\n|\\tilde{K}_{t,\\lambda}(X,Y)| \\lesssim e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)} |x-y|\\langle x_2-y_2-t(x_1-y_1)\\rangle^{(2\\sigma-1)_+}.\n\\end{equation}\nWhile using a change of variables, we notice that for all $t \\in [-T,T]$ and $0<\\lambda \\leq 1$,\n\\begin{align*}\n\\int_{\\mathbb{R}^{4n}}|\\tilde{K}_{t,\\lambda}(X,Y)|dY \\lesssim & \\ \\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}|x-y|\\langle x_2-y_2-t(x_1-y_1)\\rangle^{(2\\sigma-1)_+}dY\\\\\n\\lesssim & \\ \\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(Y)}|y|\\langle y_2-ty_1\\rangle^{(2\\sigma-1)_+}dY \\\\\n\\lesssim & \\ \\int_{\\mathbb{R}^{4n}}e^{-\\frac{\\pi}{2}|Y|^2}|\\lambda^{-1\/2}y|\\langle \\lambda^{-1\/2}(y_2-ty_1)\\rangle^{(2\\sigma-1)_+}dY \\lesssim \\lambda^{-\\frac{1+(2\\sigma-1)_+}{2}},\n\\end{align*}\nsince we have $\\langle \\mu x \\rangle \\leq \\mu \\langle x \\rangle$, when $\\mu \\geq 1$. By symmetry of the estimate (\\ref{E2}), we also have \n$$\\int_{\\mathbb{R}^{4n}}|\\tilde{K}_{t,\\lambda}(X,Y)|dX \\lesssim \\lambda^{-\\frac{1+(2\\sigma-1)_+}{2}}.$$\nSchur test for integral operators then shows that for all $0<\\lambda \\leq 1$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n$$\\|[\\pi_{\\lambda},F(x_2-tx_1)]W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\lesssim \\lambda^{-\\frac{1+(2\\sigma-1)_+}{2}} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})},$$\nwhich proves Lemma~\\ref{LEM2}.~$\\Box$\n\n\n\\bigskip\n\n\\noindent\nWe then deduce from Lemmas~\\ref{LEM1} and \\ref{LEM2} that there exists a positive constant $C_6>0$ such that for all $0<\\lambda \\leq 1$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{align*}\n& \\ C_6^{-1}\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}^2 \\\\\n+ & \\ C_6^{-1}\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2\\\\\n\\leq & \\ \\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}^2+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}^2\\\\\n+& \\ \\lambda^{-1-(2\\sigma-1)_+} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2+\\lambda\\|\\langle x_2-tx_1\\rangle^{2\\sigma}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2.\n\\end{align*}\nBy choosing a positive constant $0<\\lambda_0 \\leq 1$ such that \n$$\\lambda_0 \\leq \\frac{1}{2C_6},$$\nwe notice that one may estimate from above the following term as \n$$\\lambda\\|\\langle x_2-tx_1\\rangle^{2\\sigma}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2 \\leq (2C_6)^{-1}\\|(1+\\langle x_1 \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle x_2-tx_1 \\rangle^{2\\sigma})W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}^2,$$\nfor all $0<\\lambda \\leq \\lambda_0$. We then notice that for any $\\varepsilon>0$, there exists a positive constant $C_{\\varepsilon}>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying\n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n$$ \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})} \\leq \\varepsilon\\|\\langle x_2-tx_1\\rangle^{2\\sigma}W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}+C_{\\varepsilon}\\|W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}.$$ By recalling that \n$$\\|W_{\\lambda}u\\|_{L^2(\\mathbb{R}^{4n+1})}=\\|u\\|_{L^2(\\mathbb{R}^{2n+1})},$$ \nwe finally deduce from these estimates that for any $T>0$, there exists a positive constant $c_T>0$ independent of the parameter $0<\\lambda \\leq \\lambda_0$; and a second positive constant $C_T(\\lambda)>0$, which may depend on $\\lambda$ such that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline}\\label{eq18}\n\\|[\\langle x_1\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})} +\\|[\\langle x_2-tx_1\\rangle^{2\\sigma}]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\\\ \\leq c_T\\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}+C_T(\\lambda)\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline}\n\n\\bigskip\n\n\n\\subsection{From a priori estimates in Wick quantization to a priori estimates in Weyl quantization}\nIn the previous section, we established an a priori estimate with symbols quantized in the Wick quantization with parameter $0<\\lambda \\leq \\lambda_0 \\leq 1$. We shall need the following lemma to estimate error terms incurred by coming back from Wick quantization with parameter $0<\\lambda \\leq \\lambda_0$ to the Weyl quantization for symbols appearing in the left-hand-side of (\\ref{eq18}).\n\n\\bigskip\n\n\n\\begin{lemma}\\label{lem1}\nLet $a \\in C^{\\infty}(\\mathbb{R}^{2n})$ be a symbol whose derivatives of order $ \\geq 2$ are bounded on $\\mathbb{R}^{2n}$. Then, there exists a positive constant $C>0$ depending only on the $L^{\\infty}$-norms of a finite number of derivatives of order greater or equal to 2 of the symbol $a$; such that for all $\\lambda > 0$ and $u \\in \\mathscr{S}(\\mathbb{R}^n)$,\n$$\\|a^{\\emph{Wick}(\\lambda)}u-a^wu\\|_{L^2(\\mathbb{R}^n)} \\leq C\\Big(\\lambda+\\frac{1}{\\lambda}\\Big)\\|u\\|_{L^2(\\mathbb{R}^n)}.$$\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lem1}.} We notice from (\\ref{lay1}) and (\\ref{lay2}) in appendix that one may write that\n$$a^{\\textrm{Wick}(\\lambda)}=a^w+R_{\\lambda}^w,$$\nwith \n$$R_{\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{2n}}(1-\\theta)a''(X+\\theta Y)Y^2e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta, \\ X=(x,\\xi) \\in \\mathbb{R}^{2n},$$\nwhere \n$$\\Gamma_{\\lambda}(Y)=\\lambda |y|^2+\\frac{|\\eta|^2}{\\lambda}, \\ Y=(y,\\eta) \\in \\mathbb{R}^{2n}.$$\nIt follows that any derivative of the symbol $R_{\\lambda}$, \n$$\\partial_{X}^{\\alpha}R_{\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{2n}}(1-\\theta)\\partial_X^{\\alpha}a''(X+\\theta Y)Y^2e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta,$$\ncan be estimated from above as\n$$\\|\\partial_{X}^{\\alpha}R_{\\lambda}\\|_{L^{\\infty}(\\mathbb{R}^{2n})} \\leq \\|\\partial_X^{\\alpha}a''\\|_{L^{\\infty}(\\mathbb{R}^{2n})}\\int_{\\mathbb{R}^{2n}}(|y|^2+|\\eta|^2)e^{-2\\pi\\lambda |y|^2}e^{-2\\pi\\frac{|\\eta|^2}{\\lambda}}2^ndyd\\eta.$$\nSince \n\\begin{multline*}\n\\int_{\\mathbb{R}^{2n}}(|y|^2+|\\eta|^2)e^{-2\\pi\\lambda |y|^2}e^{-2\\pi\\frac{|\\eta|^2}{\\lambda}}2^ndyd\\eta\\\\\n=\\int_{\\mathbb{R}^{2n}}\\Big(\\frac{|y|^2}{\\lambda}+\\lambda|\\eta|^2\\Big)e^{-2\\pi (|y|^2+|\\eta|^2)}2^ndyd\\eta \n=\\mathcal{O}\\Big(\\lambda+\\frac{1}{\\lambda}\\Big),\n\\end{multline*}\nLemma~\\ref{lem1} is then a direct consequence of the $L^2$ continuity theorem in the class $S_{00}^0$.~$\\Box$\n\n\\bigskip\n\n\n\\noindent\nNotice that we may apply Lemma~\\ref{lem1} to the two symbols seen as functions of the variables $(x_1,x_2,\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n}$, \n$$\\langle x_1\\rangle^{\\frac{2\\sigma}{2\\sigma+1}} \\textrm{ and } \\langle x_2-tx_1\\rangle^{2\\sigma}\\chi_0(t),$$\nwith $\\chi_0 \\in C_0^{\\infty}(\\mathbb{R})$, $\\chi_0=1$ on $[-T,T]$, since $0<\\sigma<1$. We recall that the $t$-variable is seen as a parameter and that \nwe only consider here the Wick and Weyl quantizations in the variables $x_1,x_2$. It follows from Lemma~\\ref{lem1} and (\\ref{eq18}) that for any fixed $T>0$, there exist a positive constant $c_T>0$ independent of the parameter $0<\\lambda \\leq \\lambda_0$, and a second positive constant $C_T(\\lambda)>0$, which may depend on $\\lambda$ such that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline}\\label{eq19}\n\\big\\|(1+|x_2-tx_1|^{2\\sigma}+|x_1|^{\\frac{2\\sigma}{2\\sigma+1}})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\\\ \\leq c_T\\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}u\\|_{L^2(\\mathbb{R}^{2n+1})}+C_T(\\lambda)\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline}\nWe shall now establish a priori estimates for error terms incurred by coming back from Wick quantization with parameter $0< \\lambda \\leq \\lambda_0$ to the Weyl quantization for the symbol $$a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1).$$ \n\n\n\\bigskip\n\n\n\\begin{lemma}\\label{prop1}\nLet $T>0$ be a positive constant. Then, there exists a positive constant $C>0$ such that for all $0< \\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\emph{\\textrm{supp }} u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\nC^{-1}\\|[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\emph{\\textrm{Wick}}(\\lambda)}u-[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^wu\\|_{L^2(\\mathbb{R}^{2n+1})} \\\\\n\\leq \\lambda^{1-\\sigma} \\|\\langle x_2-tx_1\\rangle^{2\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})}+ \\lambda^{-\\sigma} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}u\\|_{L^2(\\mathbb{R}^{2n+1})}+ \\lambda^{-1}\\|u\\|_{L^2(\\mathbb{R}^{2n+1})},\n\\end{multline*}\nwith $(2\\sigma-1)_+=\\emph{\\textrm{max}}(2\\sigma-1,0)$.\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{prop1}.} Let $\\chi_0$ be a $C_0^{\\infty}(\\mathbb{R})$ function satisfying $\\chi_0=1$ on $[-T,T]$. By using again (\\ref{lay1}) and (\\ref{lay2}) in appendix, one may write that\n\\begin{multline}\\label{eq20}\n\\chi_0(t)[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{\\textrm{Wick}(\\lambda)}\\\\ =\\chi_0(t)[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^w +R_{t,\\lambda}^w,\n\\end{multline}\nwith \n$$R_{t,\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{4n}}(1-\\theta)\\chi_0(t)r_t''(X+\\theta Y)Y^2e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta, $$\nwhere \n$$r_t(x_1,x_2;\\xi_1,\\xi_2)=a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1), \\ X=(x_1,x_2;\\xi_1,\\xi_2) \\in \\mathbb{R}^{4n},$$\nand\n$$\\Gamma_{\\lambda}(Y)=\\lambda (|y_1|^2+|y_2|^2)+\\frac{1}{\\lambda}(|\\eta_1|^2+|\\eta_2|^2), \\ Y=(y_1,y_2;\\eta_1,\\eta_2) \\in \\mathbb{R}^{4n}.$$\nDefine\n\\begin{equation}\\label{eq23}\n\\tilde{r}_{1,t}(X,Y,\\theta)=\\chi_0(t)a\\big(t,\\xi_2+\\theta \\eta_2,\\xi_1+t\\xi_2+\\theta(\\eta_1+t \\eta_2)\\big) (\\nabla_x^2B_t)(x+\\theta y).y^2,\n\\end{equation}\n\\begin{equation}\\label{eq22}\n\\tilde{r}_{2,t}(X,Y,\\theta)=\\chi_0(t)(\\nabla_{\\xi}A_t)(\\xi+\\theta \\eta).\\eta \\ (\\nabla_xB_t)(x+\\theta y).y,\n\\end{equation}\n\\begin{equation}\\label{eq21}\n\\tilde{r}_{3,t}(X,Y,\\theta)=\\chi_0(t)(\\nabla_{\\xi}^2A_t)(\\xi+\\theta \\eta).\\eta^2 \\ F(x_2-tx_1+\\theta(y_2-ty_1)),\n\\end{equation}\nwith\n\\begin{equation}\\label{eq5678}\nA_t(\\xi)=a(t,\\xi_2,\\xi_1+t\\xi_2), \\ \\xi=(\\xi_1,\\xi_2) \\textrm{ and } B_t(x)=F(x_2-tx_1), \\ x=(x_1,x_2).\n\\end{equation}\nWe also define\n\\begin{equation}\\label{eq24}\nR_{1,t,\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{4n}}(1-\\theta)\\tilde{r}_{1,t}(X,Y,\\theta)e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta,\n\\end{equation}\n\\begin{equation}\\label{eq25}\nR_{2,t,\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{4n}}(1-\\theta)\\tilde{r}_{2,t}(X,Y,\\theta)e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta,\n\\end{equation}\n\\begin{equation}\\label{eq26}\nR_{3,t,\\lambda}(X)=\\int_0^1\\int_{\\mathbb{R}^{4n}}(1-\\theta)\\tilde{r}_{3,t}(X,Y,\\theta)e^{-2\\pi\\Gamma_{\\lambda}(Y)}2^ndYd\\theta.\n\\end{equation}\nRecall from (\\ref{eq00}) that there exists a positive constant $C_1>0$ such that for all $x \\in \\mathbb{R}^n$,\n\\begin{equation}\\label{eq27}\n|F(x)| \\leq C_1 \\langle x \\rangle^{2\\sigma}, \\ \\ |F'(x)| \\leq C_1 \\langle x \\rangle^{(2\\sigma-1)_+}, \\ \\ \\forall k \\geq 2, \\ F^{(k)} \\in L^{\\infty}(\\mathbb{R}^n),\n\\end{equation}\nwith $(2\\sigma-1)_+=\\textrm{max}(2\\sigma-1,0)$, since $0<\\sigma <1$. \n\n\n\\bigskip\n\n\n\\begin{lemma}\\label{lem2}\nThere exists a positive constant $C>0$ such that \n$$\\forall \\ 0<\\lambda \\leq \\lambda_0, \\forall u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1}),\\ \\|R_{1,t,\\lambda}^wu\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C \\lambda^{-1}\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.$$\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lem2}}. According to the $L^2$ continuity theorem in the class $S_{00}^0$, it is sufficient to prove that for all $\\alpha \\in \\mathbb{N}^{4n}$, there exists a positive constant $C_{2,\\alpha}>0$ such that \n$$\\forall t \\in \\mathbb{R}, \\forall \\ 0< \\lambda \\leq \\lambda_0,\\ \\|\\partial_X^{\\alpha}R_{1,t,\\lambda}\\|_{L^{\\infty}(\\mathbb{R}_X^{4n})} \\leq C_{2,\\alpha}\\lambda^{-1}.$$\nRecalling that $a$ is a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function, we deduce from (\\ref{eq23}), (\\ref{eq5678}), (\\ref{eq24}), (\\ref{eq27}) and a change of variables that for any $\\alpha \\in \\mathbb{N}^{4n}$, there exist some positive constants $C_{3,\\alpha},C_{4,\\alpha}>0$ such that\n\\begin{multline*}\n\\forall t \\in \\mathbb{R}, \\forall \\ 0<\\lambda \\leq \\lambda_0,\\ \\|\\partial_X^{\\alpha}R_{1,t,\\lambda}\\|_{L^{\\infty}(\\mathbb{R}_X^{4n})} \\\\ \\leq C_{3,\\alpha}\\int_{\\mathbb{R}^{4n}}\\big(|y_1|^2+|y_2|^2\\big)\ne^{-2\\pi(\\lambda (|y_1|^2+|y_2|^2)+\\lambda^{-1}(|\\eta_1|^2+|\\eta_2|^2))}dy_1dy_2d\\eta_1d\\eta_2=C_{4,\\alpha}\\lambda^{-1}.\n\\end{multline*}\n\n\n\\bigskip\n\n\n\\begin{lemma}\\label{lem3}\nLet $T>0$ be a positive constant. Then, there exists a positive constant $C>0$ such that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\emph{\\textrm{supp }} u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n$$\\|R_{2,t,\\lambda}^wu\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C \\lambda^{-\\sigma} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}u\\|_{L^2(\\mathbb{R}^{2n+1})},$$\nwith $(2\\sigma-1)_+=\\emph{\\textrm{max}}(2\\sigma-1,0) \\geq 0.$ \n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lem3}}. We notice from (\\ref{eq5678}) and (\\ref{eq27}) that \n\\begin{multline*}\n\\big| (\\nabla_xB_t)(x+\\theta y)\\big| \\lesssim \\langle x_2-tx_1+\\theta(y_2-ty_1)\\rangle^{(2\\sigma -1)_+}\\\\ \\lesssim \\langle x_2-tx_1\\rangle^{(2\\sigma -1)_+}\n \\langle \\theta(y_2-ty_1)\\rangle^{(2\\sigma -1)_+}\n \\lesssim \\langle x_2-tx_1\\rangle^{(2\\sigma -1)_+} \\langle y_2-ty_1\\rangle^{(2\\sigma -1)_+},\n \\end{multline*}\nwhen $0 \\leq \\theta \\leq 1$ and $t \\in \\textrm{supp }\\chi_0$. \nRecalling that $a$ is a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function, it follows from (\\ref{eq22}), (\\ref{eq5678}) and (\\ref{eq25}) that for all $0<\\lambda \\leq \\lambda_0$,\n\\begin{align*}\n|R_{2,t,\\lambda}(X)| \\lesssim & \\ \\chi_0(t) \\langle x_2-tx_1\\rangle^{(2\\sigma -1)_+}\\int_{\\mathbb{R}^{4n}}|\\eta||y|\\langle y_2-ty_1\\rangle^{(2\\sigma -1)_+}e^{-2\\pi\\Gamma_{\\lambda}(Y)}dY \\\\\n\\lesssim & \\ \\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -1)_+}\\int_{\\mathbb{R}^{4n}}|\\eta||y|\\langle \\lambda^{-\\frac{1}{2}}(y_2-ty_1)\\rangle^{(2\\sigma -1)_+}e^{-2\\pi|Y|^2}dY\\\\\n\\lesssim & \\ \\lambda^{-\\frac{(2\\sigma -1)_+}{2}}\\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -1)_+},\n\\end{align*}\nsince we have $\\langle \\mu x \\rangle \\leq \\mu \\langle x \\rangle$, when $\\mu \\geq 1$. \nAfter differentiating (\\ref{eq22}) with respect to the variable $X=(x,\\xi)$, and using the same kind of estimates, we obtain from (\\ref{eq27}) that for all $0<\\lambda \\leq \\lambda_0$,\n\\begin{align*}\n& \\ |\\partial_{x}^{\\alpha}\\partial_{\\xi}^{\\beta}R_{2,t,\\lambda}(X)| \\\\\n\\lesssim & \\ \\chi_0(t) \\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|-1)_+}\\int_{\\mathbb{R}^{4n}}|\\eta||y|\\langle y_2-ty_1\\rangle^{(2\\sigma -|\\alpha|-1)_+}e^{-2\\pi\\Gamma_{\\lambda}(Y)}dY \\\\\n\\lesssim & \\ \\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|-1)_+}\\int_{\\mathbb{R}^{4n}}|\\eta||y|\\langle \\lambda^{-\\frac{1}{2}}(y_2-ty_1)\\rangle^{(2\\sigma -|\\alpha|-1)_+}e^{-2\\pi|Y|^2}dY\\\\\n\\lesssim & \\ \\lambda^{-\\frac{(2\\sigma -|\\alpha|-1)_+}{2}}\\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|-1)_+},\n\\end{align*}\nwith \n$$(2\\sigma -|\\alpha|-1)_+=\\textrm{max}(2\\sigma -|\\alpha|-1,0).$$ \nIt follows from those estimates that the Weyl symbol\n$$\\tilde{a}_{\\lambda}(t,X)=\\lambda^{\\sigma}R_{2,t,\\lambda}(X) \\sharp[\\chi_0(t) \\langle x_2-tx_1\\rangle^{-(2\\sigma-1)_+}],$$ \nof the operator obtained by composition\n$$\\lambda^{\\sigma}R_{2,t,\\lambda}^w [\\chi_0(t) \\langle x_2-tx_1\\rangle^{-(2\\sigma-1)_+}]^w,$$\nbelongs to the class $S(1,dt^2+d\\tau^2+dX^2)$ uniformly with respect to the parameter $0<\\lambda \\leq \\lambda_0$. We may then deduce from the $L^2$ continuity theorem in the class $S_{00}^0$ that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{align*}\n\\|R_{2,t,\\lambda}^wu\\|_{L^2(\\mathbb{R}^{2n+1})} = & \\ \\|\\lambda^{-\\sigma}\\tilde{a}_{\\lambda}^w\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\\\ \\lesssim & \\ \n\\lambda^{-\\sigma} \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{align*}\nThis ends the proof of Lemma~\\ref{lem3}.~$\\Box$\n\n\\bigskip\n\n\n\\begin{lemma}\\label{lem4}\nLet $T>0$ be a positive constant. Then, there exists a positive constant $C>0$ such that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\emph{\\textrm{supp }} u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n$$\\|R_{3,t,\\lambda}^wu\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C \\lambda^{1-\\sigma} \\|\\langle x_2-tx_1\\rangle^{2\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})}.$$ \n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lem4}}. By using similar kind of estimates as in the previous lemma together with the fact that \n\\begin{multline*}\n|F(x_2-tx_1+\\theta(y_2-ty_1))| \\lesssim \\langle x_2-tx_1+\\theta(y_2-ty_1)\\rangle^{2\\sigma}\\\\ \\lesssim \\langle x_2-tx_1\\rangle^{2\\sigma}\n \\langle \\theta(y_2-ty_1)\\rangle^{2\\sigma}\n \\lesssim \\langle x_2-tx_1\\rangle^{2\\sigma} \\langle y_2-ty_1\\rangle^{2\\sigma},\n \\end{multline*}\nwhen $0 \\leq \\theta \\leq 1$;\nit follows from (\\ref{eq21}), (\\ref{eq5678}) and (\\ref{eq26}) that for all $0<\\lambda \\leq \\lambda_0$,\n\\begin{align*}\n|R_{3,t,\\lambda}(X)| \\lesssim & \\ \\chi_0(t) \\langle x_2-tx_1\\rangle^{2\\sigma}\\int_{\\mathbb{R}^{4n}}|\\eta|^2\\langle y_2-ty_1\\rangle^{2\\sigma}e^{-2\\pi\\Gamma_{\\lambda}(Y)}dY \\\\\n\\lesssim & \\ \\chi_0(t)\\langle x_2-tx_1\\rangle^{2\\sigma}\\int_{\\mathbb{R}^{4n}}\\lambda|\\eta|^2\\langle \\lambda^{-\\frac{1}{2}}(y_2-ty_1)\\rangle^{2\\sigma}e^{-2\\pi|Y|^2}dY\\\\\n\\lesssim & \\ \\lambda^{1-\\sigma}\\chi_0(t)\\langle x_2-tx_1\\rangle^{2\\sigma}.\n\\end{align*}\nAfter differentiating (\\ref{eq21}) and (\\ref{eq26}) with respect to the variable $X=(x,\\xi)$ and using similar types of estimates, we obtain from (\\ref{eq27}) that for all $0<\\lambda \\leq \\lambda_0$,\n\\begin{align*}\n& \\ |\\partial_{x}^{\\alpha}\\partial_{\\xi}^{\\beta}R_{3,t,\\lambda}(X)| \\\\\n\\lesssim & \\ \\chi_0(t) \\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|)_+}\\int_{\\mathbb{R}^{4n}}|\\eta|^2\\langle y_2-ty_1\\rangle^{(2\\sigma -|\\alpha|)_+}e^{-2\\pi\\Gamma_{\\lambda}(Y)}dY \\\\\n\\lesssim & \\ \\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|)_+}\\int_{\\mathbb{R}^{4n}}\\lambda|\\eta|^2\\langle \\lambda^{-\\frac{1}{2}}(y_2-ty_1)\\rangle^{(2\\sigma -|\\alpha|)_+}e^{-2\\pi|Y|^2}dY\\\\\n\\lesssim & \\ \\lambda^{1-\\frac{(2\\sigma -|\\alpha|)_+}{2}}\\chi_0(t)\\langle x_2-tx_1\\rangle^{(2\\sigma -|\\alpha|)_+}.\n\\end{align*} \nIt follows from those estimates that the Weyl symbol\n$$\\tilde{a}_{\\lambda}(t,X)=\\lambda^{\\sigma-1}R_{3,t,\\lambda}(X) \\sharp[\\chi_0(t) \\langle x_2-tx_1\\rangle^{-2\\sigma}],$$ \nof the operator obtained by composition\n$$\\lambda^{\\sigma-1}R_{3,t,\\lambda}^w [\\chi_0(t) \\langle x_2-tx_1\\rangle^{-2\\sigma}]^w,$$\nbelongs to the class $S(1,dt^2+d\\tau^2+dX^2)$ uniformly with respect to the parameter $0<\\lambda \\leq \\lambda_0$. We may then deduce from the $L^2$ continuity theorem in the class $S_{00}^0$ that for all $0<\\lambda \\leq \\lambda_0$ and $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n$$\\|R_{3,t,\\lambda}^wu\\|_{L^2(\\mathbb{R}^{2n+1})} =\\|\\lambda^{1-\\sigma}\\tilde{a}_{\\lambda}^w\\langle x_2-tx_1\\rangle^{2\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\lesssim \n\\lambda^{1-\\sigma} \\|\\langle x_2-tx_1\\rangle^{2\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})}.$$\nThis ends the proof of Lemma~\\ref{lem4}.~$\\Box$\n\n\\bigskip\n\n\\noindent\nBy coming back to (\\ref{eq20}), Lemma~\\ref{prop1} is then a direct consequence of Lemmas~\\ref{lem2}, \\ref{lem3} and~\\ref{lem4}.~$\\Box$\n\n\\bigskip\n\nNotice that the power $1-\\sigma$ appearing in the right-hand-side of the estimate given by Lemma~\\ref{prop1} is positive by assumption, since $0<\\sigma<1$.\nWe deduce from Lemma~\\ref{prop1} and (\\ref{eq19}) that for any fixed $T>0$, one may choose a new positive parameter $0< \\lambda_0 \\leq 1$ indexing the Wick quantization sufficiently small so that the term \n$$\\lambda^{1-\\sigma} \\|\\langle x_2-tx_1\\rangle^{2\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})},$$\nappearing in the right-hand-side of the estimate given by Lemma~\\ref{prop1}\ncan be controlled by one half times the left-hand-side terms appearing in the a priori estimate (\\ref{eq19}). For any fixed $T>0$, \nthere therefore exists a positive constant $c_T>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\nc_T^{-1}\\big\\|(1+|x_2-tx_1|^{2\\sigma}+|x_1|^{\\frac{2\\sigma}{2\\sigma+1}})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq \\|\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+}u\\|_{L^2(\\mathbb{R}^{2n+1})}\\\\\n+\\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^{w}u\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline*}\nBy noticing that, for any $\\varepsilon>0$, there exists a positive constant $C_{\\varepsilon}>0$ such that for all $(t,x_1,x_2) \\in [-T,T] \\times \\mathbb{R}^{2n}$,\n$$\\langle x_2-tx_1\\rangle^{(2\\sigma-1)_+} \\leq \\varepsilon \\langle x_2-tx_1\\rangle^{2\\sigma}+C_{\\varepsilon},$$ \nwe finally obtain that for any fixed $T>0$, there exists a positive constant $c_T>0$ such that for all $u \\in \\mathscr{S}(\\mathbb{R}_{t,x_1,x_2}^{2n+1})$ satisfying \n$$\\textrm{supp } u(\\cdot,x_1,x_2) \\subset [-T,T], \\ (x_1,x_2) \\in \\mathbb{R}^{2n},$$\nwe have\n\\begin{multline*}\nc_T^{-1}\\big\\|(1+|x_2-tx_1|^{2\\sigma}+|x_1|^{\\frac{2\\sigma}{2\\sigma+1}})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\\\ \n\\leq \\|iD_tu+[a(t,\\xi_2,\\xi_1+t\\xi_2)F(x_2-tx_1)]^wu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline*}\nThis proves the a priori estimate (\\ref{eq5}) and ends the proof of Proposition~\\ref{th1}.~$\\Box$\n\n\n\n\\section{Proof of Theorem~\\ref{TTH1}}\\label{section3}\n\n\\noindent\nThe next step in the proof of Theorem~\\ref{TTH1} is given by establishing the following hypoelliptic estimate:\n\n\\bigskip\n\n\\begin{proposition}\\label{cor1}\nLet $P$ be the operator defined in \\emph{(\\ref{yo2})} and $K$ a compact subset of~$\\mathbb{R}^{2n+1}$. Then, there exists a positive constant $C_K>0$ such that for any $u \\in C_0^{\\infty}(K)$,\n$$\\big\\|(1+|D_t|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_x|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_{v}|^{2\\sigma})u\\big\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C_K \\big(\\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}\\big).$$\n\\end{proposition}\n\n\\bigskip\n\n\n\\noindent\n\\textit{Proof of Proposition~\\ref{cor1}}. \nLet $K$ be a compact subset of $\\mathbb{R}^{2n+1}$ and denote by $\\mathcal{F}_{t,x}$ the partial Fourier transform with respect to the $t,x$ variables. Then, for any $u \\in C_0^{\\infty}(K)$, we may write that\n$$\\||D_t|^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\lesssim \\||\\tau+v\\cdot \\xi|^{\\frac{2\\sigma}{2\\sigma+1}}\\mathcal{F}_{t,x}u\\|_{L^2(\\mathbb{R}^{2n+1})}+\\||v\\cdot \\xi|^{\\frac{2\\sigma}{2\\sigma+1}}\\mathcal{F}_{t,x}u\\|_{L^2(\\mathbb{R}^{2n+1})}.$$\nNotice that there exists a positive constant $C_{K}>0$ such that for all $u \\in C_0^{\\infty}(K)$,\n$$\\||v\\cdot \\xi|^{\\frac{2\\sigma}{2\\sigma+1}}\\mathcal{F}_{t,x}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\leq C_{K} \\||D_x|^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}^{2n+1})}$$\nand that \n\\begin{multline*}\n\\||\\tau+v\\cdot \\xi|^{\\frac{2\\sigma}{2\\sigma+1}}\\mathcal{F}_{t,x}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\lesssim \\|(\\tau+v\\cdot \\xi)\\mathcal{F}_{t,x}u\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}\\\\\n\\lesssim \\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline*}\nRecalling that $a \\in L^{\\infty}(\\mathbb{R}^{2n+1})$, we obtain that for all $u \\in C_0^{\\infty}(K)$,\n\\begin{multline*}\n\\||D_t|^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}^{2n+1})} \\lesssim \\|Pu\\|_{L^2(\\mathbb{R}^{2n+1})}+ \\||D_x|^{\\frac{2\\sigma}{2\\sigma+1}}u\\|_{L^2(\\mathbb{R}^{2n+1})}\\\\ +\\|(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2(\\mathbb{R}^{2n+1})}+\\|u\\|_{L^2(\\mathbb{R}^{2n+1})}.\n\\end{multline*}\nProposition~\\ref{cor1} is then a direct consequence of Proposition~\\ref{th1}.~$\\Box$\n\n\n\n\n\\bigskip\n\nBy using Proposition~\\ref{cor1}, one can now complete the proof of Theorem~\\ref{TTH1}. Let $K$ be a compact subset of $\\mathbb{R}^{2n+1}$, $s \\in \\mathbb{R}$ and $\\chi$ be a $C_0^{\\infty}(\\mathbb{R}^{2n+1})$ function satisfying $\\chi=1$ on $K$. Setting\n$$\\Lambda=(1+|D_t|^2+|D_x|^2+|D_v|^2)^{\\frac{1}{2}},$$\nwe apply Proposition~\\ref{cor1} to the function $\\chi \\Lambda^su$ with $u \\in C_0^{\\infty}(K)$,\n$$\\big\\|(1+\\langle D_t\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_x\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_{v}\\rangle ^{2\\sigma})\\chi \\Lambda^su\\big\\|_{L^2} \\leq C_K \\big(\\|P\\chi \\Lambda^su\\|_{L^2}+\\|\\chi \\Lambda^su\\|_{L^2}\\big).$$\nNotice that \n$$\\chi \\Lambda^su= \\Lambda^s(\\chi u)+[\\chi,\\Lambda^s]u=\\Lambda^su+[\\chi,\\Lambda^s]u,$$\nsince $\\chi=1$ on $K$ and $u \\in C_0^{\\infty}(K)$. \nWe have \n$$\\|\\chi \\Lambda^su\\|_{L^2} \\leq \\|u\\|_{s}.$$\nSymbolic calculus shows that the Weyl symbol of the operator $[\\chi,\\Lambda^s]$ belongs to the class\n$S(\\lambda^{s-1},\\tilde{\\Gamma})$ with\n$$\\lambda=(1+|\\tau|^2+|\\xi|^2+|\\eta|^2)^{\\frac{1}{2}} \\textrm{ and } \\tilde{\\Gamma}=dt^2+dx^2+dy^2+\\lambda^{-2}(d\\tau^2+d\\xi^2+d\\eta^2).$$\nIt therefore follows that \n\\begin{multline*}\n\\big|\\big\\|(1+\\langle D_t \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_x\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_{v}\\rangle^{2\\sigma})\\chi \\Lambda^su\\big\\|_{L^2} \n\\\\ -\\big\\|(1+\\langle D_t \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_x\\rangle^{\\frac{2\\sigma}{2\\sigma+1}}+\\langle D_{v}\\rangle^{2\\sigma})u\\big\\|_{s}\\big| \n\\lesssim \\|\\langle D_{v}\\rangle ^{(2\\sigma-1)_+}u\\|_{s},\n\\end{multline*}\nwith $(2\\sigma-1)_+=\\textrm{max}(2\\sigma-1,0)$.\nOn the other hand, we have \n\\begin{align*}\n\\|P\\chi \\Lambda^su\\|_{L^2} \\leq & \\ \\|P \\Lambda^su\\|_{L^2}+\\|P[\\chi,\\Lambda^s]u\\|_{L^2}\\\\ \\leq & \\ \\|Pu\\|_{s}+ \\|[P,\\Lambda^s]u\\|_{L^2} +\\|[P,[\\chi,\\Lambda^s]]u\\|_{L^2}+\\|[\\chi,\\Lambda^s]Pu\\|_{L^2}.\n\\end{align*}\nNotice that \n$$[P,\\Lambda^s]=[v,\\Lambda^s] \\cdot \\partial_x+[a(t,x,v),\\Lambda^s](-\\tilde{\\Delta}_v)^{\\sigma}.$$\nSymbolic calculus directly shows that \n$$ \\|[P,\\Lambda^s]u\\|_{L^2} \\lesssim \\|\\langle D_{v}\\rangle ^{(2\\sigma-1)_+}u\\|_{s}$$\nand \n$$\\|[\\chi,\\Lambda^s]Pu\\|_{L^2}=\\|[\\chi,\\Lambda^s]\\Lambda^{-s}\\Lambda^sPu\\|_{L^2} \\lesssim \\|Pu\\|_{s}.$$\nIt follows that \n\\begin{align*}\n& \\ \\big\\|(1+|D_t|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_x|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_{v}|^{2\\sigma})u\\big\\|_{s} \\\\ \\lesssim & \\\n \\|Pu\\|_{s}+\\|[P,[\\chi,\\Lambda^s]]u\\|_{L^2}\n+\\|\\langle D_{v}\\rangle ^{(2\\sigma-1)_+}u\\|_{s}.\n\\end{align*}\nIt remains to study the double commutator \n\\begin{multline*}\n[P,[\\chi,\\Lambda^s]]=[\\partial_t,[\\chi,\\Lambda^s]]+[v \\cdot \\partial_x,[\\chi,\\Lambda^s]]+[a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma},[\\chi,\\Lambda^s]]\n=[\\partial_t,[\\chi,\\Lambda^s]]\\\\ +v[\\partial_x,[\\chi,\\Lambda^s]]+[v,[\\chi,\\Lambda^s]]\\partial_x+a(t,x,v)[(-\\tilde{\\Delta}_v)^{\\sigma},[\\chi,\\Lambda^s]]+[a(t,x,v),[\\chi,\\Lambda^s]](-\\tilde{\\Delta}_v)^{\\sigma}.\n\\end{multline*}\nBy using the fact that $a$ is a $C_b^{\\infty}(\\mathbb{R}^{2n+1})$ function and $0<\\sigma <1$, symbolic calculus directly shows that \n\\begin{multline*}\n\\|[\\partial_t,[\\chi,\\Lambda^s]]u\\|_{L^2}+\\|[v,[\\chi,\\Lambda^s]]\\partial_xu\\|_{L^2}+\\|a(t,x,v)[(-\\tilde{\\Delta}_v)^{\\sigma},[\\chi,\\Lambda^s]]u\\|_{L^2}\\\\ +\\|[a(t,x,v),[\\chi,\\Lambda^s]](-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2} \\lesssim \\|u\\|_{s},\n\\end{multline*}\nsince we recall that the Weyl symbol of the operator $[\\chi,\\Lambda^s]$ belongs to the class $S(\\lambda^{s-1},\\tilde{\\Gamma})$. It follows that \n\\begin{multline*}\n\\|[P,[\\chi,\\Lambda^s]]u\\|_{L^2} \\lesssim \\|[v,[\\partial_x,[\\chi,\\Lambda^s]]]u\\|_{L^2}+\\|[\\partial_x,[\\chi,\\Lambda^s]]]vu\\|_{L^2}+\\|u\\|_{s}\\\\\n\\lesssim \\|vu\\|_{s}+\\|u\\|_{s}\\lesssim \\|\\chi vu\\|_{s}+\\|u\\|_{s} \\lesssim \\|u\\|_{s},\n\\end{multline*}\nsince $\\chi=1$ on $K$ and $u \\in C_0^{\\infty}(K)$.\nBy using that for any $\\varepsilon>0$, there exists a positive constant $C_{\\varepsilon}>0$ such that\n$$\\|\\langle D_{v}\\rangle ^{(2\\sigma-1)_+}u\\|_{s} \\leq \\varepsilon \\||D_{v}|^{2\\sigma}u\\|_{s}+C_{\\varepsilon}\\|u\\|_{s},$$\nwe finally obtain that there exists a positive constant $C_K>0$ such that for any $u \\in C_0^{\\infty}(K)$,\n$$\\big\\|(1+|D_t|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_x|^{\\frac{2\\sigma}{2\\sigma+1}}+|D_{v}|^{2\\sigma})u\\big\\|_{s} \\leq C_K \\big(\\|Pu\\|_{s}+\\|u\\|_{s}\\big).$$\nThis ends the proof of Theorem~\\ref{TTH1}.~$\\Box$\n\n\n\\section{Proof of Corollary~\\ref{ev1}}\\label{section4}\n\n\nThis section is devoted to the proof of Corollary~\\ref{ev1}. Let $P$ be the operator defined in (\\ref{yo2}) and $N_0 \\in \\mathbb{N}$. For simplicity, we shall assume that $(t_0,x_0) =(0,0)$ and consider a function $u \\in H^{-N_0}(\\mathbb{R}^{2n+1})$ satisfying \n\\begin{equation}\\label{sev20}\nPu \\in H^s_{\\textrm{loc}, (t_0,x_0)}(\\mathbb{R}^{2n+1}_{t,x,v}),\n\\end{equation}\nwith $s \\geq 0$. Let $\\varphi$ be a $C_0^\\infty(\\mathbb{R}_{t,x}^{n+1})$ function supported in the set \n$$\\{(t,x) \\in \\mathbb{R}^{n+1} : |(t,x)|<2\\}$$\nsatisfying $\\varphi=1$ in a neighborhood of $(t_0,x_0) =(0,0)$ and \n\\begin{equation}\\label{ts9}\n\\varphi(t,x) Pu \\in H^s(\\mathbb{R}_{t,x,v}^{2n+1}).\n\\end{equation}\nFor convenience only, we shall assume that \n\\begin{equation}\\label{ts8}\n \\varphi(t,x)=1,\n\\end{equation}\nfor all $|(t,x)| \\leq 3\/2$.\nWe consider $\\chi$ a $C^\\infty(\\mathbb{R}^{n+1}_{t,x})$ function satisfying $-N_0-2 \\le \\chi \\le s$, \n$\\chi = s$ on $\\{|(t,x)| \\leq 1\\}$, $\\chi = -N_0-2$ when $|(t,x)| \\ge 3\/2$, and define\n\\begin{equation}\\label{sev6}\nM_\\delta(t,x, \\zeta) =\\frac{\\langle \\zeta \\rangle^{\\chi(t,x)}}{\n(1+ \\delta \\langle \\zeta \\rangle)^{N_0+s+2}}, \\ \\zeta=(\\tau,\\xi,\\eta) \\in \\mathbb{R}^{2n+1},\n\\end{equation}\nwith $\\langle \\zeta \\rangle = (1+|\\zeta|^2)^{1\/2}$ and $0 <\\delta \\leq 1$. We recall that $\\tau$, $\\xi$ and $\\eta$ stand respectively for the dual variables of $t$, $x$ and $v$. It directly follows from this definition that for any $\\alpha \\in \\mathbb{N}^{n+1}$, $\\beta \\in \\mathbb{N}^{2n+1}$, there exist some positive constants $C_{\\alpha, \\beta} >0$ such that for all $0< \\delta \\leq 1$, $(t,x) \\in \\mathbb{R}^{n+1}$ and $\\zeta \\in \\mathbb{R}^{2n+1}$,\n\\begin{align}\\label{symbol-est}\n\\big| \\partial^\\alpha_{t,x} \\partial_{\\zeta}^\\beta\nM_\\delta(t,x, \\zeta) \\big|\\le C_{\\alpha, \\beta} \n\\big(\\log \\langle \\zeta \\rangle\\big)^{|\\alpha|} \\langle \\zeta \\rangle^{-|\\beta|}\nM_\\delta(t,x, \\zeta).\n\\end{align}\nNotice that \n$$M_{\\delta}(t,x,\\zeta) \\leq \\langle \\zeta \\rangle^s.$$\nLet $\\varepsilon>0$ be a positive constant and $k \\geq 0$ a nonnegative integer. The symbol $M_{\\delta}$ therefore belongs to the symbol class $S^{s+\\varepsilon}$ uniformly with respect to the parameter $0<\\delta \\leq 1$. We recall that the notation $S^m$, with $m \\in \\mathbb{R}$, stands for the symbol class\n$$S^{m}=S(\\langle \\zeta \\rangle^m,\\Gamma), \\quad \\Gamma=dz^2+\\frac{d\\zeta^2}{\\langle \\zeta \\rangle^{2}},$$\nwith $z=(t,x,v) \\in \\mathbb{R}^{2n+1}$. The symbol $M_{\\delta}$ also belongs to the symbol class $\\tilde{S}^{s}$,\n\\begin{equation}\\label{tt1}\n\\tilde{S}^{s}=S(\\langle \\zeta \\rangle^s,\\tilde{\\Gamma}), \\quad \\tilde{\\Gamma}=\\langle \\zeta \\rangle dz^2+\\frac{d\\zeta^2}{\\langle \\zeta \\rangle^{2}},\n\\end{equation}\nuniformly with respect to the parameter $0<\\delta \\leq 1$.\nFurthermore, notice that \n$$M_{\\delta}(t,x,\\zeta) \\leq \\delta^{-N_0-s-2} \\langle \\zeta \\rangle^{-N_0-2},$$\nwhen $0<\\delta \\leq 1$. It follows that for each fixed $0<\\delta \\leq 1$, the symbol $M_\\delta$ also belongs to the class\n$S^{-N_0-2+\\varepsilon}$. \nMore specifically, we deduce from (\\ref{sev6}) and the Leibniz formula that for any $\\alpha \\in \\mathbb{N}^{2n+1}$, $\\beta \\in \\mathbb{N}^{2n+1}$, we may write \n\\begin{equation}\\label{symbol-est1}\n\\partial^\\alpha_{z} \\partial_{\\zeta}^\\beta\\big[\\langle v \\rangle^{-k} M_\\delta(t,x, \\zeta)\\big]=a_{\\alpha,\\beta, \\delta}(z,\\zeta)\\langle v \\rangle^{-k}M_\\delta(t,x, \\zeta),\n\\end{equation}\nwith $a_{\\alpha,\\beta, \\delta}$ a symbol belonging to the class $S^{\\varepsilon-|\\beta|}$\nuniformly with respect to the parameter $0<\\delta \\leq 1$. \nWe begin by establishing the following lemma:\n\n\n\n\n\\bigskip\n\n\\begin{lemma}\\label{llop1}\nLet $0<\\varepsilon<1$, $k \\geq 0$, $N \\in \\mathbb{N}$ and $A \\in S^m$, $m \\in \\mathbb{R}$. Then, there exists a symbol $B_{\\delta}$ belonging to the class $S^{m-(1-\\varepsilon)}$ uniformly with respect to the parameter $0<\\delta \\leq 1$ such that \n$$\\emph{\\textrm{Op}}(A\\langle v \\rangle^{-k}M_{\\delta})-A(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)-\\emph{\\textrm{Op}}(B_{\\delta}\\langle v \\rangle^{-k}M_{\\delta}) \\in \\emph{\\textrm{Op}}(S^{-N}),$$\nuniformly with respect to the parameter $0<\\delta \\leq 1$.\n\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\nHere $A(z,D_z)$ or $\\textrm{Op}(A)$ stands for the standard quantization of the symbol $A(z,\\zeta)$,\n$$A(z,D_z)u(z)=\\int_{\\mathbb{R}^{2n+1}}e^{2\\pi i(z-\\tilde{z}).\\zeta}A(z,\\zeta)u(\\tilde{z})d\\zeta d\\tilde{z}, \\ z=(t,x,v) \\in \\mathbb{R}^{2n+1}.$$\n\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{llop1}}. Since $A \\in S^m$ and $M_{\\delta} \\in S^{s+\\varepsilon}$ uniformly with respect to the parameter $0<\\delta \\leq 1$, symbolic calculus shows that $\\langle v \\rangle^{-k}M_{\\delta} \\in S^{s+\\varepsilon}$ uniformly with respect to the parameter $0<\\delta \\leq 1$ and\n$$A(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z) -\\sum_{0\\le|\\alpha|\\leq [m+s+\\varepsilon]+N}\\frac{1}{\\alpha !}\\textrm{Op}\\big((\\partial^\\alpha_\\zeta A) D_z^\\alpha\\big(\\langle v \\rangle^{-k}M_{\\delta}\\big)\\big) \\in \\textrm{Op}(S^{-N}),$$\nwith $[m+s+\\varepsilon]$ being the integer part of $m+s+\\varepsilon \\in \\mathbb{R}$. We may therefore write from (\\ref{symbol-est1}) that \n$$\\textrm{Op}(A\\langle v \\rangle^{-k}M_{\\delta})-A(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)-\\textrm{Op}(B_{\\delta}\\langle v \\rangle^{-k}M_{\\delta}) \\in \\textrm{Op}(S^{-N}),$$\nwith \n$$B_{\\delta}(z,\\zeta)=-\\sum_{1\\le|\\alpha|\\leq [m+s+\\varepsilon]+N}\\frac{1}{\\alpha !}(\\partial^\\alpha_\\zeta A)(z,\\zeta)a_{\\alpha,0,\\delta}(z,\\zeta).$$\nSince $\\partial^\\alpha_\\zeta A \\in S^{m-|\\alpha|}$ and $a_{\\alpha,0,\\delta} \\in S^{\\varepsilon}$ uniformly with respect to $0<\\delta \\leq 1$, the symbol \n$B_{\\delta}$ belongs to the class $S^{m-(1-\\varepsilon)}$ uniformly with respect to $0<\\delta \\leq 1$. This proves Lemma~\\ref{llop1}.~$\\Box$\n\n\n\\bigskip\n\\noindent\n\n\n\\bigskip\n\n\\begin{lemma}\\label{llop2}\nLet $k \\geq 0$, $N \\in \\mathbb{N}$ and $A \\in S^m$, $m \\in \\mathbb{R}$. Then, there exists a symbol $A_{\\delta}$ belonging to the class $S^{m}$ uniformly with respect to the parameter $0<\\delta \\leq 1$ such that \n$$\\emph{\\textrm{Op}}(A\\langle v \\rangle^{-k}M_{\\delta})-A_{\\delta}(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z) \\in \\emph{\\textrm{Op}}(S^{-N}),$$\nuniformly with respect to the parameter $0<\\delta \\leq 1$.\n\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{llop2}}. Let $0<\\varepsilon<1$. Recalling that the symbol $\\langle v \\rangle^{-k}M_{\\delta}$ belongs to the class $S^{s+\\varepsilon}$ uniformly with respect to the parameter $0<\\delta \\leq 1$, Lemma~\\ref{llop2} follows from $k_0$ iterations of Lemma~\\ref{llop1} with $m+s+\\varepsilon-k_0(1-\\varepsilon)<-N$ on the successive remainder terms $\\textrm{Op}(B_{\\delta}\\langle v \\rangle^{-k}M_{\\delta})$.~$\\Box$\n\n\n\n\n\n\n\n\\bigskip\n\n\\noindent\nWe shall now use the following commutator argument: \n\n\n\n\n\\bigskip\n\n\\begin{lemma}\\label{lop1}\nLet $0<\\varepsilon<1$, $k \\geq 1$ and $u$ the function defined in \\emph{(\\ref{sev20})}. Then, there exists a positive constant $C_{k,\\varepsilon}>0$ such that for all $0<\\delta \\leq 1$, \n$$\\|[\\partial_t + v\\cdot \\nabla_x \\, , \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)] u\\|_{L^2} \\leq C_{k,\\varepsilon} \\| \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{\\varepsilon} +C_{k,\\varepsilon}\\|u\\|_{-N_0}.$$\n\\end{lemma}\n\n\\bigskip\n\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lop1}}. Let $k \\geq 1$ be a positive integer and $0<\\varepsilon<1$. We consider the commutator \n$$[\\partial_t + v\\cdot \\nabla_x \\, , \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)].$$\nSymbolic calculus shows that the symbol of the commutator \n$$ [\\partial_t + v\\cdot \\nabla_x , \\langle v \\rangle^{-k}M_\\delta(t,x,D_z)]=\\langle v \\rangle^{-k} [\\partial_t + v\\cdot \\nabla_x , M_\\delta(t,x,D_z)],$$ \nin the standard quantization is \n$$\\langle v \\rangle^{-k} \\big((\\partial_t \\chi)(t,x) + v\\cdot (\\nabla_x\\chi)(t,x)\\big) M_\\delta(t,x, \\zeta)\\log \\langle \\zeta \\rangle-\\langle v \\rangle^{-k} \\xi \\cdot (\\nabla_{\\eta}M_{\\delta})(t,x,\\zeta).$$\nThis symbol may be written as \n\\begin{multline}\\label{sev21b}\n\\log \\langle \\zeta \\rangle\\Big(\\langle v \\rangle^{-1}(\\partial_t \\chi)(t,x) + \\frac{v}{\\langle v \\rangle}\\cdot (\\nabla_x\\chi)(t,x)\\Big) \\langle v \\rangle^{-(k-1)}M_\\delta(t,x, \\zeta)\\\\ - \\langle v \\rangle^{-1}\\xi \\cdot (\\nabla_{\\eta})\\big[\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,\\zeta)\\big].\n\\end{multline} \nBy using that $\\log \\langle \\zeta \\rangle \\in S^{\\varepsilon}$, it follows from (\\ref{symbol-est1}) and (\\ref{sev21b}) that this symbol may also be written as\n \\begin{equation}\\label{sev21}\nA_{\\delta}(z,\\zeta)\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,\\zeta),\n\\end{equation} \nwith $A_{\\delta}$ belonging to the class $S^{\\varepsilon}$ uniformly with respect to $0<\\delta \\leq 1$. Then, we deduce from Lemma~\\ref{llop2} that \n\\begin{equation}\\label{sev1}\n[\\partial_t + v\\cdot \\nabla_x ,\\langle v \\rangle^{-k} M_\\delta(t,x,D_z)]-\\tilde{A}_{\\delta}(z,D_z)\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,D_z) \\in \\textrm{Op}(S^{-N_0}),\n\\end{equation}\nwith $\\tilde{A}_{\\delta}$ a new symbol belonging to the class $S^{\\varepsilon}$ uniformly with respect to $0<\\delta \\leq 1$.\nThis implies that \n\\begin{multline}\\label{sev2}\n\\|[\\partial_t + v\\cdot \\nabla_x , \\langle v \\rangle^{-k}M_\\delta(t,x,D_z)]u-\\tilde{A}_{\\delta}(z,D_z)\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,D_z)u\\|_{L^2} \\\\ \\lesssim \\|u\\|_{-N_0}.\n\\end{multline}\nOn the other hand, we have\n\\begin{equation}\\label{rum2}\n\\|\\tilde{A}_{\\delta}(z,D_z)\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,D_z)u\\|_{L^2} \\lesssim \\|\\langle v \\rangle^{-(k-1)}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon},\n\\end{equation}\nsince $\\tilde{A}_{\\delta} \\in S^{\\varepsilon}$ uniformly with respect to $0<\\delta \\leq 1$.\nIt follows from (\\ref{sev2}), (\\ref{rum2}) and the triangular inequality that the estimate \n$$\\|[\\partial_t + v\\cdot \\nabla_x \\, , \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)] u\\|_{L^2} \\lesssim \\| \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{\\varepsilon} +\\|u\\|_{-N_0},$$\nholds uniformly with respect to the parameter $0<\\delta \\leq 1$. This proves Lemma~\\ref{lop1}.~$\\Box$\n\n\n\n\\bigskip\n\n\\noindent\nOne can now estimate from above the following second commutator: \n\n\n\n\n\\bigskip\n\n\\begin{lemma}\\label{lop2}\nLet $0<\\varepsilon<1$, $k \\geq 0$ and $u$ be the function defined in \\emph{(\\ref{sev20})}. Then, there exists a positive constant $C_{k,\\varepsilon}>0$ such that for all $0<\\delta \\leq 1$, \n\\begin{multline*}\n\\|[a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)] u\\|_{L^2}\\\\ \\leq C_{k,\\varepsilon} \n \\| \\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon} \\langle v \\rangle^{-k} \nM_\\delta(t,x,D_z) u \\|_{L^2}+C_{k,\\varepsilon}\\|u\\|_{-N_0},\n\\end{multline*}\nwith $(2\\sigma-1)_+=\\mbox{\\rm max}(2\\sigma-1,0)$.\n\\end{lemma}\n\n\\bigskip\n\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lop2}}. Notice that the symbol of the operator $a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}$ in the standard quantization does not belong to the class $S^{2\\sigma}$ in $\\mathbb{R}_{z,\\zeta}^{4n+2}$. Handling this commutator term \n$$[a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)]$$\nneeds therefore some caution. By using that \n$$[a(t,x,v),\\langle v \\rangle^{-k}]=0 \\textrm{ and } [(-\\tilde{\\Delta}_v)^{\\sigma},M_\\delta(t,x,D_z)]=0,$$ \nwe may write that \n\\begin{multline}\\label{ryu1}\n[a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)]=a(t,x,v)[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}]M_\\delta(t,x,D_z)\\\\\n+ \\langle v \\rangle^{-k} [a(t,x,v), M_\\delta(t,x,D_z)](-\\tilde{\\Delta}_v)^{\\sigma}.\n\\end{multline}\nLet $0<\\varepsilon<1$. Regarding the first term in the right-hand-side of (\\ref{ryu1}), we first notice that the symbol of the operator \n$[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}]\\langle v \\rangle^{k} \\langle D_v \\rangle^{-(2\\sigma-1)_+-\\varepsilon}$ \nbelongs to the class\n$$S(1,dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2).$$\nIndeed, the symbol of the operator $(-\\tilde{\\Delta}_v)^{\\sigma}$ belongs to $S(\\langle \\eta \\rangle^{2\\sigma},dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2)$, the symbol of the operator \n$\\langle D_v \\rangle^{-(2\\sigma-1)_+-\\varepsilon}$ belongs to $S(\\langle \\eta \\rangle^{-(2\\sigma-1)_+-\\varepsilon},dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2)$\n and\n$\\langle v \\rangle^{m} \\in S(\\langle v \\rangle^{m},dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2)$ when $m \\in \\mathbb{R}$. Symbolic calculus shows that the symbol of the commutator \n$[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}]$ belongs to $S(\\langle \\eta \\rangle^{2\\sigma-1}\\langle v \\rangle^{-k},dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2)$ and therefore that the symbol of the operator $[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}]\\langle v \\rangle^{k}\\langle D_v \\rangle^{-(2\\sigma-1)_+-\\varepsilon}$ \nbelongs to the class\n$$S(\\langle \\eta \\rangle^{2\\sigma-1-(2\\sigma-1)_+-\\varepsilon},dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2) \\subset S(1,dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2).$$\nRecalling that by assumption $a \\in C_b^{\\infty}(\\mathbb{R}_{t,x,v}^{2n+1})$, we obtain that \n\\begin{align}\\label{ryu2}\n& \\ \\|a(t,x,v)[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}]M_\\delta(t,x,D_z)u\\|_{L^2} \\\\ \\notag\n\\lesssim & \\ \\|[(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k}] \\langle v \\rangle^{k}\\langle D_v \\rangle^{-(2\\sigma-1)_+-\\varepsilon}\\langle D_v \\rangle^{(2\\sigma-1)_++\\varepsilon} \\langle v \\rangle^{-k}M_\\delta(t,x,D_z)u\\|_{L^2} \\\\ \\notag\n\\lesssim & \\ \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon} \\langle v \\rangle^{-k}M_\\delta(t,x,D_z)u\\|_{L^2}.\n\\end{align}\nRecalling that $M_{\\delta} \\in S^{s+\\varepsilon}$ uniformly with respect to $0<\\delta \\leq 1$ and that by assumption $a \\in S^0$, symbolic calculus shows that the symbol of the operator \n$$\\langle v \\rangle^{-k}[a(t,x,v), M_\\delta(t,x,D_z)],$$ \nwrites as\n\\begin{align}\\label{ryu3}\n& \\ -\\langle v \\rangle^{-k} \\sum_{1\\le|\\alpha|\\leq [s+\\varepsilon]+N_0+2}\\frac{1}{\\alpha !}(\\partial^\\alpha_\\zeta M_{\\delta}) D_z^\\alpha a\\\\ \n= & \\ -\\sum_{1\\le|\\alpha|\\leq [s+\\varepsilon]+N_0+2}\\frac{1}{\\alpha !}(D_z^\\alpha a)\\partial^\\alpha_\\zeta\\big[\\langle v \\rangle^{-k} M_{\\delta}\\big],\\notag\n\\end{align}\nup to a term in $S^{-N_0-2}$. This latter term in the class $S^{-N_0-2}$ gives rise to a term bounded by \n$$\\|(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{-N_0-2} \\lesssim \\|u\\|_{-N_0},$$\nbecause $0<\\sigma <1$,\nwhile estimating from above the term \n$$\\|\\langle v \\rangle^{-k} [a(t,x,v), M_\\delta(t,x,D_z)](-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2}.$$\nBy coming back to (\\ref{ryu3}), we deduce from (\\ref{symbol-est1}) that this symbol reduces to a term $A_{\\delta}(z,\\zeta)\\langle v \\rangle^{-k} M_{\\delta}$ with \n$A_{\\delta}$ belonging to $S^{\\varepsilon-1}$ uniformly with respect to $0<\\delta \\leq 1$. By using Lemma~\\ref{llop2}, we notice that \n$$\\textrm{Op}(A_{\\delta}\\langle v \\rangle^{-k}M_{\\delta})-\\tilde{A}_{\\delta}(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z) \\in \\textrm{Op}(S^{-N_0-2}),$$\nwith $\\tilde{A}_{\\delta}$ a new symbol belonging to the class $S^{\\varepsilon-1}$ uniformly with respect to $0<\\delta \\leq 1$.\nIt follows that \n\\begin{multline}\\label{ryu4}\n\\|\\textrm{Op}(A_{\\delta}\\langle v \\rangle^{-k}M_{\\delta})(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2} \\lesssim \\|(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{-N_0-2}\\\\ + \\|\\tilde{A}_{\\delta}(z,D_z)\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2}. \n\\end{multline}\nSince $[M_{\\delta}(t,x,D_z),(-\\tilde{\\Delta}_v)^{\\sigma}]=0$ and $\\tilde{A}_{\\delta} \\in S^{\\varepsilon-1}$ uniformly with respect to $0<\\delta \\leq 1$, we deduce from (\\ref{ryu4}) that\n\\begin{multline}\\label{ryu5}\n\\|\\textrm{Op}(A_{\\delta}\\langle v \\rangle^{-k}M_{\\delta})(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2} \\lesssim \\|u\\|_{-N_0} + \\|\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)(-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{\\varepsilon-1}\\\\\n\\lesssim \\|u\\|_{-N_0} + \\|\\langle v \\rangle^{-k}(-\\tilde{\\Delta}_v)^{\\sigma}\\langle v \\rangle^{k}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1}. \n\\end{multline}\nNotice that \n\\begin{multline}\\label{ryu6}\n\\|\\langle v \\rangle^{-k}(-\\tilde{\\Delta}_v)^{\\sigma}\\langle v \\rangle^{k}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1} \n\\leq \\|(-\\tilde{\\Delta}_v)^{\\sigma}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1}\\\\\n+\\|[\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1}.\n\\end{multline}\nWe may write\n\\begin{align}\\label{ryu111}\n& \\ \\|(-\\tilde{\\Delta}_v)^{\\sigma}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1} \\\\ \n\\lesssim & \\ \\|\\langle D_{z}\\rangle^{-(1-\\varepsilon)}(-\\tilde{\\Delta}_v)^{\\sigma}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon}\\langle D_v\\rangle^{2\\sigma-1+\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2} \\notag \\\\\n\\lesssim & \\ \\|\\langle D_v\\rangle^{2\\sigma-1+\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}\\lesssim \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}, \\notag\n\\end{align}\nwith $\\langle D_z \\rangle=(1+D_t^2+|D_x|^2+|D_v|^2)^{1\/2}$, since the operator \n$$\\langle D_{z}\\rangle^{-(1-\\varepsilon)}(-\\tilde{\\Delta}_v)^{\\sigma}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon},$$ \nis bounded on $L^2$.\nNotice also that\n\\begin{align}\\label{ryu112}\n& \\ \\|[\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{\\varepsilon-1} \\\\\n\\lesssim & \\ \\notag \\|\\langle D_{z}\\rangle^{-(1-\\varepsilon)}[\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}\\\\\n\\lesssim & \\ \\notag \\|\\langle D_{z}\\rangle^{-(1-\\varepsilon)}[\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon}\\langle D_v\\rangle^{2\\sigma-1+\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2} \\\\\n\\lesssim & \\ \\|\\langle D_v\\rangle^{2\\sigma-1+\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}\\lesssim \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon}\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}. \\notag\n\\end{align}\nIndeed, we just need to check that the operator \n\\begin{multline*}\n\\langle D_{z}\\rangle^{-(1-\\varepsilon)}[\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon}\\\\\n=\\langle D_{z}\\rangle^{-(1-\\varepsilon)}\\langle D_v\\rangle^{1-\\varepsilon} \\langle D_v\\rangle^{-1+\\varepsilon} [\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon},\n\\end{multline*}\nis bounded on $L^2$. This is actually the case since the operator \n$$\\langle D_{z}\\rangle^{-(1-\\varepsilon)}\\langle D_v\\rangle^{1-\\varepsilon},$$\nis obviously bounded on $L^2$ since $0<1-\\varepsilon<1$, and that symbolic calculus easily shows that the symbol of the operator \n$$\\langle D_v\\rangle^{-1+\\varepsilon} [\\langle v \\rangle^{-k},(-\\tilde{\\Delta}_v)^{\\sigma}]\\langle v \\rangle^{k}\\langle D_v\\rangle^{-2\\sigma+1-\\varepsilon},$$\nbelongs to the class $S(1,dv^2+\\langle \\eta \\rangle^{-2}d\\eta^2)$. It follows from (\\ref{ryu5}), (\\ref{ryu6}), (\\ref{ryu111}) and (\\ref{ryu112}) that \n\\begin{align}\\label{ryu7}\n& \\ \\|\\langle v \\rangle^{-k} [a(t,x,v), M_\\delta(t,x,D_z)](-\\tilde{\\Delta}_v)^{\\sigma}u\\|_{L^2} \\\\ \\lesssim & \\ \n\\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon} \\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)u\\|_{L^2}+ \\|u\\|_{-N_0}. \\notag\n\\end{align}\nTogether with (\\ref{ryu1}) and (\\ref{ryu2}), we finally obtain the estimate\n\\begin{align*}\n&\\|[a(t,x,v)(-\\tilde{\\Delta}_v)^{\\sigma}, \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)] u\\|_{L^2} \\\\\n&\\lesssim \\| \\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon} \\langle v \\rangle^{-k} \nM_\\delta(t,x,D_z) u \\|_{L^2} +\\|u\\|_{-N_0},\n\\end{align*}\nwhich proves Lemma~\\ref{lop2}.~$\\Box$\n\n\\bigskip\n\n\\noindent\nSumming up the two previous lemmas provides the following estimate:\n\n\n\\bigskip\n\n\\begin{lemma}\\label{lop3}\nLet $0<\\varepsilon<1$, $k \\geq 1$, $P$ be the operator defined in \\emph{(\\ref{yo2})} and $u$ the function defined in \\emph{(\\ref{sev20})}. Then, there exists a positive constant $C_{k,\\varepsilon}>0$ such that for all $0<\\delta \\leq 1$, \n\\begin{multline*}\n\\|[P, \\langle v \\rangle^{-k} M_\\delta(t,x,D_z)] u\\|_{L^2} \\leq C_{k,\\varepsilon} \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon}\\langle v \\rangle^{-k} \nM_\\delta(t,x,D_z) u \\|_{L^2} \\\\ +C_{k,\\varepsilon} \\| \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{\\varepsilon} +C_{k,\\varepsilon}\\|u\\|_{-N_0},\n\\end{multline*}\nwith $(2\\sigma-1)_+=\\mbox{\\rm max}(2\\sigma-1,0)$.\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\nResuming our proof of Corollary~\\ref{ev1}, we may first deduce from Proposition~\\ref{th1} that for all $w \\in \\mathscr{S}(\\mathbb{R}_{t,x,v}^{2n+1})$ satisfying\n\\begin{equation}\\label{rie1}\n\\textrm{supp } w(\\cdot,x,v) \\subset [-3,3], \\ (x,v) \\in \\mathbb{R}^{2n},\n\\end{equation}\nwe have\n\\begin{equation}\\label{rie2}\n\\|\\langle D_x \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}w\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}w\\|_{L^2} \\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2}.\n\\end{equation}\nThese estimates are maximal hypoelliptic estimates and we therefore get an upper bound for the transport term\n\\begin{multline}\\label{rie3}\n\\|(\\partial_t+ v\\cdot \\nabla_x)w\\|_{L^2} \\lesssim \\|P w \\|_{L^2}+\\|a(t,x,v)(-\\tilde{\\Delta}_v^\\sigma)w\\|_{L^2}\\\\ \\lesssim \\|P w \\|_{L^2}+\\|\\langle D_v\\rangle^{2\\sigma}w\\|_{L^2}\n\\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2},\n\\end{multline}\nsince $a \\in C_b^{\\infty}(\\mathbb{R}^{2n+1})$. Notice from (\\ref{rie2}) and (\\ref{rie3}) that for those functions $w$, \n\\begin{align}\\label{rie4}\n& \\ \\|\\langle v \\rangle^{-1}\\Lambda_{t,x}^{-\\frac{1}{2\\sigma +1}}D_t w\\|_{L^2}\\\\\n\\lesssim & \\ \\|\\langle v \\rangle^{-1}\\Lambda_{t,x}^{-\\frac{1}{2\\sigma +1}}(\\partial_t+ v\\cdot \\nabla_x) w\\|_{L^2}+ \\|\\langle v \\rangle^{-1}\\Lambda_{t,x}^{-\\frac{1}{2\\sigma +1}} v\\cdot \\nabla_x w\\|_{L^2} \\notag \\\\ \\notag\n\\lesssim & \\ \\|(\\partial_t+ v\\cdot \\nabla_x) w \\|_{L^2} + \\|\\langle D_x\\rangle^{\\frac{2\\sigma}{2\\sigma+1}} w\\|_{L^2} \\lesssim \n \\|Pw\\|_{L^2}+\\|w\\|_{L^2},\n\\end{align}\nwith $\\Lambda_{t,x}=(1+D_t^2+|D_x|^2)^{1\/2}$. It follows from (\\ref{rie2}) and (\\ref{rie4}) that \n\\begin{multline*}\n\\|\\Lambda_{t,x}^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2}\n\\lesssim \\|\\langle v \\rangle^{-1}\\Lambda_{t,x}^{-\\frac{1}{2\\sigma +1}}D_tw\\|_{L^2}+\\|\\langle v \\rangle^{-1}\\Lambda_{t,x}^{-\\frac{1}{2\\sigma +1}}D_xw\\|_{L^2}\\\\ \\lesssim\n \\|Pw\\|_{L^2}+\\|w\\|_{L^2}+\\|\\langle D_x \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}w\\|_{L^2} \\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2}.\n\\end{multline*}\nAnother use of (\\ref{rie2}) shows that \n\\begin{equation}\\label{rie5}\n\\|\\Lambda_{t,x}^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}w\\|_{L^2} \\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2}.\n\\end{equation}\nNotice furthermore that \n\\begin{multline*}\n\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2} \\lesssim \\|\\Lambda_{t,x}^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2}+\n\\|\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2} \n\\lesssim \\|\\Lambda_{t,x}^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2} \\\\ +\n\\|[\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}},\\langle v \\rangle^{-1}]w\\|_{L^2}+\\|\\langle v \\rangle^{-1}\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}}w\\|_{L^2}\n\\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2},\n\\end{multline*}\nbecause \n$$\\|\\langle v \\rangle^{-1}\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}}w\\|_{L^2} \\lesssim \\|\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}}w\\|_{L^2} \\lesssim \\|\\langle D_v \\rangle^{2\\sigma}w\\|_{L^2} \\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2}$$\nand that the operator $[\\langle D_v \\rangle^{\\frac{2\\sigma}{2\\sigma +1}},\\langle v \\rangle^{-1}]$ is bounded on $L^2$ since symbolic calculus shows that its symbol belongs to the class $S^0$.\nWe therefore obtain from (\\ref{rie5}) the estimate\n\\begin{equation}\\label{rie6}\n\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-1}w\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}w\\|_{L^2} \\lesssim \\|Pw\\|_{L^2}+\\|w\\|_{L^2},\n\\end{equation}\nwhich may be extended to any function $w \\in H^{2-\\varepsilon}(\\mathbb{R}^{2n+1})$ satisfying \\eqref{rie1} and \n$$v \\cdot \\nabla_x w \\in L^2(\\mathbb{R}^{2n+1}),$$ \nwith $\\varepsilon>0$ such that $\\mbox{\\rm max}(2\\sigma,1) \\leq 2-\\varepsilon$. This is possible since $0<\\sigma <1$.\nTake now a $C_0^{\\infty}(\\mathbb{R})$ function $\\psi$ such that \n$$\\textrm{supp }\\psi \\subset [-3,3] \\textrm{ and } \\psi(t) = 1 \\textrm{ if }|t| < 2.$$ \nRecalling that the function $u$ defined in (\\ref{sev20}) belongs to $H^{-N_0}(\\mathbb{R}^{2n+1})$ and that for any $\\varepsilon >0$ such that $\\mbox{\\rm max}(2\\sigma,1) \\leq 2-\\varepsilon$, the symbol $M_{\\delta}$ belongs to $S^{-N_0-2+\\varepsilon}$ for each fixed $0<\\delta \\leq 1$, we notice that\n$$M_{\\delta}(t,x,D_z)u \\in H^{2-\\varepsilon}(\\mathbb{R}^{2n+1}),$$\nfor any $0<\\delta \\leq 1$.\nIt follows that the estimate (\\ref{rie6}) may be applied to function\n\\begin{equation}\\label{ts2}\n\\psi(t)\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u,\n\\end{equation}\nwith an integer $k \\geq 1$. We obtain that \n\\begin{multline}\\label{rie7}\n\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\psi(t)\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}\\psi(t)\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \n\\lesssim \\|P\\psi(t)\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}+\\|\\psi(t)\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}.\n\\end{multline}\nNotice that the choice of function $\\chi$ in (\\ref{sev6}) implies that the symbol \n$$(1-\\psi(t))\\langle v \\rangle^{-l}M_{\\delta}(t,x,\\zeta),$$ \nwith $l \\geq 0$, belongs to the class $S^{-N_0-2}$ uniformly with respect to the parameter $0<\\delta \\leq 1$. Recalling that $0<\\sigma<1$, it follows that \n\\begin{align}\\label{rie8}\n\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}(1-\\psi(t))\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \\lesssim \\|(1-\\psi(t))\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{\\frac{2\\sigma}{2\\sigma +1}} \\lesssim \\|u\\|_{-N_0} \\notag\n\\end{align}\nand\n\\begin{align}\\label{rie9}\n\\|\\langle D_v \\rangle^{2\\sigma}(1-\\psi(t))\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \\lesssim \\|(1-\\psi(t))\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{2\\sigma} \\lesssim \\|u\\|_{-N_0}, \\notag\n\\end{align}\nuniformly with respect to $0<\\delta \\leq 1$. Moreover, since the commutator \n$$[P,\\psi(t)]=\\psi'(t),$$ \nis bounded on $L^2$ and that \n\\begin{align*}\n& \\ \\|P\\psi(t)\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\\n\\leq & \\ \\|\\psi(t)P\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} + \\|[P,\\psi(t)]\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\\n\\lesssim & \\ \\|P\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}+ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2},\n\\end{align*}\nwe deduce from (\\ref{rie7}), (\\ref{rie8}), (\\ref{rie9}) and the triangular inequality that\n\\begin{multline}\\label{rie10}\n\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \n\\lesssim \\|P\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}+\\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}+ \\|u\\|_{-N_0},\n\\end{multline}\nuniformly with respect to $0<\\delta \\leq 1$. Let $0<\\varepsilon_0<1$ such that \n\\begin{equation}\\label{ts1} \n0 < \\varepsilon_0 < \\frac{2\\sigma}{2\\sigma+1} \\textrm{ and } (2\\sigma-1)_++\\varepsilon_0 <2\\sigma.\n\\end{equation}\nWe may use Lemma~\\ref{lop3} to obtain the estimate\n\\begin{align}\n\\label{rie11} & \\ \\|P\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \\notag\n \\leq & \\ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2} + \\|[P,\\langle v \\rangle^{-k} M_\\delta(t,x,D_z)]u\\|_{L^2}\\\\ \\notag\n \\lesssim & \\ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2}+ \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon_0}\\langle v \\rangle^{-k} \nM_\\delta(t,x,D_z) u \\|_{L^2} \\\\ \\notag\n& \\ \\quad + \\| \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{\\varepsilon_0} +\\|u\\|_{-N_0},\n\\end{align}\nwhich holds uniformly with respect to $0<\\delta \\leq 1$. It follows from (\\ref{rie10}) and (\\ref{rie11}) that\n\\begin{align}\\label{rie12}\n& \\ \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2}+\\|\\langle D_v \\rangle^{2\\sigma}\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \n\\lesssim & \\ \\notag \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2}+ \\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon_0}\\langle v \\rangle^{-k} \nM_\\delta(t,x,D_z) u \\|_{L^2} \\\\\n& \\ \\quad \\notag + \\|\\langle D_z \\rangle^{\\varepsilon_0} \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{L^2} \n+ \\|u\\|_{-N_0},\n\\end{align}\nuniformly with respect to $0<\\delta \\leq 1$. Since for any $\\varepsilon>0$, there exists $C_{\\varepsilon}>0$ such that \n\\begin{multline*}\n\\|\\langle D_v\\rangle^{(2\\sigma-1)_++\\varepsilon_0}\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u \\|_{L^2}\\\\ \\leq \\varepsilon \\|\\langle D_v \\rangle^{2\\sigma}\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2}\n+C_\\varepsilon \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2},\n\\end{multline*}\nbecause $(2\\sigma-1)_++\\varepsilon_0 <2\\sigma$, and\n$$ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) u\\|_{L^2} \\lesssim \\|\\langle v \\rangle^{-k+1} M_\\delta(t,x,D_z) u\\|_{L^2} \\lesssim \\|\\langle D_z \\rangle^{\\varepsilon_0}\\langle v \\rangle^{-k+1} M_\\delta(t,x,D_z) u\\|_{L^2},$$\nbecause $\\varepsilon_0>0$, we deduce from (\\ref{rie12}) that \n\\begin{multline}\\label{rie13}\n \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\\n\\lesssim \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2} + \\|\\langle D_z \\rangle^{\\varepsilon_0} \\langle v \\rangle^{-k+1} \nM_\\delta(t,x,D_z) u \\|_{L^2} \n+ \\|u\\|_{-N_0},\n\\end{multline}\nuniformly with respect to $0<\\delta \\leq 1$. We shall need the following instrumental lemma:\n\n\\bigskip\n\n\n\\begin{lemma}\\label{lop4} \nLet $s_10, \\exists C_{\\varepsilon}>0, \\forall w \\in \\mathscr{S}(\\mathbb{R}^{2n+1}), \\ \\|\\langle v \\rangle^{l} w\\|_{s_1} \\leq \\varepsilon \\| w \\|_{s_2} + C_{\\varepsilon} \\|\\langle v \\rangle^m w\\|_{-N}.$$\n\\end{lemma}\n\n\\bigskip\n\n\\noindent\n\\textit{Proof of Lemma~\\ref{lop4}}. Let $m \\in \\mathbb{R}$. Since the symbol of the operator $\\langle D_z \\rangle^{m}$ belongs to the class $S(\\langle \\zeta \\rangle^{m},dz^2+\\langle \\zeta \\rangle^{-2}d\\zeta^2)$ and \n$\\langle v \\rangle^m \\in S(\\langle v \\rangle^{m},dz^2+\\langle \\zeta \\rangle^{-2}d\\zeta^2),$\nsymbolic calculus shows that the symbol of the operator \n$$\\langle D_{z}\\rangle^{-s_2} \\langle v \\rangle^{l}\\langle D_{z}\\rangle^{2s_1 }\\langle v \\rangle^{-l},$$ \nbelongs to the class $S(\\langle \\zeta \\rangle^{2s_1-s_2},dz^2+\\langle \\zeta \\rangle^{-2}d\\zeta^2)$.\nIt follows from the Cauchy-Schwarz inequality that for any $\\varepsilon>0$, there exists $C_{\\varepsilon}>0$ such that \n\\begin{align*}\n \\|\\langle v \\rangle^{l} w\\|_{s_1}^2 = & \\ \\|\\langle D_{z}\\rangle^{s_1} \\langle v \\rangle^{l} w \\|_{L^2}^2 = (\\langle D_{z}\\rangle^{s_2} w,\\langle D_{z}\\rangle^{-s_2} \\langle v \\rangle^{l}\\langle D_{z}\\rangle^{2s_1 }\\langle v \\rangle^{l} w)_{L^2}\\\\\n= & \\ (\\langle D_{z}\\rangle^{s_2} w,\\langle D_{z}\\rangle^{-s_2} \\langle v \\rangle^{l}\\langle D_{z}\\rangle^{2s_1 }\\langle v \\rangle^{-l}\\langle v \\rangle^{2l} w)_{L^2}\\\\\n\\leq & \\ \\|\\langle D_{z}\\rangle^{s_2} w\\|_{L^2}\\|\\langle D_{z}\\rangle^{-s_2} \\langle v \\rangle^{l}\\langle D_{z}\\rangle^{2s_1 }\\langle v \\rangle^{-l}\\langle v \\rangle^{2l} w\\|_{L^2}\\\\\n\\lesssim & \\ \\|w\\|_{s_2}\\|\\langle v \\rangle^{2l} w\\|_{s_1-(s_2-s_1)}\\\\\n\\lesssim & \\ \\varepsilon \\|w\\|_{s_2}^2+C_{\\varepsilon}\\|\\langle v \\rangle^{2l} w\\|_{s_1-(s_2-s_1)}^2.\n\\end{align*}\nNotice that by assumption $s_1-(s_2-s_1)0$,\n\\begin{equation}\\label{ts4}\n \\exists C_{\\varepsilon}>0, \\forall w \\in \\mathscr{S}(\\mathbb{R}^{2n+1}), \\ \\|\\langle v \\rangle^{2} w\\|_{\\varepsilon_0} \\leq \\varepsilon \\| w \\|_{\\frac{2\\sigma}{2\\sigma+1}} + C_{\\varepsilon} \\|\\langle v \\rangle^m w\\|_{-N_0-s-1}.\n\\end{equation}\nWe then choose the integer $k=m-1 \\geq 1$ in (\\ref{ts2}). Recalling that the symbol $M_{\\delta}$ belongs to the class $S^{s+1}$ uniformly with respect to $0<\\delta \\leq 1$, we obtain that \n$$ \\|\\langle v \\rangle^m w\\|_{-N_0-s-1}=\\|M_\\delta(t,x,D_z) u\\|_{-N_0-s-1} \\lesssim \\|u\\|_{-N_0},$$\n uniformly with respect to $0<\\delta \\leq 1$ with $w=\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u.$ \nThe estimate (\\ref{ts4}) extends by density. It follows that \n\\begin{multline}\\label{ts5}\n\\|\\langle D_z \\rangle^{\\varepsilon_0} \\langle v \\rangle^{-k+1} M_\\delta(t,x,D_z) u \\|_{L^2}=\\|\\langle v \\rangle^{2}w \\|_{\\varepsilon_0} \\leq \\varepsilon \\| w \\|_{\\frac{2\\sigma}{2\\sigma+1}}\\\\ + C_{\\varepsilon} \\|\\langle v \\rangle^m w\\|_{-N_0-s-1}\n\\lesssim \\varepsilon \\| \\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u \\|_{L^2} + C_{\\varepsilon} \\|u\\|_{-N_0}.\n\\end{multline} \nWe deduce from (\\ref{rie13}) and (\\ref{ts5}) that \n\\begin{equation}\\label{rie14}\n \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\\n\\lesssim \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2} + \\|u\\|_{-N_0},\n\\end{equation}\nuniformly with respect to $0<\\delta \\leq 1$. Notice from (\\ref{ts8}) and (\\ref{sev6}) that the symbol of the operator \n$$\\langle v \\rangle^{-k}M_{\\delta}(t,x,D_z)(1-\\varphi(t,x))\\langle v \\rangle,$$ \nwith $k \\geq 1$, belongs to the class $S^{-N_0-2}$ uniformly with respect to $0<\\delta \\leq 1$. Recalling from (\\ref{tt1}) that $M_{\\delta} \\in \\tilde{S}^s$ uniformly with respect to $0<\\delta \\leq 1$, we obtain that \n\\begin{align}\\label{ts10}\n& \\ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z) Pu\\|_{L^2} \\\\ \\notag\n\\leq & \\ \\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z)\\varphi Pu\\|_{L^2}+\\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z)(1-\\varphi)Pu\\|_{L^2} \n\\\\ \\notag\n\\leq & \\ \\|M_\\delta(t,x,D_z)\\varphi Pu\\|_{L^2}+\\|\\langle v \\rangle^{-k} M_\\delta(t,x,D_z)(1-\\varphi)\\langle v \\rangle\\langle v \\rangle^{-1}Pu\\|_{L^2}\n\\\\ \\notag\n\\lesssim & \\ \\|\\varphi Pu\\|_{s} +\\|\\langle v \\rangle^{-1}Pu\\|_{-N_0-2}.\n\\end{align}\nSince obviously $\\|\\langle v \\rangle^{-1}Pu\\|_{-N_0-2} \\lesssim \\|u\\|_{-N_0}$, we deduce from (\\ref{rie14}) and (\\ref{ts10}) that for all $0<\\delta \\leq 1$,\n\\begin{equation}\\label{tt2}\n \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \n\\lesssim \\|\\varphi Pu\\|_{s} + \\|u\\|_{-N_0} <+\\infty,\n\\end{equation}\nsince by assumption $u \\in H^{-N_0}(\\mathbb{R}^{2n+1})$ and $\\varphi Pu \\in H^{s}(\\mathbb{R}^{2n+1})$. Let $\\phi$ be $C_0^\\infty(\\mathbb{R}_{t,x}^{n+1})$ function such that $\\textrm{supp }\\phi \\subset \\{|(t,x)|< 1\\}$. Since the commutator $[\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}},\\phi(t,x)]$ is obviously bounded on $L^2$, we deduce from (\\ref{tt2}) that for all $0<\\delta \\leq 1$, \n\\begin{align}\\label{tt3}\n& \\ \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\phi(t,x)\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \\notag\n \\leq & \\ \\|[\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}},\\phi(t,x)]\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\ \\notag\n & \\ \\quad \\quad +\\|\\phi(t,x)\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2}\\\\ \\notag\n \\lesssim & \\ \\|\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2}\n +\\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\\\\\notag\n \\lesssim & \\ \\|\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\langle v \\rangle^{-(k+1)} M_\\delta(t,x,D_z) u\\|_{L^2} \\lesssim \\|\\varphi Pu\\|_{s} + \\|u\\|_{-N_0} <+\\infty.\n\\end{align}\nNotice from (\\ref{sev6}) that \n$$\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}} M_\\delta(t,x,D_z) u=\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}}\\frac{\\langle D_z \\rangle^{s} }{\n(1+ \\delta \\langle D_z \\rangle)^{N_0+s+2}} u$$\nand that \n$$\\langle D_{z}\\rangle^{\\frac{2\\sigma}{2\\sigma +1}}\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}} M_\\delta(t,x,D_z) u \\longrightarrow\n\\langle D_z \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}} \\langle D_z \\rangle^{s}u,$$\nin $\\mathscr{S}'(\\mathbb{R}^{2n+1})$ when $\\delta \\to 0$. Because of the weak compactness of the unit ball in $L^2$, it follows from (\\ref{tt3}) that \n\\begin{equation}\\label{aay10}\n\\langle D_z \\rangle^{\\frac{2\\sigma}{2\\sigma+1}}\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}} \\langle D_z \\rangle^{s}u \\in L^2(\\mathbb{R}^{2n+1}),\n\\end{equation}\nfor any $C_0^\\infty(\\mathbb{R}_{t,x}^{n+1})$ function $\\phi$ such that $\\textrm{supp }\\phi \\subset U_0=\\{|(t,x)|< 1\\}$.\n\n\nTo complete the proof of Corollary~\\ref{ev1}, we notice that the following two assertions \n\\begin{equation}\\label{aay1}\n\\forall \\phi \\in C_0^\\infty(U_0), \\ \\frac{\\phi(t,x)}{\\langle v \\rangle^k} u \\in H^{r+m}(\\mathbb{R}^{2n+1}_{t,x,v})\n\\end{equation}\nand\n\\begin{equation}\\label{aay2}\n\\forall \\phi \\in C_0^\\infty(U_0), \\ \\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\langle D_z \\rangle^{m} u \\in H^r(\\mathbb{R}_{t,x,v}^{2n+1})\n\\end{equation}\nare equivalent if $u \\in H_{-N_0}(\\mathbb{R}^{2n+1}_{t,x,v})$ and $m, r \\in \\mathbb{R}$. Assume first that (\\ref{aay1}) holds. Let $\\phi$ be a $C_0^{\\infty}(U_0)$ function and $\\tilde{\\phi}$ another $C_0^{\\infty}(U_0)$ function such that $\\tilde{\\phi}=1$ on a neighborhood of $\\textrm{supp }\\phi$. Symbolic calculus allows to write\n\\begin{align}\\label{aay3}\n & \\ \\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\langle D_z \\rangle^{m}-\\langle D_z \\rangle^{m} \\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\\\ \\notag\n= & \\ -\\sum_{0 < |\\alpha| < m +N_0+[r]+1} \\frac{1}{\\alpha!} \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big) +R \\\\ \\notag\n= & \\ -\\sum_{0 < |\\alpha| < m +N_0+[r]+1} \\frac{1}{\\alpha!} \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\tilde{\\phi} +\\tilde{R}\n\\end{align}\nwith $R, \\tilde{R} \\in \\textrm{Op}(S^{-N_0-[r]-1})$ and therefore \n\\begin{equation}\\label{aay4}\n\\|\\tilde{R} u\\|_{r} \\lesssim \\|u\\|_{-N_0-[r]-1+r} \\lesssim \\|u\\|_{-N_0} <+\\infty. \n\\end{equation}\nIn order to prove that (\\ref{aay2}) holds, it is sufficient to check that \n\\begin{multline*}\n\\Big\\| \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\tilde{\\phi}u\\Big\\|_r \\\\\n=\\Big\\| \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle v \\rangle^{k}\\langle D_z \\rangle^{-m}\\langle D_z \\rangle^{m}\\frac{\\tilde{\\phi}(t,x)}{\\langle v \\rangle^k}u\\Big\\|_r\n<+\\infty,\n\\end{multline*}\nwhen $0 < |\\alpha| < m +N_0+[r]+1$. This is actually the case since symbolic calculus shows that the symbol of the operator \n$$\\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle v \\rangle^{k}\\langle D_z \\rangle^{-m},$$\nbelongs to $S^0$ when $0 < |\\alpha| < m +N_0+[r]+1$ and that by (\\ref{aay1}),\n$$\\langle D_z \\rangle^{m}\\frac{\\tilde{\\phi}(t,x)}{\\langle v \\rangle^k}u \\in H^r(\\mathbb{R}_{t,x,v}^{2n+1}).$$\nIt follows that (\\ref{aay1}) implies (\\ref{aay2}).\nConversely, we may write for any function $\\phi$ in $C_0^{\\infty}(U_0)$, \n\\begin{align*}\n & \\ \\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\langle D_z \\rangle^{m}-\\langle D_z \\rangle^{m} \\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\\\ \\notag\n= & \\ -\\sum_{0 < |\\alpha| < m +N_0+[r]+1} \\frac{1}{\\alpha!} \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle D_z \\rangle^{-m}\\langle v \\rangle^{k} \\langle v \\rangle^{-k} \\langle D_z \\rangle^{m}+R \\\\ \\notag\n= & \\ -\\sum_{0 < |\\alpha| < m +N_0+[r]+1} \\frac{1}{\\alpha!} \\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle D_z \\rangle^{-m}\\langle v \\rangle^{k}\\frac{\\tilde{\\phi}(t,x)}{\\langle v \\rangle^{k}} \\langle D_z \\rangle^{m} +\\tilde{R}\n\\end{align*}\nwith $R, \\tilde{R} \\in \\textrm{Op}(S^{-N_0-[r]-1})$ if $\\tilde{\\phi}$ is a $C_0^{\\infty}(U_0)$ function satisfying $\\tilde{\\phi}=1$ on a neighborhood of $\\textrm{supp }\\phi$. This implies that \n$$\\|\\tilde{R} u\\|_{r} \\lesssim \\|u\\|_{-N_0-[r]-1+r} \\lesssim \\|u\\|_{-N_0} <+\\infty. $$\nIn order to prove that (\\ref{aay1}) holds, it is sufficient to check that \n$$\\Big\\|\\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle D_z \\rangle^{-m}\\langle v \\rangle^{k}\\frac{\\tilde{\\phi}(t,x)}{\\langle v \\rangle^{k}} \\langle D_z \\rangle^{m}u\\Big\\|_r\n<+\\infty,$$\nwhen $0 < |\\alpha| < m +N_0+[r]+1$. This is actually the case since symbolic calculus shows that the symbol of the operator \n$$\\textrm{Op}\\Big(D_z^{\\alpha}\\left(\\frac{ \\phi(t,x) }{\\langle v \\rangle^k}\\right) \\partial_{\\zeta}^{\\alpha} \\big(\\langle \\zeta \\rangle^{m}\\big)\\Big)\\langle D_z \\rangle^{-m}\\langle v \\rangle^{k},$$\nbelongs to $S^0$ when $0 < |\\alpha| < m +N_0+[r]+1$ and that by (\\ref{aay2}),\n$$\\frac{\\tilde{\\phi}(t,x)}{\\langle v \\rangle^k}\\langle D_z \\rangle^{m}u \\in H^r(\\mathbb{R}_{t,x,v}^{2n+1}).$$\nIt follows that (\\ref{aay2}) implies (\\ref{aay1}).\n\n\nBy finally coming back to (\\ref{aay10}), we deduce that the function defined in (\\ref{sev20}) satisfies \n$$\\frac{\\phi(t,x)}{\\langle v \\rangle^{k+1}} u \\in H^{s+\\frac{2\\sigma}{2\\sigma+1}}(\\mathbb{R}^{2n+1}),$$\nfor any $C_0^\\infty(\\mathbb{R}_{t,x}^{n+1})$ function $\\phi$ such that $\\textrm{supp }\\phi \\subset U_0=\\{|(t,x)|< 1\\}$, that is \n$$\\frac{u}{\\langle v \\rangle^{k+1}} \\in H^{s+\\frac{2\\sigma}{2\\sigma+1}}_{\\textrm{loc}, (t_0,x_0)}(\\mathbb{R}^{2n+1}_{t,x,v}),$$\nwith $(t_0,x_0)=(0,0)$. This ends the proof of Corollary~\\ref{ev1}.~$\\Box$\n\n\n\n\n\n\\section{Appendix}\n\n\\subsection{The Boltzmann equation}\\label{kkboltz}\nThe Boltzmann equation \\cite{17,19} describes the behavior of a dilute gas when the only interactions taken into account are binary collisions. It reads as the evolution equation \n\\begin{equation}\\label{xiaoe1}\n\\begin{cases}\n\\partial_tf+v\\cdot\\nabla_{x}f=Q(f,f)\\\\\nf|_{t=0}=f_0,\n\\end{cases}\n\\end{equation}\nfor the density distribution of the particles $f=f(t,x,v) \\geq 0$ at time $t$, having position $x \\in \\mathbb{R}^d$ and velocity $v \\in \\mathbb{R}^d$.\nThe term appearing in the right-hand-side of this equation $Q(f,f)$ is the so-called quadratic Boltzmann collision operator associated to the Boltzmann bilinear operator \n\\begin{equation}\\label{xiaoeq1}\nQ(g, f)=\\int_{\\mathbb{R}^d}\\int_{S^{d-1}}B(v-v_{*},\\sigma) \\big(g'_* f'-g_*f\\big)d\\sigma dv_*,\n\\end{equation}\nacting on $L^1(\\mathbb{R}^d) \\times L^1(\\mathbb{R}^d)$, $d \\geq 2$, \nwhere $f'_*=f(t,x,v'_*)$, $f'=f(t,x,v')$, $f_*=f(t,x,v_*)$, $f=f(t,x,v)$, \n$$v'=\\frac{v+v_*}{2}+\\frac{|v-v_*|}{2}\\sigma,\\quad v_*'=\\frac{v+v_*}{2}-\\frac{|v-v_*|}{2}\\sigma,$$\nfor $\\sigma$ belonging to the unit sphere $S^{d-1}$. Those relations between pre and post collisional velocities follow from the conservations of momentum and kinetic energy\n$$\\quad v+v_{\\ast}=v'+v_{\\ast}', \\quad |v|^2+|v_{\\ast}|^2=|v'|^2+|v_{\\ast}'|^2,$$\nin the binary collisions.\nThe Boltzmann operator has the fundamental properties of conserving mass, momentum and energy\n$$\\int_{\\mathbb{R}^d}Q(f,f)\\phi(v)dv=0, \\quad \\phi(v)=1,v,|v|^2,$$\nand to satisfy to the so-called Boltzmann's H theorem\n$$-\\frac{d}{dt} \\int_{\\mathbb{R}^d}f \\log f dv=-\\int_{\\mathbb{R}^d}Q(f,f)\\log f dv \\geq 0.$$\nThe functional $-\\int f \\log f$ is the entropy of the solution and the Boltzmann's H theorem implies that at some point $x \\in \\mathbb{R}^d$, any equilibrium distribution function, i.e., any function maximizing the entropy, has the form of a locally Maxwellian distribution\n$$M(\\rho,u,T)(v)=\\frac{\\rho}{(2\\pi T)^{d\/2}}e^{-\\frac{|u-v|^2}{2T}},$$\nwhere $\\rho, u, T$ are respectively the density, mean velocity and temperature of the gas at the point $x$, defined by\n$$\\rho=\\int_{\\mathbb{R}^d}f(v)dv, \\quad u=\\frac{1}{\\rho}\\int_{\\mathbb{R}^d}vf(v)dv, \\quad T=\\frac{1}{N\\rho}\\int_{\\mathbb{R}^d}|u-v|^2f(v)dv.$$ \nFor further details on the physical background and derivation of the Boltzmann equation, we refer the reader to the extensive expositions \\cite{17,19,villani2}. \n\n\nFor monatomic gas, the non-negative cross section $B(z,\\sigma)$ defining the collision operator (\\ref{xiaoeq1}) depends only on $|z|$ and the scalar product \n$\\frac{z}{|z|}\\cdot \\sigma.$\nFurthermore, the cross section is assumed to be supported in the set where $\\frac{z}{|z|} \\cdot \\sigma \\geq 0$. This condition is not a restriction when dealing with the quadratic Boltzmann operator $Q(f,f)$. As noticed in~\\cite{alexandre0}, one may indeed reduce to this case after a symmetrization of the cross section since the term $f'f_*'$ appearing in the Boltzmann operator is invariant under the mapping $\\sigma \\rightarrow -\\sigma$. \n\n\nMore specifically, we consider cross sections of the type\n\\begin{equation}\\label{xiaoeq1.01}\nB(v-v_*,\\sigma)=\\Phi(|v-v_*|)b\\Big(\\frac{v-v_*}{|v-v_*|} \\cdot \\sigma\\Big), \\quad \\cos \\theta=\\frac{v-v_*}{|v-v_*|} \\cdot \\sigma, \\quad 0 \\leq \\theta \\leq \\frac{\\pi}{2},\n\\end{equation}\nwith a kinetic factor\n\\begin{equation}\\label{xiaosa0}\n\\Phi(|v-v_*|)=|v-v_*|^{\\gamma}, \\quad \\gamma \\in ]-d,+\\infty[,\n\\end{equation}\nand a factor related to the collision angle with a singularity \n\\begin{equation}\\label{xiaosa1}\n(\\sin \\theta)^{d-2} b(\\cos \\theta) \\sim K \\theta^{-1-2s},\n\\end{equation} \nwhen $\\theta \\to 0$, $\\theta>0$, for some constants $K>0$ and $0 < s <1$. Notice that this singularity is not integrable\n$$\\int_0^{\\frac{\\pi}{2}}(\\sin \\theta)^{d-2}b(\\cos \\theta)d\\theta=+\\infty.$$\nThis non-integrability property plays a major r\\^ole regarding the qualitative behavior of the solutions of the Boltzmann equation. Indeed, as first observed by L.~Desvillettes for the Kac equation~\\cite{D95}, grazing collisions that account for the non-integrability of the angular factor near $\\theta \\sim 0^+$, \ndo induce smoothing effects for the solutions of the non-cutoff Kac equation, or more generally for the solutions of the non-cutoff Boltzmann equation; whereas those solutions are at most as regular as the initial data, see e.g. \\cite{36}, when the collision cross section is assumed to be integrable, or after removing the singularity by using a cutoff function (Grad's angular cutoff assumption).\n\n\n\n\nBeing concerned with a close to equilibrium framework, the setting of the problem can be formulated as follows. First of all, without loss of generality, we consider the fluctuation around\n$$\\mu(v)=(2\\pi)^{-\\frac{d}{2}}e^{-\\frac{|v|^2}{2}},$$\nthe unique normalized equilibrium with mass 1, momentum 0 and temperature 1, by setting \n$$f=\\mu+\\sqrt{\\mu}g.$$ \nSince $Q(\\mu,\\mu)=0$ by the conservation of the kinetic energy\n$$|v|^2+|v_{\\ast}|^2=|v'|^2+|v_{\\ast}'|^2,$$\nthe Boltzmann collision operator can be split into three terms\n$$Q(\\mu+\\sqrt{\\mu}g,\\mu+\\sqrt{\\mu}g)=Q(\\mu,\\sqrt{\\mu}g)+Q(\\sqrt{\\mu}g,\\mu)+Q(\\sqrt{\\mu}g,\\sqrt{\\mu}g),$$\nwhose linearized part is \n$$Q(\\mu,\\sqrt{\\mu}g)+Q(\\sqrt{\\mu}g,\\mu).$$\nDefine\n$$\\mathcal{L}g=\\mathcal{L}_1g+\\mathcal{L}_2g,$$\nwith \n$$\\mathcal{L}_1g=-\\mu^{-1\/2}Q(\\mu,\\mu^{1\/2}g), \\quad \\mathcal{L}_2g=-\\mu^{-1\/2}Q(\\mu^{1\/2}g,\\mu),$$\nthe original Boltzmann equation is reduced to the Cauchy problem for the fluctuation\n$$\\begin{cases}\n\\partial_tg+v\\cdot\\nabla_{x}g+\\mathcal{L}g=\\mu^{-1\/2}Q(\\sqrt{\\mu}g,\\sqrt{\\mu}g)\\\\\ng|_{t=0}=g_0.\n\\end{cases}$$\nThis linearized Boltzmann operator $\\mathcal{L}$ is known (see e.g. \\cite{17}) to be an unbounded symmetric operator on $L^2$ (acting in the velocity variable) such that its Dirichlet form satisfies to\n$$\\mathbf{D}g=(\\mathcal{L}g,g)_{L^2} \\geq 0,$$\nand that \n$$\\mathbf{D}g=0 \\Leftrightarrow g=\\mathbf{P}g,$$ \nwhere\n$$\\mathbf{P}g=(a+b \\cdot v+c|v|^2)\\mu^{1\/2},$$ \nwith $a,c \\in \\mathbb{R}$, $b \\in \\mathbb{R}^d$, being the orthogonal projection onto the space of the so-called collisional invariants\n$$\\textrm{Span}\\big\\{\\mu^{1\/2},v_1 \\mu^{1\/2},...,v_d\\mu^{1\/2},|v|^2\\mu^{1\/2}\\big\\}.$$ \nFurthermore, recent works in particular those by P.L.~Lions \\cite{lions}, R.~Alexandre, L.~Desvillettes, C.~Villani, B.~Wennberg \\cite{alexandre0}, C.~Mouhot~\\cite{44}, C.~Mouhot, R.M.~Strain \\cite{strain},\nR. Alexandre, Y. Morimoto, C.-J. Xu, S. Ukai, T. Yang \\cite{amuxy5-1}\n or P.T.~Gressman, R.M. Strain \\cite{gress1,gress2} have highlighted that the non-cutoff Boltzmann operator enjoys remarkable coercive properties. Indeed, the results proved in~\\cite{44,strain} show that the linearized Boltzmann operator enjoys the following coercive estimate \n\\begin{equation}\\label{xiaosa2}\n(\\mathcal{L}g,g)_{L^2(\\mathbb{R}^d)} \\gtrsim \\|\\langle v \\rangle^{s+\\frac{\\gamma}{2}}(g-\\mathbf{P}g)\\|_{L^2(\\mathbb{R}^d)}^2+\\|g-\\mathbf{P}g\\|_{H^s_{\\textrm{loc}}(\\mathbb{R}^d)}^2,\n\\end{equation}\nwith $\\langle v \\rangle=(1+|v|^2)^{1\/2}$. This coercive estimate was improved in \\cite{amuxy5-1} by showing that the latter term $\\|g-\\mathbf{P}g\\|_{H^s_{\\textrm{loc}}(\\mathbb{R}^d)}^2$ in (\\ref{xiaosa2}) may be substituted by the global improvement \n$$\\|\\langle v \\rangle^{\\frac{\\gamma}{2}}(g-\\mathbf{P}g)\\|_{H^s(\\mathbb{R}^d)}^2.$$\nSee~\\cite{amuxy5-1,gress1,gress2} for optimal coercive estimates in appropriate weighted anisotropic Sobolev spaces. Regarding those coercive estimates,\nthe studies \\cite{alexandre0,lions} have decisively made clear that the smoothing\neffect of the non-cutoff Boltzmann equation is produced by the term\n$$\\int_{\\mathbb{R}^{2d}}\\int_{S^{d-1}}B(v-v_{*},\\sigma)g'_*(f- f')^2d\\sigma dvdv_*,$$\nthanks to the celebrated cancellation lemma, see~\\cite{alexandre0} (Lemma~1 and Proposition~2). The discovery of these special features of the non-cutoff Boltzmann operator have led those authors to conjecture that this collision operator behaves and induces smoothing effects as a fractional Laplacian $-(-\\Delta)^s$. See \\cite{alexandre0}, p.~331. This conjecture accounts for the choice of the operator (\\ref{yo2}) for the linearized non-cutoff Boltzmann equation. Notice that this operator is only a simplified model which does not account for the actual anisotropic features of the linearized non-cutoff Boltzmann equation and that current investigations are led to determine the exact microlocal structure of this operator. \nSee the first works in this direction by R.~Alexandre~\\cite{a1,a3,alexandre1}. \n\n\n\n\n\n\n\\subsection{Wick calculus}\\label{appendix}\n\n\n\nThe purpose of this appendix is to recall few facts and basic properties of the Wick quantization with parameter that are used in this paper. We refer the reader to the works of N.~Lerner \\cite{lerner}, \\cite{cubo} and \\cite{birkhauser} for thorough and extensive presentations of this quantization and some of its applications. \n\n\nThe main property of the Wick quantization is its property of positivity, i.e., that non-negative Hamiltonians define non-negative operators\n$$a \\geq 0 \\Rightarrow a^{\\textrm{Wick}(\\lambda)} \\geq 0.$$\nWe recall that this is not the case for the Weyl quantization nor the standard\nquantization and refer to \\cite{lerner} for an example of non-negative Hamiltonian defining an operator which is not non-negative. Before defining properly the Wick quantization, we first need to recall the definition of the wave-packets transform with parameter $\\lambda>0$ of a function $u \\in \\mathscr{S}(\\mathbb{R}^n)$, \n$$W_{\\lambda}u(y,\\eta)=(u,\\varphi_{y,\\eta}^{\\lambda})_{L^2(\\mathbb{R}^n)}, \\ (y,\\eta) \\in \\mathbb{R}^{2n}.$$\nwhere \n$$\\varphi_{y,\\eta}^{\\lambda}(x)=(2\\lambda)^{n\/4}e^{- \\pi \\lambda |x-y|^2}e^{2i \\pi (x-y) \\cdot \\eta}, \\ x \\in \\mathbb{R}^n,$$\nwith $|\\cdot|$ standing for the Euclidean norm and $x \\cdot y$ the standard dot product on $\\mathbb{R}^n$. With this definition, one can check (Lemma 1.3 in \\cite{cubo}) that \nthe mapping $u \\mapsto W_{\\lambda}u$ is continuous from $\\mathscr{S}(\\mathbb{R}^n)$ to $\\mathscr{S}(\\mathbb{R}^{2n})$, isometric from $L^{2}(\\mathbb{R}^n)$ to $L^2(\\mathbb{R}^{2n})$ and that we have the\nreconstruction formula\n\\begin{equation}\\label{lay0.1}\n\\forall u \\in \\mathscr{S}(\\mathbb{R}^n), \\forall x \\in \\mathbb{R}^n, \\ u(x)=\\int_{\\mathbb{R}^{2n}}{W_{\\lambda}u(y,\\eta)\\varphi_{y,\\eta}^{\\lambda}(x)dyd\\eta}.\n\\end{equation}\nBy denoting by $\\Sigma_Y^{\\lambda}$ the operator defined by the Weyl quantization of the symbol \n$$p_Y(X)=2^n e^{-2\\pi\\Gamma_{\\lambda}(X-Y)}, \\ X=(x,\\xi) \\in \\mathbb{R}^{2n}, \\ Y=(y,\\eta) \\in \\mathbb{R}^{2n},$$\nwith \n$$\\Gamma_{\\lambda}(X)=\\lambda |x|^2+\\frac{|\\xi|^2}{\\lambda}, \\ X=(x,\\xi) \\in \\mathbb{R}^{2n},$$\nwhich is a rank-one orthogonal projection,\n$$\\big{(}\\Sigma_Y^{\\lambda} u\\big{)}(x)=W_{\\lambda}u(Y)\\varphi_Y^{\\lambda}(x)=(u,\\varphi_Y^{\\lambda})_{L^2(\\mathbb{R}^n)}\\varphi_Y^{\\lambda}(x),$$\nwe define the Wick quantization with parameter $\\lambda>0$ of any $L^{\\infty}(\\mathbb{R}^{2n})$ symbol $a$ as\n\\begin{equation}\\label{lay0.2}\na^{\\textrm{Wick}(\\lambda)}=\\int_{\\mathbb{R}^{2n}}{a(Y)\\Sigma_Y^{\\lambda} dY}.\n\\end{equation}\nMore generally, one can extend this definition when the symbol $a$ belongs to $\\mathscr{S}'(\\mathbb{R}^{2n})$ by defining the operator $a^{\\textrm{Wick}(\\lambda)}$ for any $u$ and $v$ in $\\mathscr{S}(\\mathbb{R}^{n})$ by\n$$\\langle a^{\\textrm{Wick}(\\lambda)}u,\\overline{v}\\rangle_{\\mathscr{S}'(\\mathbb{R}^{n}), \\mathscr{S}(\\mathbb{R}^{n})}=\\langle a(Y),(\\Sigma_Y^{\\lambda}u,v)_{L^2(\\mathbb{R}^n)}\\rangle_{\\mathscr{S}'(\\mathbb{R}^{2n}), \\mathscr{S}(\\mathbb{R}^{2n})},$$\nwhere $\\langle \\textrm{\\textperiodcentered},\\textrm{\\textperiodcentered}\\rangle_{\\mathscr{S}'(\\mathbb{R}^n),\\mathscr{S}(\\mathbb{R}^n)}$ denotes the duality bracket between the\nspaces $\\mathscr{S}'(\\mathbb{R}^n)$ and $\\mathscr{S}(\\mathbb{R}^n)$. \nNotice from Proposition~3.1 in~\\cite{cubo} that \n\\begin{equation}\\label{lay10}\na^{\\textrm{Wick}(\\lambda)}=W_{\\lambda}^*a^{\\mu}W_{\\lambda}, \\ 1^{\\textrm{Wick}(\\lambda)}=\\textrm{Id}_{L^2(\\mathbb{R}^n)},\n\\end{equation}\nwhere $a^{\\mu}$ stands for the multiplication operator by $a$ on $L^2(\\mathbb{R}^{2n})$.\nThe Wick quantization with parameter $\\lambda>0$ is a positive quantization\n\\begin{equation}\\label{lay0.5}\na \\geq 0 \\Rightarrow a^{\\textrm{Wick}(\\lambda)} \\geq 0. \n\\end{equation}\nIn particular, real Hamiltonians get quantized in this quantization by formally self-adjoint operators and one has (Proposition 3.1 in \\cite{cubo}) that $L^{\\infty}(\\mathbb{R}^{2n})$ symbols define bounded operators on $L^2(\\mathbb{R}^n)$ such that \n\\begin{equation}\\label{lay0}\n\\|a^{\\textrm{Wick}(\\lambda)}\\|_{\\mathcal{L}(L^2(\\mathbb{R}^n))} \\leq \\|a\\|_{L^{\\infty}(\\mathbb{R}^{2n})}.\n\\end{equation}\nAccording to Proposition~3.2 in~\\cite{cubo} (see (51) in~\\cite{cubo}), the Wick and Weyl quantizations of a symbol $a$ are linked by the following identities\n\\begin{equation}\\label{lay1bis}\na^{\\textrm{Wick}(\\lambda)}=\\tilde{a}^w,\n\\end{equation}\nwith\n\\begin{equation}\\label{lay2bis}\n\\tilde{a}(X)=\\int_{\\mathbb{R}^{2n}}{a(X+Y)e^{-2\\pi \\Gamma_{\\lambda}(Y)}2^ndY}, \\ X \\in \\mathbb{R}^{2n},\n\\end{equation}\nand\n\\begin{equation}\\label{lay1}\na^{\\textrm{Wick}(\\lambda)}=a^w+r_{\\lambda}(a)^w,\n\\end{equation}\nwhere $r_{\\lambda}(a)$ stands for the symbol\n\\begin{equation}\\label{lay2}\nr_{\\lambda}(a)(X)=\\int_0^1\\int_{\\mathbb{R}^{2n}}{(1-\\theta)a''(X+\\theta Y)Y^2e^{-2\\pi \\Gamma_{\\lambda}(Y)}2^ndYd\\theta}, \\ X \\in \\mathbb{R}^{2n}.\n\\end{equation}\nWe recall that we use here the following normalization for the Weyl quantization\n\\begin{equation}\\label{lay3}\n(a^wu)(x)=\\int_{\\mathbb{R}^{2n}}{e^{2i\\pi(x-y)\\cdot\\xi}a\\big(\\frac{x+y}{2},\\xi\\big)u(y)dyd\\xi}.\n\\end{equation}\nLet us finally mention that the operator $\\pi_{\\lambda}=W_{\\lambda}W_{\\lambda}^*$ is an orthogonal projection on a closed proper subspace of $L^2(\\mathbb{R}^{2n})$ whose kernel is given by\n\\begin{equation}\\label{yo1}\nK_{\\lambda}(X,Y)=e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}e^{i\\pi (x-y)\\cdot(\\xi+ \\eta)}, \\ X=(x,\\xi) \\in \\mathbb{R}^{2n}, \\ Y=(y,\\eta) \\in \\mathbb{R}^{2n}. \n\\end{equation}\nIndeed, for any $u \\in \\mathscr{S}(\\mathbb{R}_x^n)$ and $v \\in \\mathscr{S}(\\mathbb{R}_Y^{2n})$, we may write\n\\begin{multline*}\n(W_{\\lambda}u,v)_{L^2(\\mathbb{R}_Y^{2n})}=\\int_{\\mathbb{R}_Y^{2n}}\\Big(\\int_{\\mathbb{R}_x^n}u(x)\\overline{\\varphi_{Y}^{\\lambda}}(x)dx\\Big)\\overline{v}(Y)dY\\\\ =\n\\int_{\\mathbb{R}_x^{n}}u(x)\\Big(\\overline{\\int_{\\mathbb{R}_Y^{2n}}\\varphi_{Y}^{\\lambda}(x)v(Y)dY}\\Big)dx\n=(u,W_{\\lambda}^*v)_{L^2(\\mathbb{R}_x^n)}.\n\\end{multline*}\nIt follows that \n$$(W_{\\lambda}^*v)(x)=\\int_{\\mathbb{R}_Y^{2n}}\\varphi_{Y}^{\\lambda}(x)v(Y)dY.$$\nWriting \n\\begin{multline*}\n(\\pi_{\\lambda}v)(X)=(W_{\\lambda}W_{\\lambda}^*v)(X)=\\int_{\\mathbb{R}_t^n}\\Big(\\int_{\\mathbb{R}_Y^{2n}}\\varphi_{Y}^{\\lambda}(t)v(Y)dY\\Big)\\overline{\\varphi_{X}^{\\lambda}}(t)dt\\\\\n=\\int_{\\mathbb{R}_Y^{2n}}\\Big(\\int_{\\mathbb{R}_t^{n}}\\varphi_{Y}^{\\lambda}(t)\\overline{\\varphi_{X}^{\\lambda}}(t)dt\\Big)v(Y)dY=\\int_{\\mathbb{R}_Y^{2n}}K_{\\lambda}(X,Y)v(Y)dY,\n\\end{multline*}\nwith \n$$K_{\\lambda}(X,Y)=\\int_{\\mathbb{R}_t^{n}}\\varphi_{Y}^{\\lambda}(t)\\overline{\\varphi_{X}^{\\lambda}}(t)dt.$$\nBy using the identity\n$$2|t-y|^2+2|t-x|^2=|x-y|^2+|x+y-2t|^2,$$\nan explicit computation gives that \n\\begin{align*}\nK_{\\lambda}(X,Y)=& \\ (2\\lambda)^{\\frac{n}{2}}\\int_{\\mathbb{R}_t^{n}}e^{-\\pi \\lambda |t-y|^2}e^{-\\pi \\lambda |t-x|^2}e^{2i\\pi(t-y)\\cdot \\eta}e^{-2i\\pi(t-x)\\cdot \\xi}dt\\\\\n= & \\ (2\\lambda)^{\\frac{n}{2}}e^{-\\frac{\\pi}{2}\\lambda |x-y|^2}e^{2i\\pi (x\\cdot \\xi-y\\cdot \\eta)}\\int_{\\mathbb{R}_t^{n}}e^{-\\frac{\\pi}{2} \\lambda |x+y-2t|^2}e^{2i\\pi t \\cdot (\\eta-\\xi)}dt.\n\\end{align*}\nSince \n\\begin{multline*}\n\\int_{\\mathbb{R}_t^{n}}e^{-\\frac{\\pi}{2} \\lambda |x+y-2t|^2}e^{2i\\pi t \\cdot (\\eta-\\xi)}dt=\\int_{\\mathbb{R}_t^{n}}e^{-2\\pi \\lambda |t-\\frac{x+y}{2}|^2}e^{2i\\pi t \\cdot (\\eta-\\xi)}dt\n\\\\ =e^{i\\pi (x+y) \\cdot (\\eta-\\xi)}\\int_{\\mathbb{R}_t^{n}}e^{-2\\pi \\lambda |t|^2}e^{2i\\pi t \\cdot (\\eta-\\xi)}dt=(2\\lambda)^{-\\frac{n}{2}}e^{i\\pi (x+y) \\cdot (\\eta-\\xi)}e^{-\\frac{\\pi}{2\\lambda} |\\eta-\\xi|^2},\n\\end{multline*}\nit follows that \n\\begin{align*}\nK_{\\lambda}(X,Y)= & \\ e^{-\\frac{\\pi}{2}\\lambda |x-y|^2}e^{-\\frac{\\pi}{2\\lambda} |\\eta-\\xi|^2}e^{2i\\pi (x\\cdot \\xi-y\\cdot \\eta)}e^{i\\pi (x+y) \\cdot (\\eta-\\xi)}\\\\\n= & \\ e^{-\\frac{\\pi}{2}\\Gamma_{\\lambda}(X-Y)}e^{i\\pi (x-y)\\cdot(\\xi+ \\eta)}.\n\\end{align*}\n\n\n\n\\bigskip\n\\bigskip\n\n\\noindent\n\\textbf{Acknowledgements.} \nThe research of the second author was\nsupported by Grant-in-Aid for Scientific Research No.22540187,\nJapan Society of the Promotion of Science. \nThe third author is very grateful to the Japan Society for the Promotion of Science and JSPS London for supporting his stay at Kyoto University during the summer 2010. He would like to thank Kyoto University and JSPS very warmly for their very kind hospitality and the exceptional working surroundings in which this work was done. Finally, all the three authors are very grateful to the referees for enriching comments and relevant suggestions which have helped to improve significantly the presentation of this article.\n\n\n\n\n\n\n\n\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section*{Abstract}\nEmbryonic development is driven by spatial patterns of gene expression\nthat determine the fate of each cell in the embryo. While gene\nexpression is often highly erratic, embryonic development is usually\nexceedingly precise. In particular, gene expression boundaries are\nrobust not only against intra-embryonic fluctuations such as noise in\ngene expression and protein diffusion, but also against\nembryo-to-embryo variations in the morphogen gradients, which provide\npositional information to the differentiating cells. How development\nis robust against intra- and inter-embryonic variations is not\nunderstood. A common motif in the gene regulation networks that\ncontrol embryonic development is mutual repression between pairs of\ngenes.\nTo assess the role of mutual repression in the robust formation\nof gene expression patterns, we have performed large-scale stochastic\nsimulations of a minimal model of two mutually repressing gap genes \nin {\\it Drosophila}, {\\it hunchback} (\\textit{hb}\\xspace) and {\\it knirps} (\\textit{kni}\\xspace).\nOur model includes not only mutual repression between \\textit{hb}\\xspace and \\textit{kni}\\xspace,\nbut also the stochastic and cooperative activation of \\textit{hb}\\xspace by the anterior\nmorphogen Bicoid (Bcd\\xspace) and of \\textit{kni}\\xspace by the posterior morphogen Caudal\n(Cad\\xspace), as well as the diffusion of Hb\\xspace and Kni\\xspace between neighboring\nnuclei.\nOur analysis reveals that mutual repression can markedly increase the \nsteepness and precision of the gap gene expression boundaries. \nIn contrast to other mechanisms such as spatial averaging and cooperative \ngene activation, mutual repression thus allows for gene-expression \nboundaries that are both steep and precise. Moreover,\nmutual repression dramatically enhances their robustness against\nembryo-to-embryo variations in the morphogen levels. Finally, our\nsimulations reveal that diffusion of the gap proteins plays a critical\nrole not only in reducing the width of the gap gene expression\nboundaries via the mechanism of spatial averaging, but also in\nrepairing patterning errors that could arise because of the\nbistability induced by mutual repression.\n\n\n\\pagebreak\n\n\\section*{Introduction}\n\\pdfbookmark[1]{Introduction}{BookmarkIntro}\nThe development of multicellular organisms requires spatially\ncontrolled cell differentiation. The positional information for the\ndifferentiating cells is typically provided by spatial concentration gradients of\nmorphogen proteins. In the\nclassical picture of morphogen-directed patterning, cells translate\nthe morphogen concentration into spatial gene-expression domains via\nsimple threshold-dependent readouts\n\\cite{Wolpert1969,Wolpert1994,Driever1988a,Driever1988b}. Yet, while\nembryonic development is exceedingly precise, this mechanism is not\nvery robust against intra- and inter-embryonic variations\n\\cite{Houchmandzadeh2002,Gregor2007,Gregor2007b}: the spatial patterns\nof the target genes do not scale with the size of the embryo and the\nboundaries of the expression domains are susceptible to fluctuations\nin the morphogen levels and to the noise in gene\nexpression. Intriguingly, the target genes of morphogens often\nmutually repress each other, as in the gap-gene system of the fruit\nfly {\\em\n Drosophila} \\cite{Jackle1986,Clyde2003,\n Jaeger2004,Surkova2008,Manu2009PlosCompBiol,Manu2009PlosBiol,Vakulenko2009}.\nTo elucidate the role of mutual repression in the robust formation of\ngene expression patterns, we have performed extensive\nspatially-resolved stochastic simulations of the gap-gene system of\n{\\em Drosophila melanogaster}. Our results show that mutual repression between\ntarget genes can markedly enhance both the steepness and the precision\nof gene-expression boundaries. Furthermore, it makes them robust against\nembryo-to-embryo variations in the morphogen gradients.\n\nThe fruit fly \\textit{Drosophila melanogaster} is arguably the\nparadigm of morphogenesis. During the first 90 minutes after\nfertilization it is a syncytium, consisting of a cytoplasm that\ncontains rapidly diving nuclei, which are not yet encapsulated by\ncellular membranes. Around cell cycle 10 the nuclei migrate towards\nthe cortex of the embryo and settle there to read out the\nconcentration gradient of the morphogen protein Bicoid (Bcd\\xspace), which\nforms from the anterior pole after fertilization \\cite{Driever1988a}.\nOne of the target genes of Bcd\\xspace is the gap gene {\\it hunchback} (\\textit{hb}\\xspace),\nwhich is expressed in the anterior half of the embryo. In spite of\nnoise in gene expression, the midembryo boundary of the \\textit{hb}\\xspace expression\ndomain is astonishingly sharp. By cell cycle 11, the \\textit{hb}\\xspace mRNA\nboundary varies by about one nuclear spacing only \\cite{Porcher2010,\nHe2011, Perry2011}, while by cell cycle 13 a similarly sharp\noundary is observed for the protein level \\cite{Houchmandzadeh2002,Gregor2007,He2008}.\nThis precision is higher than the best achievable precision for a \ntime-averaging based readout mechanism of the Bcd gradient \\cite{Gregor2007}.\nInterestingly, the\nstudy of Gregor {\\it et al.} revealed that the Hb concentrations in\nneighboring nuclei exhibit correlations and the authors suggested that\nthis implies a form of spatial averaging that enhances the precision\nof the posterior Hb boundary \\cite{Gregor2007}. Two recent simulation\nstudies suggest that the mechanism of spatial averaging is based on\nthe diffusion of Hb itself \\cite{Erdmann2009,Okabe-Oho2009}; as shown\nanalytically in \\cite{Erdmann2009}, Hb diffusion between neighboring\nnuclei reduces the super-Poissonian part of the noise in its\nconcentration. In essence, diffusion reduces noise by washing out\nbursts in gene expression. However, the mechanism of spatial averaging\ncomes at a cost: it tends to lessen the steepness of the expression\nboundaries.\n\nBcd induces the expression of not only \\textit{hb}\\xspace, but a number of gap genes,\nand pairs of gap genes tend to repress each other\nmutually. Interestingly, repression between directly neighboring gap\ngenes is weak, whereas repression between non-adjacent genes is strong\n\\cite{Kraut1991}. \\textit{hb}\\xspace forms a strongly repressive pair with {\\it knirps}\n(\\textit{kni}\\xspace) which is expressed further towards the posterior pole; \nboth genes play a prominent role in the later positioning of\ndownstream pair-rule gene stripes \\cite{Clyde2003}.\nIt has\nbeen argued that mutual repression can enhance robustness to\nembryo-to-embryo variations in morphogen levels\n\\cite{Manu2009PlosCompBiol,Manu2009PlosBiol,Vakulenko2009} and sharpen\na morphogen-induced transition between the two mutually \nrepressing genes in a non-stochastic background \\cite{Saka2007,Ishihara2008}.\nHowever, mutual repression can also lead to bistability\n\\cite{Cherry2000,Kepler2001,Warren2004,Warren2005,Papatsenko2011}. While bistablity may buffer\nagainst inter-embryo variations and rapid intra-embryo fluctuations in\nmorphogen levels, it may also cause stochastic switching between\ndistinct gene expression patterns, which would be highly detrimental.\nTherefore, the precise role of mutual repression in the robust\nformation of gene-expression patterns remains to be elucidated.\n\nWhile the role of antagonistic interactions in the formation of\ngene-expression patterns has been studied using mean-field models\n\\cite{Manu2009PlosCompBiol,Papatsenko2011,Ishihara2005,Zinzen2006,Zinzen2007},\nto address the question whether mutual repression enhances the\nrobustness of these patterns against noise arising from the inherent\nstochasticity of biochemical reactions a stochastic model is\nessential. We have therefore performed large-scale stochastic\nsimulations of a minimal model of mutual repression between \\textit{hb}\\xspace and\n\\textit{kni}\\xspace. Our model includes the stochastic and cooperative activation\nof \\textit{hb}\\xspace by Bcd\\xspace and of \\textit{kni}\\xspace by the posterior morphogen Caudal (Cad\\xspace)\n\\cite{Rivera-Pomar1995,Schulz1995}. Moreover, Hb\\xspace and Kni\\xspace can diffuse\nbetween neighboring nuclei and repress each other's expression,\ngenerating two separate spatial domains interacting at midembryo (see\nFig.\\xspace \\ref{Fig1}). We analyze the stability of these domains by\nsystematically varying the diffusion constants of the Hb\\xspace and Kni\\xspace\nproteins, the strength of mutual repression and the Bcd\\xspace and Cad\\xspace\nactivator levels. To quantify the importance of mutual\nrepression, we compare the results to those of a system containing\nonly a single gap gene, which is regulated by its morphogen only;\nthis is the ``system without mutual repression''. While our model is\nsimplified---it neglects, {\\it e.g.}, the interactions of \\textit{hb}\\xspace and \\textit{kni}\\xspace\nwith {\\it kr\\\"{u}ppel} (\\textit{kr}\\xspace) and {\\it giant} (\\textit{gt}\\xspace)\n\\cite{Jaeger2011}---it does allow us to elucidate the mechanism by\nwhich mutual repression can enhance the robust formation of gene\nexpression patterns.\n\nOne of the key findings of our analysis is that mutual repression\nenhances the robustness of the gene expression domains against\nintra-embryonic fluctuations arising from the intrinsic\nstochasticity of biochemical reactions. Specifically, mutual\nrepression increases the precision of gene-expression\nboundaries: it reduces the variation $\\Delta x$ in their\npositions due to these fluctuations. At the same time, mutual\nrepression also enhances the steepness of the expression boundaries.\nTo understand the interplay between steepness, precision and\nintra-embryonic fluctuations (biochemical noise), it is instructive\nto recall that the width $\\Delta x$ of a boundary of the expression\ndomain of a gene $g$ is, to first order, given by\n\\begin{equation}\n\\label{EqDeltaX}\n\\Delta x =\\frac{\\sigma_G (x_t)}{|\\Avg{G(x_t)}^\\prime|},\n\\end{equation}\nwhere $\\sigma_G(x_t)$ is the standard deviation of the copy number $G$\nof protein G\\xspace and $|\\Avg{G(x_t)}^\\prime|$ is the magnitude of the\ngradient of $G$ at the boundary position $x_t$\n\\cite{Gregor2007,Tostevin2007,Erdmann2009}. Steepness thus refers to\nthe slope of the average concentration profile,\n$|\\Avg{G(x_t)}^\\prime|$, while precision refers to $\\Delta x$,\nwhich is the standard deviation in the position at which $G$\ncrosses a specified threshold value, here taken to be the\nhalf-maximal average expression level of $G$. \n\nThe simulations reveal, perhaps surprisingly, that mutual repression\nhardly affects the noise $\\sigma_G(x_t)$ at the expression boundaries\nof \\textit{hb}\\xspace and \\textit{kni}\\xspace. Moreover, mutual repression can strongly enhance the\nsteepness $|\\Avg{G(x_t)}^\\prime|$ of these boundaries: the steepness\nof the boundaries in a system with mutual repression can, depending on\nthe diffusion constant, be twice as large as that in the system\nwithout mutual repression. Together with Eq. \\ref{EqDeltaX}, these\nobservations predict that mutual repression can significantly enhance\nthe precision of the boundaries, i.e. decrease $\\Delta x$, which\nis indeed precisely what the simulations reveal. Interestingly, there\nexists an optimal diffusion constant that minimizes the boundary width\n$\\Delta x$, as has been observed for a system without mutual\nrepression \\cite{Erdmann2009}. While the minimal $\\Delta x$ of the\nsystem with mutual repression is only marginally lower than that of\nthe system without it, this optimum is reached at a lower value of the\ndiffusion constant, where the steepness of the boundaries is much\nhigher. We find that these observations are robust,\ni.e. independent of the precise parameters of the model, such as\nmaximum expression level, size of the bursts of gene expression, and\nthe cooperativity of gene activation.\n\nOur results also show that mutual repression can strongly buffer\nagainst embryo-to-embryo variations in the morphogen levels by\nsuppressing boundary shifts via a mechanism that is akin to that of\n\\cite{Howard2005, Morishita2009}. A more detailed analysis reveals\nthat when the regions where Bcd\\xspace and Cad\\xspace activate \\textit{hb}\\xspace and \\textit{kni}\\xspace\nrespectively overlap, bistability can arise in the overlap zone. Yet,\nthe mean waiting time for switching is longer than the lifetime of the\nmorphogen gradients, which means that the \\textit{hb}\\xspace and \\textit{kni}\\xspace expression\npatterns are stable on the relevant developmental time scales. This\nalso means, however, that when errors are formed during development,\nthese cannot be repaired. Here, our simulations reveal another\nimportant role for diffusion: without diffusion a spotty phenotype\nemerges in which the nuclei in the overlap zone randomly express either\nHb\\xspace or Kni\\xspace; diffusion can anneal these patterning defects, leading\nto well-defined expression domains of Hb\\xspace and Kni\\xspace. \nFinally, we also study a scenario where \\textit{hb}\\xspace and \\textit{kni}\\xspace are\nactivated by Bcd\\xspace only. While this scheme is not robust against\nembryo-to-embryo variations in the morphogen levels, mutual\nrepression does enhance boundary precision and steepness also\nin this scenario.\n\n\\section*{Results}\n\\pdfbookmark[1]{Results}{BookmarkResults}\n\\subsection*{Model}\n\\pdfbookmark[2]{Model}{BookmarkModel}\nWe consider the embryo in the syncytial blastoderm stage at late cell cycle\n14, ca. $2~\\unit{h}$ after fertilization. In this stage the majority of the nuclei\nforms a cortical layer and \\textit{hb}\\xspace and \\textit{kni}\\xspace expression\ncan be detected \\cite{Surkova2008}. Our model is an extension of the\none presented in \\cite{Erdmann2009}. It is based on a cylindrical\narray of diffusively coupled reaction volumes which represent the\nnuclei, with periodic boundary conditions in the angular ($\\phi$)\nand reflecting boundaries in the axial ($x$) direction. The\ndimensions of the cortical array are $N_x=N_\\phi=64$, with equal\nspacing of the nuclei $\\ell=8.5 \\unit{\\mu m}$ in both directions. For a given embryo length\n$L$, this implies a cylinder radius\n$R=\\frac{L}{2\\pi}\\simeq\\frac{L}{6}$, which is close to the\nexperimentally observed ratio.\nThe resulting number of $N=4096$ nuclei roughly corresponds to the\nexpected number of cortical nuclei at cell cycle 14 if non-dividing\npolyploid yolk nuclei are taken into account \\cite{Foe1983} (see\nText S1\\xspace for details); we also emphasize, however, that none of\nthe results presented below depend on the precise number of nuclei.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=3.27in]{Figure1.pdf}\n\\end{center}\n\\caption{ {\\bf The model.} \\subfig{A} Cartoon of our model. Bcd\\xspace\n activates \\textit{hb}\\xspace, while its antagonist \\textit{kni}\\xspace is activated by Cad\\xspace. The\n gap genes \\textit{hb}\\xspace and \\textit{kni}\\xspace repress each other mutually. In each nuclear\n compartment we simulate the genetic promoters of both \\textit{hb}\\xspace and\n \\textit{kni}\\xspace. Activation is cooperative: In the default setting, 5 morphogen proteins have to bind to\n the promoter to initiate gene expression. Hb\\xspace and Kni\\xspace both form\n homodimers, which can bind to the other gene's promoter to totally\n block expression, irrespective of the number of bound morphogen\n proteins. Both dimers and monomers travel between neighboring\n nuclear compartments via diffusion. \\subfig{B} Protein copy number\n profiles along the AP axis in a typical simulation in steady state,\n with parameter values as in Table S2 in Text S1\\xspace. Plotted are\n the morphogen gradients Bcd\\xspace ($\\Avg{B}$, solid green line) and Cad\\xspace \n ($\\Avg{C}$, solid red line) and the resulting Hb\\xspace ($H$) and Kni\\xspace ($K$) total copy \n number profiles for different times. The dashed green and red lines show the\n Hb\\xspace ($\\Avg{H}$) and Kni\\xspace ($\\Avg{K}$) profiles averaged over time and the \n circumference of the (cylindrical) system.\n }\n\\label{Fig1}\n\\end{figure}\n\nIn each nuclear volume we simulate the activation of the gap genes \\textit{hb}\\xspace\nand \\textit{kni}\\xspace by the morphogens Bcd\\xspace and Cad\\xspace, respectively, and mutual\nrepression between \\textit{hb}\\xspace and \\textit{kni}\\xspace (see Fig.\\xspace \\ref{Fig1}). In what\nfollows, we will refer to Hb\\xspace and Kni\\xspace as repressors and to Bcd\\xspace and\nCad\\xspace as activators. Our model of gene regulation bears\nsimilarities to those of\n\\cite{Bolouri2003,Janssens2006,Zinzen2006,Zinzen2007,Papatsenko2011},\nin the sense that it is based on a statistical mechanical model of\ngene regulation by transcription factors, allowing the computation\nof promoter-site occupancies. However, the models of\n\\cite{Bolouri2003,Janssens2006,Zinzen2006,Zinzen2007,Papatsenko2011}\nare mean-field models, which cannot capture the effect of\nintra-embryonic fluctuations due to biochemical noise arising from\nthe inherent stochasticity of biochemical reactions. This requires a\nstochastic model; moreover, it necessitates a model in which the\ntransitions between the promoter states are taken into account\nexplicitly, since these transitions form a major source of noise in\ngene expression, as we will show. To limit the number of\ncombinatorial promoter states, we have therefore studied a minimal model\nthat only includes Bcd\\xspace, Cad\\xspace, Hb\\xspace and Kni\\xspace.\nFollowing \\cite{Erdmann2009}, we assume that Bcd\\xspace and Cad\\xspace bind\nstochastically and cooperatively to $n_{max}$ sites on their target\npromoters. To obtain a lower bound on the precision of the \\textit{hb}\\xspace and\n\\textit{kni}\\xspace expression domains, we assume that the activating morphogens Bcd\\xspace\nand Cad\\xspace bind to their promoters with a diffusion-limited rate $k_{\\rm on}^{\\rm A}=4\\pi\\alpha D_A\/V$,\nwhere $\\alpha$ is the dimension of a binding site, $D_A$ is the \ndiffusion constant of the morphogen, and $V$ is the nuclear volume \n(see ``Materials \\& Methods'' for parameter values). Since\nthe morphogen-promoter association rate is assumed to be diffusion\nlimited, cooperativity of \\textit{hb}\\xspace and \\textit{kni}\\xspace activation is tuned via the\ndissociation rate $k_{\\rm off,n}^{\\rm A}=a\/b^n$, which decreases with\nincreasing number $n$ of promoter-bound morphogen molecules. The\nbaseline parameters are set such that the half-maximal activation\nlevel of \\textit{hb}\\xspace and \\textit{kni}\\xspace is at midembryo, and the effective Hill\ncoefficient for gene activation is around 5 \\cite{Erdmann2009}; while\nwe will vary the Hill coefficient, this is our baseline\nparameter. Again to obtain a lower bound on the precision of the\ngap-gene expression boundaries, transcription and translation is\nconcatenated in a single step. Mutual repression between \\textit{hb}\\xspace and \\textit{kni}\\xspace\noccurs via binding of Hb\\xspace to the \\textit{kni}\\xspace promoter, which blocks the\nexpression of \\textit{kni}\\xspace irrespective of the number of bound Cad\\xspace molecules,\nand vice versa. To assess the importance of bistability, Hb\\xspace and Kni\\xspace\ncan homodimerize and bind to their target promoters only in their\ndimeric form, which is a prerequisite for bistability in the\nmean-field limit \\cite{Cherry2000}. Both the monomers and dimers\ndiffuse between neighboring nuclei and are also degraded; the\neffective degradation rate $\\mu_{\\rm eff}$ is such that the gap-gene\nexpression domains can form sufficiently rapidly on the time scale of\nembryonic development ($\\approx 10-20~\\unit{min}$ \\cite{Foe1983}).\nIn the absence of mutual repression, our model\nbehaves very similarly to that of \\cite{Erdmann2009}, even though our\nmodel contains both monomers and dimers instead of only monomers.\n\n\nMotivated by experiment \\cite{Driever1988a,Houchmandzadeh2002,Gregor2007b},\nand in accordance with the diffusion-degradation model,\nwe adopt an exponential shape for the stationary Bcd\\xspace profile; we thus \ndo not model the establishment of the gradient \\cite{Sample2010}.\nTo elucidate the role of mutual repression,\nit will prove useful to take our model to be\nsymmetric: the Cad\\xspace profile is the mirror image of the Bcd\\xspace\nprofile, and \\textit{hb}\\xspace and \\textit{kni}\\xspace repress each other equally\nstrongly. Diffusion of Bcd\\xspace and Cad\\xspace between nuclei induce\nfluctuations in their copy numbers on the time scale\n$\\tau_d=\\ell^2\/(4D_A)\\simeq 6~\\unit{s}$. Because $\\tau_d$ is much\nsmaller than the time scale for promoter binding, $1\/k_{\\rm\n on}^{\\rm A}\\simeq 360~\\unit{s}$, fluctuations in the copy number\nof Bcd\\xspace and Cad\\xspace are effectively averaged out by slow binding of Bcd\\xspace\nand Cad\\xspace to their respective promoters, \\textit{hb}\\xspace and \\textit{kni}\\xspace\n\\cite{Erdmann2009}. To elucidate the importance of the threshold\npositions for \\textit{hb}\\xspace and \\textit{kni}\\xspace activation, we will scale the morphogen\ngradients by a global dosage factor $A$; this procedure will also\nallow us to study the robustness of the system against\nembryo-to-embryo variations in the morphogen levels.\n\nWe simulate the model using the Stochastic Simulation Algorithm (SSA)\nof Gillespie \\cite{Gillespie1976, Gillespie1977}. Diffusion is\nimplemented into the scheme via the next-subvolume method used in\nMesoRD \\cite{Elf2004, Hattne2005}.\nA recent version of our code is available at GitHub and can be\naccessed via \\url{http:\/\/ggg.amolf.nl} .\n\n\\subsection*{Characteristics of gap-gene expression boundaries}\n\\pdfbookmark[2]{Characteristics of gap-gene expression boundaries}{BookmarkBoundaries}\nThree key characteristics of gene expression boundaries are 1) the\nnoise in the protein concentration at the boundary; 2) the steepness\nof the boundary; 3) the width of the boundary. While these quantities\nmay make intuitive sense, their definitions are not unambiguous. Equally important, different definitions will\nreveal different properties of the system.\n\n\\mysubsubsection{Decomposing the noise} Let's consider the variance in the copy number $G$ of protein G\\xspace\n at position $x$ along the anterior-posterior (AP) axis. We define\n its mean copy number, averaged over all embryos, circumferential positions\n $\\phi$ and all times, at the anterior-posterior position $x$ as\n\\begin{eqnarray}\n\\ETPAvg{G}(x)&\\equiv&\\frac{1}{N_e}\\frac{1}{T}\\frac{1}{N_\\phi} \\sum_{e=0}^{N_e-1}\\sum_{t=0}^{T-1}\\sum_{\\phi=0}^{N_\\phi-1}G_e(\\phi,x,t),\n\\end{eqnarray}\nwhere $G_e(x,\\phi,t)$ is the copy number of protein G\\xspace in embryo $e$ at position $x$\n and angle $\\phi$ in the\ncircumferential direction (perpendicular to the AP-axis) at time\n$t$. Here, we introduce the convention that the overline denotes\nan average in time, while the ensemble brackets with\na subscript $\\phi$ \ndenote an average along the $\\phi$ direction and that with a\nsubscript $e$ an average over all embryos. The variance in the copy number\n$G\\equiv G_e(x,\\phi,t)$ is then given by\n\\begin{eqnarray}\n\\sigma^2_G(x)&=&\\ETPAvg{(G-\\ETPAvg{G})^2}\\\\\n&=&\\Avg{\\TPAvg{G^2}}_e-\\Avg{\\overline{\\Avg{G}_\\phi^2}}_e+\\Avg{\\overline{\\Avg{G}_\\phi^2}}_e-\\Avg{\\TPAvg{G}^2}_e+\\Avg{\\TPAvg{G}^2}_e-\\ETPAvg{G}^2\\\\\n&=&\\overbrace{\\Avg{\\overline{\\sigma^2_G}}_e(x)+\\Avg{\\sigma^2_{\\Avg{G}_\\phi}}_e(x)}^{\\text{mean\n intra-embryonic\n noise}}\\hspace{2.76cm}+\\overbrace{\\sigma^2_{\\Avg{\\overline{G}}_\\phi}(x)}^{\\text{inter-embryonic variations}}\n\\label{EqNoiseDecom}\n\\end{eqnarray}\nThe total variance in the copy number can thus be decomposed into\nintra-embryonic fluctuations averaged over all embryos and\ninter-embryonic variations. The former can, furthermore, be decomposed into\n$\\Avg{\\overline{\\sigma^2_G}}_e(x)$, which is the time-averaged mean of\nthe variance in $G$ along the circumferential direction,\n$\\overline{\\sigma^2_G}(x)$, averaged over all embryos, and\n$\\Avg{\\sigma^2_{\\Avg{G}_\\phi}}_e(x)$, which is the variance in time\nover the mean of $G$ along the circumferential direction,\n$\\sigma^2_{\\Avg{G}_\\phi}(x)$, again averaged over all embryos. These\nintra-embryonic terms capture different types of dynamics. If the\nexpression boundary is rough\nbut its average position does not fluctuate in time,\nthen $\\overline{\\sigma^2_G}(x)$ will be large yet\n$\\sigma^2_{\\Avg{G}_\\phi}(x)$ will be small. Conversely, when the\nboundary is smooth but its average position does fluctuate in time, then\n$\\overline{\\sigma^2_G}(x)$ will be small yet\n$\\sigma^2_{\\Avg{G}_\\phi}(x)$ will be large. Naturally, a combination\nof the two is also possible. The third term,\n$\\sigma^2_{\\Avg{\\overline{G}}_\\phi}(x)$, captures the embryo-to-embryo\nvariations in the average over time and $\\phi$ of the protein-copy\nnumber. Similarly, we can decompose the fluctuations in the boundary\nposition $x_t$ as\n\\begin{eqnarray}\n\\Delta x &=&\\sigma_{x_t}\\\\\n&=&\\sqrt{\\Avg{\\overline{\\sigma^2_{x_t}}}_e+\\Avg{\\sigma^2_{\\Avg{x_t}_\\phi}}_e+\\sigma^2_{\\Avg{\\overline{x_t}}_\\phi}}\n\\label{EqXtDecom}\n\\end{eqnarray}\nThe two different contributions to the intra-embryonic variance, $\\Avg{\\overline{\\sigma^2_{x_t}}}_e+\\Avg{\\sigma^2_{\\Avg{x_t}_\\phi}}_e$, are illustrated\n in Fig.\\xspace \\ref{Fig2}.\nHere and in the next section, we will study the robustness of the\nsystem against intra-embryonic fluctuations, while in the section\n``Robustness to inter-embryonic variations: Mutual repression can\nbuffer against correlated morphogen level variations'' we will study\nthe robustness against inter-embryonic variations in the morphogen levels.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\vspace{0.33in}\n\\includegraphics[width=3.27in]{Figure2.pdf}\n\\end{center}\n\\caption{ {\\bf Two different contributions to the intra-embryonic\n variance in the boundary position.} The total variance of the\n gap gene expression boundary position $x_t$ due to intra-embryonic\n fluctuations, $\\sigma^2_{x_t, intra}$, can be decomposed into two\n contributions: $\\sigma^2_{\\Avg{x_t}_\\phi}$, the variance in time\n of the circumferential mean of $x_t$, and\n $\\overline{\\sigma^2_{x_t}}$, the time-average of the variance of\n $x_t$ along the circumference of the embryo. The sketch illustrates\n two extremal cases: If the boundary is very smooth along the\n circumference at any moment in time, concerted movements of the boundary\n will dominate the total variance, i.e. $\\sigma^2_{x_t, intra}\n \\simeq \\sigma^2_{\\Avg{x_t}_\\phi}$ (left side). If, in contrast,\n the boundary is rough but its mean position does not fluctuate\n much in time, then $\\sigma^2_{x_t, intra} \\simeq\n \\overline{\\sigma^2_{x_t}}$ (right side). Naturally, a combination\n of the two types of fluctuations is possible. }\n\\label{Fig2}\n\\end{figure}\n\n{\\mysubsubsection{Intra-embryonic fluctuations} Fig.\\xspace S2 in Text S1\\xspace\n shows the decomposition of the noise in the Hb\\xspace copy number\n $H$ and the threshold position $x_t$ of the Hb\\xspace boundary, as a\n function of the diffusion constant. We show the intra-embryonic\n fluctuations for one given embryo (with the baseline parameter set);\n how $\\Delta x$ (the boundary variance originating from\n intra-embryonic fluctuations) changes with embryo-to-embryo\n variations in the morphogen levels is addressed in section ``Overlap\n of morphogen activation domains does not corrupt robustness to\n intrinsic fluctuations''. Fig.\\xspace S2 shows that by far the dominant\n contribution to the intra-embryonic noise in the copy number and\n threshold position is the time average of the variance in these\n observables along the circumferential direction; the variance in\n time of the $\\phi$-average of these quantities is indeed very\n small. The picture that emerges is that the expression boundary is\n rough, even when the diffusion constant $D$ is large,\n i.e. $D=1~\\unit{\\mu m^2\/s}$. An analysis of the spatial correlation function\n at midembryo $\\TPAvg{\\delta H(0)\\delta H(\\phi)}(x_t)$, where $\\delta\n H(\\phi)=H(x_t,\\phi,t)-\\TPAvg{H}$, revealed that the correlation\n length $\\xi_\\phi$ is on the order of a few nuclei, which corresponds\n to the diffusion length $\\lambda=\\sqrt{D\/\\mu_{eff}}$ a protein can\n diffuse with diffusion constant $D$ before it is degraded with a\n rate $\\mu_{eff}$; the correlation length is thus small compared to\n the circumference. One possible source of coherent\n fluctuations in the mean copy number $\\Avg{X}_\\phi$ and boundary\n position $\\Avg{x_t}_\\phi$ are temporal variations of the morphogen\n profiles. However, in our model, these profiles are static---we\n argued that the morphogen fluctuations are fast on the timescale\n of gene expression, and are thus effectively integrated out. The\n small correlation length $\\xi_\\phi$ then indeed means that the\n varations in the mean over $\\phi$, $\\Avg{\\dots}_\\phi$, will be\n small. This leads to an interesting implication for\n experiments, which we discuss in the Discussion section.\n\n\n \\mysubsubsection{The boundary steepness} Now that we have\n characterized the fluctuations in the copy number and the boundary\n position, the next question is how fluctuations in the copy number\n affect the steepness of the boundary. In particular, a\n gene-expression boundary can be shallow either because at each\n moment in time the interface is shallow, or because at each moment\n in time the interface is sharp yet the interface fluctuates in\n time, leading to a smooth profile. The question is thus how much\n the gradient of the mean concentration profile,\n $\\TPAvg{G}^\\prime$, and the mean of the gradient,\n $\\TPAvg{G^\\prime}$, differ (here the prime denotes the spatial\n derivative). Fig.\\xspace S3 in Text S1\\xspace shows both quantities as a function of the\n diffusion constant. It is seen that while the average of the\n gradient is larger than the gradient of the average (as it\n should), the difference is around a factor of 2. We thus conclude that\n the steepness of the expression boundary at each moment in time\n does not differ very much from the steepness of the average\n concentration profile.\n\nIn the rest of the manuscript, we will predominantly focus on the\nproperties of individual embryos, and average quantities are\ntypically averages over time and the circumference. For brevity, therefore, \n$\\Avg{\\dots}=\\TPAvg{\\dots}$, unless stated otherwise.\n\\subsection*{Robustness to intra-embryonic fluctuations: Mutual repression allows for steeper profiles without raising the noise level at the boundary}\n\\pdfbookmark[2]{Robustness to intra-embryonic fluctuations: Mutual repression allows for steeper profiles without raising the noise level at the boundary}{BookmarkRobustnessIntra}\n\n\\mysubsubsection{Mutual repression shifts boundaries apart}\nFig.\\xspace \\ref{Fig3}A shows the average Hb\\xspace and Kni\\xspace steady-state\nprofiles along the anterior-posterior (AP) axis as a function of their\ndiffusion constant $D$ for a system with mutual repression. The inset\nshows the morphogen-activation profiles, which are the spatial profiles of\nthe probability that the \\textit{hb}\\xspace and \\textit{kni}\\xspace promoters have 5 copies of their\nrespective morphogens bound. Without mutual repression, thus when Hb\\xspace\nand Kni\\xspace cannot bind to their respective target promoters, these profiles\ndescribe the probability that \\textit{hb}\\xspace and \\textit{kni}\\xspace are activated by their\nrespective morphogens. Indeed, without mutual repression and without\nHb\\xspace and Kni\\xspace diffusion, the Hb\\xspace and Kni\\xspace concentration profiles would\nbe proportional to their respective morphogen-activation profiles\n\\cite{Erdmann2009}, which means that they would precisely intersect at\nmidembryo. In contrast, Fig.\\xspace \\ref{Fig3}A shows that the Hb\\xspace and Kni\\xspace\nconcentration profiles are shifted apart in the system with mutual\nrepression. There is already a finite separation for $D=0$, which\nincreases further as $D$ is increased.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=6in]{Figure3.pdf}\n\\end{center}\n\\caption{ {\\bf The effect of mutual repression on the average protein\n concentrations and their standard deviations.} \\subfig{A} Time-\n and circumference-averaged Hb\\xspace ($\\Avg{H}$, solid lines) and Kni\\xspace\n ($\\Avg{K}$, dashed lines) total protein copy number profiles along\n the AP axis for various diffusion constants $D$ in a system with\n mutual repression. The inset shows for both the \\textit{hb}\\xspace and the \\textit{kni}\\xspace\n promoter the probability that the promoter binds 5 morphogen\n proteins irrespective of whether the antagonistic gap protein is\n bound to it (meaning that the promoter is activated by the\n morphogen, even though it may be repressed by the antogonistic gap\n protein); these ``morphogen-activation'' profiles are identical for\n all $D$ values. \\subfig{B} Profiles of the probability\n $\\Avg{H^0_5}$ that the \\textit{hb}\\xspace promoter is induced, meaning that it has\n 5 copies of Bcd\\xspace bound to it and no Kni\\xspace dimer (solid lines), and the\n probability $\\Avg{H^1_5}$ that \\textit{hb}\\xspace is activated by Bcd\\xspace yet\n repressed by Kni\\xspace, in which case \\textit{hb}\\xspace is indeed not expressed (dashed\n lines). \\subfig{C} AP profiles of the time- and\n circumference-averaged standard deviation of the total gap protein\n copy number for Hb\\xspace ($\\sigma_H$, solid lines) and Kni\\xspace ($\\sigma_K$,\n dashed lines). \\subfig{D} Normalized standard deviation\n $\\sigma_{H}(x)\/\\Avg{H}_{max}$ versus the normalized mean\n $\\Avg{H}(x)\/\\Avg{H}_{max}$; $\\Avg{H}(x)$ is the averaged total\n Hb\\xspace copy number at $x$ and $\\Avg{H}_{max}$ is the maximum of this\n average over all $x$. The grey dashed line represents the\n Poissonian limit (PL) given by\n $\\sqrt{(1+f_D)\\Avg{H}(x)}\/\\Avg{H}_{max}$, where $f_D$ is the\n fraction of proteins in dimers. }\n\\label{Fig3}\n\\end{figure}\n\nIn Fig.\\xspace \\ref{Fig3}B we show the profile of the probability\n$\\Avg{H^0_5}$ that the \\textit{hb}\\xspace promoter is induced, meaning that it has 5\ncopies of Bcd\\xspace bound to it and no Kni\\xspace, and the profile of the\nlikelihood $\\Avg{H^1_5}$ that \\textit{hb}\\xspace is activated by Bcd\\xspace, yet repressed\nby Kni\\xspace, in which case \\textit{hb}\\xspace is not expressed. It is seen that\nrepression by \\textit{kni}\\xspace almost fully inhibits \\textit{hb}\\xspace expression beyond the\nhalf-activation point, where \\textit{hb}\\xspace would be expressed without \\textit{kni}\\xspace\nrepression (see inset Panel A). Indeed, mutual repression effectively\ncuts off protein production beyond midembryo. The production\nprobability therefore changes more abruptly along the AP axis, leading\nto a higher steepness of the protein profiles near midembryo. For\n$D>0$, repressor influx over the midplane increases, and as a result\nthe regions of expression inhibiton are enlarged and the concentration\nprofiles shift apart further.\n\n\\mysubsubsection{Noise reduction via spatial averaging}\nFig.\\xspace \\ref{Fig3}C shows the standard deviation of the protein copy\nnumber along the AP axis for both Hb\\xspace ($\\sigma_H$) and Kni\\xspace\n($\\sigma_K$). It is seen that the noise increases close to the\nhalf-activation point where promoter-state fluctuations are strongest\n\\cite{Tkacik2008,VanZon2006,So2011}. This is also observed in Fig.\\xspace\n\\ref{Fig3}D, which shows the normalized standard deviation $\\sigma_H \/\n\\Avg{H}_{max}$ versus the normalized mean $\\Avg{H}\/\\Avg{H}_{max}$\nof the average Hb\\xspace copy number; here, $\\Avg{H}_{max}$ is the\nmaximum average concentration of Hb\\xspace. The noise maximum close to mid\nembryo diminishes with increasing $D$, approaching the\nPoissonian limit. Note that the Poissonian limit here is given by\n$\\sigma_P=\\sqrt{(1+f_D)\\Avg{H}}$, where $f_D=2\\Avg{H_D}\/\\Avg{H}$ is\nthe fraction of dimerized Hb\\xspace proteins with respect to the total Hb\\xspace\ncopy number (see Text S1\\xspace for details). Clearly, the\nspatial averaging mechanism described in\n\\cite{Erdmann2009,Okabe-Oho2009} reduces the noise also in our system,\nwhich differs from those in \\cite{Erdmann2009,Okabe-Oho2009}\nby the presence of both gap gene monomers and dimers instead\nof monomers only.\n\n\\mysubsubsection{Mutual repression reduces the boundary width by\n increasing the steepness}\nFig.\\xspace \\ref{Fig4} quantifies the impact of spatial averaging and\nmutual repression on the Hb\\xspace boundary width $\\Delta x$, comparing it\nto that of the system without mutual repression. To first order, the\nboundary precision $\\Delta x$ is related to the standard\ndeviation in the protein copy number at the boundary, $\\sigma_H(x_t)$,\nand the steepness of the boundary, $|\\Avg{H(x_t)}^\\prime|$, via\nEq. \\ref{EqDeltaX} \\cite{Tostevin2007,Gregor2007,Erdmann2009}. The\nnoise $\\sigma_H(x_t)$ decreases with increasing $D$ due to spatial\naveraging in an almost identical manner for the systems with and\nwithout mutual repression (Fig.\\xspace \\ref{Fig4}, top panel); indeed,\nperhaps surprisingly, mutual repression has little effect on the noise\nat the boundary. Increasing $D$ also lessens the steepness of the\nprotein profiles, thus reducing the slope $|\\Avg{H(x_t)}^\\prime|$\n(Fig.\\xspace \\ref{Fig4}, middle panel). While without mutual repression this\nreduction is monotonic, in the case with mutual repression the\nsteepness first rises because increasing $D$ increases the influx of\nthe antagonistic repressor into the regions where the gap genes are\nactivated by their respective morphogens, which, for low values of\n$D$, {\\em steepens} the effective gene-activation profile\n$\\Avg{H^1_5}(x)$ by most strongly reducing gene expression near\nmidembryo; after the steepness has reached its maximum at\n$D=0.032~\\unit{\\mu m^2\/s}$, it drops for higher diffusion constants, because\nthe diffusion of the gap-gene proteins now flattens their\nconcentration profiles. Most importantly, with mutual repression\n$|\\Avg{H(x_t)}^\\prime|$ reaches significantly higher values for all\n$D\\leq 1.0~\\unit{\\mu m^2\/s}$. At $D=0.032~\\unit{\\mu m^2\/s}$ the profile is roughly twice\nas steep as in the case without repression. Interestingly, for\n$D\\lesssim 0.1~\\unit{\\mu m^2\/s}$, our simulation results for the steepness of\nthe profiles as normalized by their maximal values agree with those\nmeasured experimentally by Surkova {\\em et al.} in cell cycle 14\n\\cite{Surkova2008}: In both simulation and experiment, the\nconcentration drops from 90\\% to 10\\% of the maximal values over\n5-10\\% of the embryo length.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=3.27in]{Figure4.pdf}\n\\end{center}\n\\caption{ {\\bf The effect of mutual repression on the precision and\n steepness of the Hb\\xspace boundary.} The figure shows the\n time- and circumference-average of the standard deviation of the\n total Hb\\xspace copy number at the boundary $\\sigma_H(x_t)$ (upper panel),\n the slope of the total Hb\\xspace copy number profile at the boundary\n $|\\Avg{H}'(x_t)|$ (middle panel) and the Hb\\xspace boundary width $\\Delta\n x$ (lower panel) as a function of the diffusion constant $D$ of the\n gap proteins. Red solid lines show the case without (NR) and green\n solid lines the case with mutual repression (R); the red and green\n dashed lines show the limiting values without diffusion of the gap\n proteins. The grey dashed lines in the boundary width plot are the\n values based on the approximation $\\Delta\n x=\\sigma_H(x_t)\/|\\Avg{H(x_t)}'|$. Note that for $D<3.2~\\unit{\\mu m^2\/s}$,\n mutual repression enhances the steepness of the boundary, which in\n turn enhances the precision of the boundary.\n The black dotted line marks the $D$-value\n where the boundary is both steep and precise due to mutual repression.}\n\\label{Fig4}\n\\end{figure}\n\nBoth with and without Hb\\xspace-Kni\\xspace mutual repression the trade-off between\nnoise and steepness reduction leads to an optimal diffusion constant\n$D_{min}$ that maximizes boundary precision, i.e. minimizes $\\Delta x$\n(Fig.\\xspace \\ref{Fig4}, lower panel). Mutual repression enhances the\nprecision for $D\\leq 1.0~\\unit{\\mu m^2\/s}$ because in this regime decreasing $D$\nincreases the steepness markedly while it has only little effect on\nthe noise as compared to the system without mutual repression.\nConversely, $\\Delta x$ is increased by mutual repression for $D\\geq\n10~\\unit{\\mu m^2\/s}$ because it reduces the steepness. The minimum in the case\nwith repression is marginally lower than that without\n($D_{min,R}\/D_{min,NR}\\simeq0.86$), but located at a lower $D$-value\n($1.0~\\unit{\\mu m^2\/s}$ vs. $3.2~\\unit{\\mu m^2\/s}$). Most importantly, at\n$D=0.32~\\unit{\\mu m^2\/s}$, the system with mutual repression produces a profile\nthat is twice as steep as that of the system without it at\n$D_{min,NR}=3.2~\\unit{\\mu m^2\/s}$, whereas the precision $\\Delta x$ is\nessentially the same in both cases. Clearly, mutual repression can\nstrongly enhance the steepness of gene-expression boundaries without\ncompromising their precision.\n\n\\mysubsubsection{Influence of Hill coefficient}\nA key parameter controlling the precision of the gap-gene\nexpression boundaries, is the degree of cooperativity by which the\ngap genes are activated by their respective morphogens---this\ndetermines the profile steepness of the average gap-gene\npromoter activity. To investigate this, we have lowered the\neffective Hill coefficient from its baseline value of 5 by reducing\nthe number $n_{\\rm max}$ of morphogen molecules that are required to\nbind the promoter to activate gene expression. To isolate the effect\nof varying the {\\em mean} gene-activation profiles $\\Avg{H_{n_{\\rm\nmax}}^0}(x)$ and $\\Avg{K_{n_{\\rm max}}^0}(x)$, we varied, upon\nvarying $n_{\\rm max}$, the association and dissociation rates such\nthat 1) the average gene activation probabilities near midembryo,\n$\\Avg{H_{n_{\\rm max}}^0}(L\/2)$ and $\\Avg{K_{n_{\\rm max}}^0}(L\/2)$, are\nunchanged and 2) the waiting-time distribution for the gene\non-to-off transition is unchanged (since the average activation\nprobability is fixed, the mean off-to-on rate is also unchanged, although\nthe waiting-time distribution is not; see also Fig.\\xspace S5 in Text S1\\xspace).\nWe observe that mutual repression\nmarkedly enhances the steepness of the gap-gene expression\nboundaries, also with a\nlower Hill coefficient for gene activation (Fig.\\xspace S6 in Text S1\\xspace). However, lowering the Hill coefficient\nreduces the steepness of the gene-activation profiles, causing the\ntwo antagonistic gene-activation profiles to overlap more. As a\nresult, in each of the two gap-gene expression domains, more of the\nantagonist is present, which tends to increase the noise in gene\nexpression by occassionally shutting off gene production. This, as\nexplained in more detail later, is\nparticularly detrimental when the diffusion constant is low. Indeed,\nwhen the effective Hill\ncoefficient of gene activation is 3 or lower, mutual repression {\\em\nincreases} $\\Delta x$ when the diffusion constant is low,\ni.e. below approximately $0.1~\\unit{\\mu m^2\/s}$. Nonetheless, the {\\em minimal} $\\Delta\nx$ is still lower with mutual repression, and, consequently, also\nwith a lower Hill coefficient for gene activation, mutual repression\ncan enhance both the steepness and the precision of gene-expression\nboundaries.\n\n\\mysubsubsection{Influence of the repression strength}\nAs a standard we assume very tight binding of the Hb\\xspace and Kni\\xspace dimers,\n``the repressors'', to their respective promoters. To test how this\nassumption affects our results we performed simulations in which we\nsystematically varied the repressor-promoter dissociation rate $k^{\\rm\n R}_{off}$ in the range $[5.27\\cdot 10^{-4}\\unit{\/s},5.27\\cdot\n10^{2}\\unit{\/s}]$, keeping the diffusion constant at $D=1.0~\\unit{\\mu m^2\/s}$\n(the value that minimizes the boundary width at $k^{\\rm\n R}_{off}=5.27\\cdot 10^{-3}\\unit{\/s}$) and all other parameters the\nsame as before. Fig.\\xspace \\ref{Fig5} shows the noise, steepness and\nboundary precision as a function of the repressor-promoter\ndissociation rate. For high dissociation rates, these quantities equal\nthose in the system without mutual repression (dashed lines). Yet, as\nthe dissociation rate is decreased, the steepness rises markedly at\n$k^{\\rm R}_{off}=1\/{\\rm s}$. In contrast, the noise $\\sigma_H(x_t)$\nfirst decreases with decreasing $k^{\\rm R}_{off}$, passing through\na minimum at $k^{\\rm R}_{off}=0.1\/{\\rm s}$ before rising to a level that\nis higher than that in a system without mutual repression. This\nminimum arises because on the one hand increasing the affinity of\nthe repressor (the antagonist) makes the operator-state fluctuations\nof the activator (the morphogen) less important---increasing\nrepressor binding drives the concentration profiles of Hb\\xspace and Kni\\xspace\naway from midembryo, where the promoter-state fluctuations of the\nactivators are strongest; on the other hand, when the repressor\nbinds too strongly, then slow repressor unbinding leads to\nlong-lived promoter states where gene expression is shut off,\nincreasing noise in gene expression; this phenomenon is similar to\nwhat has been observed in Refs. \\cite{VanZon2006} and\n\\cite{Morelli2008}, where slower binding of the gene regulatory\nproteins to the promoter increases noise in gene expression and\ndecreases the stability of a toggle switch, respectively. The interplay between the\nnoise and the steepness yields a marked reduction of the boundary\nwidth $\\Delta x$; indeed, even in the limit of very tight repressor binding,\nmutual repression significantly enhances the precision of the\nboundary.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=3.27in]{Figure5.pdf}\n\\end{center}\n\\caption{ {\\bf The effect of varying repression strength on the precision\n and steepness of the Hb\\xspace boundary.} Shown are the time- and\n circumference average of the standard deviation of the total Hb\\xspace copy\n number at the boundary $\\sigma_H(x_t)$ (upper panel), the steepness\n of the boundary $|\\Avg{H}'(x_t)|$ (middle panel) and the Hb\\xspace\n boundary width $\\Delta x$ (lower panel) as a function of\n $k_{\\rm off}^{\\rm R}$, the promoter-dissociation rate of Hb\\xspace and Kni\\xspace. The solid green line are values obtained from the\n boundary position distribution, the dashed grey line the ones\n calculated from the approximation $\\Delta\n x=\\sigma_H(x_t)\/|\\Avg{H(x_t)}'|$. Straight dashed lines mark\n the limits for the case without mutual repression ($k_{on}^{\\rm R}=k_{off}^{\\rm R}=0$).}\n\\label{Fig5}\n\\end{figure}\n\n\n\\mysubsubsection{Influence of expression level}\nSince the precise gap protein expression level is not known, we also varied the\nmaximal protein copy number $N$ by varying the maximal expression\nrate $\\beta$ (see Text S1\\xspace). Fig.\\xspace S9 in Text S1\\xspace shows the output\nnoise and slope at the boundary position, and the boundary precision $\\Delta x$, \nas a function of the diffusion constant for three different\nexpression levels. It is seen that for low diffusion constant, the\nprecision is independent of $N$, while for higher diffusion constant it scales roughly with\n$1\/\\sqrt{N}$. This can be understood by noting that the steepness of\nthe gene-expression boundary scales to a good approximation\nwith $N$ independently of $D$, while the noise $\\sigma$ scales with\n$N$ when the diffusion constant is small, but with $\\sqrt{N}$ when\nthe diffusion constant is large (see also Eq. \\ref{EqDeltaX}). The\nscaling of the noise with $N$ is due to the fact that for low $D$\nthe noise in the copy number is dominated by the noise coming from\nthe promoter-state fluctuations, which scales linearly with $N$,\nwhile for high $D$, diffusion washes out the expression bursts resulting from\nthe promoter-state flucutations, leaving only the noise coming from\nthe Poissonian fluctuations arising from transcription and\ntranslation, which scales with the square root of $N$\n\\cite{Erdmann2009}. In Text S1\\xspace we also study the importance of\nbursts arising in the transcription-translation step (see Fig.\\xspace S8 in Text S1\\xspace); however, we\nfind that for a typical burst size, these bursts do not dramatically\naffect boundary precision.\n\n\\FloatBarrier\n\n\\subsection*{Robustness to inter-embryonic variations: Mutual repression can buffer against correlated morphogen level variations}\n\\pdfbookmark[2]{Robustness to inter-embryonic variations: Mutual repression can buffer against correlated morphogen level variations}{BookmarkRobustnessInter}\nAlthough the Bcd\\xspace copy number at midembryo has been determined\nexperimentally \\cite{Gregor2007}, the measured value is not\nnecessarily the half-activation threshold of \\textit{hb}\\xspace. Indeed, in vivo the\nHb\\xspace profile is shaped by other forces, like mutual repression. In the\n\\textit{kni}\\xspace-\\textit{kr}\\xspace double mutant, the Hb boundary at midembryo shifts\nposteriorly \\cite{Manu2009PlosBiol}. Moreover, gap gene domain\nformation has been observed at strongly reduced Bcd\\xspace levels,\nsuggesting that Bcd\\xspace might be present in excess\n\\cite{Ochoa-Espinosa2009}. Also from a theoretical point of view it\nis not obvious that a precisely centered morphogen-activation\nthreshold is optimal, in terms of robustness against both\nintra-embryonic fluctuations and inter-embryonic variations. Here, we\nstudy the effect of changing the threshold position where \\textit{hb}\\xspace and \\textit{kni}\\xspace\nare half-maximally activated by their respective morphogens, Bcd\\xspace and\nCad\\xspace. While the threshold positions could be varied by changing the\nthreshold morphogen concentrations for half-maximal gap-gene\nactivation (for example by changing the morphogen-promoter\ndissociation rates), we will vary these positions by changing the\namplitude of the morphogen profiles by a factor $A$. This procedure\nnot only preserves the promoter-activation dynamics at the\nboundaries---a key determinant for the noise at the boundaries---but\nalso allows us to study the importance of mutual repression in\nensuring robustness against embryo-to-embryo variations. Indeed, we\nwill examine not only how changing the threshold position affects the\nprecision of the gap-gene expression boundaries, $\\Delta x(A)$,\nbut also how the average boundary positions vary with morphogen\ndosage, $x_t(A)$, and how the latter gives rise to embryo-to-embryo\nvariations in the boundary position $\\Delta x_t(\\Delta A)$ due to\nembryo-to-embryo variations in the morphogen dosage $\\Delta A$.\n\n\\mysubsubsection{Double-activation induces bistability}\nWe first consider the scenario in which the amplitudes of both\nmorphogens are scaled by the same factor $A$. When $A=1$, the\nposition at which \\textit{hb}\\xspace and \\textit{kni}\\xspace are half-maximally activated by their\nrespective morphogens coincide at midembryo, meaning that the\ndomains in which \\textit{hb}\\xspace and \\textit{kni}\\xspace are activated beyond half-maximum are\nadjoining, but do not overlap---this is the scenario discussed in\nthe previous sections. When $A>1$, the position at which \\textit{hb}\\xspace is\nhalf-maximally activated by its morphogen is shifted posteriorly,\nwhile that of \\textit{kni}\\xspace is shifted anteriorly, creating an overlap between\nthe two regions where \\textit{hb}\\xspace and \\textit{kni}\\xspace are activated.\nIn this ``double-activated region'' both \\textit{hb}\\xspace and \\textit{kni}\\xspace are\nactivated by their respective morphogens, yet they also mutually\nrepress each other. This may lead to bistability. To probe whether\nthis is the case, we performed a bifurcation analysis of the\nmean-field chemical-rate equations of isolated nuclei, implying that\n$D=0$ (see Fig.\\xspace S1 in Text S1\\xspace). In addition, we performed stochastic\nsimulations of isolated nuclei with different morphogen levels\ncorresponding to different positions along the AP axis. All other\nparameter values were the same as in the full-scale simulation. We\nrecorded long trajectories of the order parameter $\\Delta N \\equiv\nH-K$, the difference between the total Hb\\xspace and total Kni\\xspace copy numbers,\nin the stationary state. From each trajectory we computed the\ndistribution $P(\\Delta N)$ of the probability that the system is in a\nstate with copy number difference $\\Delta N$. This defines a ``free\nenergy'' $G(\\Delta N)\\equiv -\\ln P(\\Delta N)$, with minima of\n$G(\\Delta N)$ corresponding to maximally probable values of $\\Delta\nN$ \\cite{Warren2004,Warren2005}. For a bistable system, $G(\\Delta\nN)$ resembles a double-well potential with minima located at a\npositive value of $\\Delta N=\\Delta N_{\\rm H}$ and a negative value of\n$\\Delta N=\\Delta N_{\\rm K}$, respectively. At midembryo the morphogen\nlevels of Bcd\\xspace and Cad\\xspace are the same and hence the biochemical network\nin the nuclei in the midplane is symmetric, which means that, if this\nnetwork is bistable, $G(\\Delta N)$ resembles a symmetric double-well\npotential with $\\Delta N_H=-\\Delta N_K$ and $\\Delta G \\equiv G(\\Delta\nN_H) - G(\\Delta N_K)=0$. Away from the middle, the morphogen levels\ndiffer, and one state will become more stable than the other; if the\nother state is, however, still metastable, then $G(\\Delta N)$ will\nresemble an asymmetric double-well potential, with $\\Delta G$ being\nnegative if the \\textit{hb}\\xspace-dominant state is more stable than the\n\\textit{kni}\\xspace-dominant state, and vice versa. The emergence of such a\n``spatial switch'' along the AP axis is also captured by our\nmean-field, bifurcation analysis (see Text S1\\xspace) and was recently\nalso shown in the mean-field analysis of Papatsenko and Levine for the same pair of\nmutually repressing genes\\cite{Papatsenko2011}.\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=3.27in]{Figure6.pdf}\n\\end{center}\n\\caption{ {\\bf Emergence of bistability in double-activated regions.}\n The ``free energy'' difference $\\Delta G\\equiv G(\\Delta N_H) -\n G(\\Delta N_K)$ as a function of $x$, the distance of the nucleus\n from the anterior pole, for different amplitudes of the morphogen\n gradients $A$; here, $G(\\Delta N)\\equiv -\\ln(P(\\Delta N))$, where\n $P(\\Delta N)$ is the stationary distribution of the order parameter\n $\\Delta N=H-K$; $\\Delta N_H\\simeq -\\Delta N_K \\approx 800$ correspond to\n the minima of $G(\\Delta N)$. Negative values of $\\Delta G$\n represent a strong bias towards the high-Hb\\xspace state, while positive\n values correspond to high-Kni\\xspace states. The insets shows $G(\\Delta\n N)$ as a function of $\\Delta N$ at the positions indicated by the\n numbers in their corners (values in [\\unit{\\% EL}]; colors correspond to\n main plot). The data is obtained from simulations of single nuclei\n with morphogen levels corresponding to the ones at position $x$ in\n the full system; this is equivalent to the full system without\n diffusion between neighboring nuclei. Note the bistable behavior in\n a wide region of the embryo for higher $A$ values. }\n\\label{Fig6}\n\\end{figure}\n\nFig.\\xspace \\ref{Fig6} shows $\\Delta G$ as a function of the position along\nthe AP axis, for different amplitudes $A$ of the morphogen gradients. The\ninset shows the energy profiles $G(\\Delta N)$ for different positions\nalong the AP axis. For $A=1$, $G(\\Delta N)$ always exhibits one\nminimum only, irrespective of the position along the AP axis; at midembryo,\nthis minimum is located at $\\Delta N=0$, while moving towards\nthe anterior (posterior) the energy minimum rapidly shifts to $\\Delta\nN \\approx +800 (-800)$, reflecting that in the anterior (posterior) half of\nthe embryo \\textit{hb}\\xspace (\\textit{kni}\\xspace) is essentially fully expressed. For $A=2$,\n$G(\\Delta N)$ develops into a double-well potential at midembryo,\nwith two pronounced minima at $\\Delta N\\approx 800$ and $\\Delta\nN\\approx-800$, respectively. These two minima correspond to a state in\nwhich \\textit{hb}\\xspace is highly expressed ($\\Avg{H}\\approx 800$) and \\textit{kni}\\xspace is\nstrongly repressed ($\\Avg{K}\\approx 0$) and another state in which\n\\textit{kni}\\xspace is highly expressed and \\textit{hb}\\xspace strongly repressed, respectively. The\nfact that the two energy mimima are equal indicates that both of these\nstates are equally likely. Moving away from midembryo, however, one\ngap-gene expression state rapidly becomes more stable than the other,\nand bistability is lost, yielding a potential with one minimum located\nat $\\Delta N\\approx 800$ in the anterior half and a potential with one minimum located\nat $\\Delta N\\approx -800$ in the posterior half of the embryo.\nInterestingly, for $A=4$\nand $A=8$ a wide region of bistability develops\naround midembryo. In this region, $\\Delta G \\approx 0$, meaning that \nthe high-\\textit{hb}\\xspace---low-\\textit{kni}\\xspace state and the low-\\textit{hb}\\xspace---high-\\textit{kni}\\xspace state are\nequally stable. These two states are equally likely because in this\nregion both the \\textit{hb}\\xspace and \\textit{kni}\\xspace promoters are fully activated by their\nrespective morphogens. It can also be seen that the width of this\nbistable region increases with the amplitude of the morphogen\ngradients, as expected.\n\\pagebreak\n\n\\mysubsubsection{Slow switching ensures a low noise level while diffusion avoids error locking}\nThe bistability observed for $A>1$ and $D=0$ raises an important\nquestion, namely whether the nuclei can switch between the two\ngap-gene expression states on the time scale of embryonic\ndevelopment. This question is particularly pertinent for the higher\nmorphogen amplitudes, where these two states are equally likely ($\\Delta\nG\\approx 0$) over a wide region of the embryo (Fig.\\xspace \\ref{Fig6}): random\nswitching between the two distinct gap-gene expression states in this\nwide region would then lead to dramatic fluctuations in the positions\nof the \\textit{hb}\\xspace and \\textit{kni}\\xspace expression boundaries, which clearly would be\ndetrimental for development. We therefore computed \\cite{Warren2005} from the recorded\nswitching trajectories the average waiting time for switching,\n$\\tau_{s}$, at midembryo ($\\Delta\nG\\simeq 0$) for different values of $A$; for $A\\geq 2$, we find\n$\\tau_{s} \\simeq\n6~\\unit{h}$ (see Table S1 in Text S1\\xspace). During cell cycle 14, approximately 2-3 hours\nafter fertilization, the Bcd\\xspace gradient disappears \\cite{Drocco2011}, suggesting that the\nspontaneous switching rate is indeed low on the relevant time scale of\ndevelopment.\n\nWith diffusion of Hb\\xspace and Kni\\xspace between neighboring nuclei ($D>0$), the\ntime scale for switching will be even longer. Diffusion couples\nneighboring nuclei, creating larger spatial domains with the same\ngap-gene expression state. This reduces the probability that a nucleus\nin the overlap region flips to the other gap-gene expression\nstate. The latter can be understood from the extensive studies on the\nswitching behavior of the ``general toggle switch''\n\\cite{Warren2004,Warren2005,Allen2005,Lipshtat2006,Loinger2007,Morelli2008}, which is highly\nsimilar to the system studied here---indeed, the toggle switch\nconsists of two genes that mutually repress each other. These studies\nhave revealed that the ensemble of transition states, which separate\nthe two stable states, is dominated by configurations where both\nantagonistic proteins are present in low copy numbers. Clearly, the\nprobability that in a given nucleus not only the minority gap protein,\nbut also the majority gap protein reaches a low copy number, is\nreduced by the diffusive influx of that majority species from the\nneighboring nuclei, which are in the same gap-gene expression\nstate. In essence, diffusion increases the effective system size, with\nits spatial dimension given by $\\lambda=\\sqrt{D\/\\mu_{\\rm eff}}$; in\nfact, since the stability of the toggle switch depends exponentially\non the system size \\cite{Warren2004,Warren2005}, we expect the stability $\\tau_{s}$\nto scale with the diffusion constant as $\\tau_{s}\\sim e^D$. We thus\nconclude that random switching between the two gap-gene expression\nstates, the high-\\textit{hb}\\xspace---low-\\textit{kni}\\xspace and low-\\textit{hb}\\xspace---high-\\textit{kni}\\xspace states, is not\nlikely to occur on the time scale of early development.\n\nThe observation that the switching rate is low raises another\nimportant question: if errors are formed during development, can they\nbe corrected? We observe in the simulations with $D=0$ that when we\nallow the gap-gene expression patterns to develop starting from\ninitial conditions in which the Hb\\xspace and Kni\\xspace copy numbers are both\nzero, in the overlap (bistable) region a spotty gap-gene expression\npattern emerges, consisting of nuclei that are either in the\nhigh-\\textit{hb}\\xspace---low-\\textit{kni}\\xspace state or in the low-\\textit{hb}\\xspace---high-\\textit{kni}\\xspace state. When\nthe diffusion constant of Hb\\xspace and Kni\\xspace is zero, then these defects are\nessentially frozen in, precisely because of the low switching\nrate. Interestingly, however, we find in the simulations that a finite\ndiffusion constant {\\em can} anneal these defects. This may seem to\ncontradict the statement made above that diffusion lowers the\nswitching rate. The resolution of this paradox is that while diffusion\nlowers the switching rate for nuclei that are surrounded by nuclei\nthat are in the same gap-gene expression state, it enhances the\nswitching rate for nuclei that are surrounded by nuclei with a\ndifferent gap-gene expression state; this is indeed akin to spins in\nan Ising system below the critical point. The mechanism for the\nformation of the gap-gene expression patterns, then, depends on the\ndiffusion constant. When $D$ is small yet finite, $00$ and $A\\leq 4$, \ndue to the diffusive influx of Hb\\xspace and Kni\\xspace from the regions outside\nthe overlap region. When $A=8$, the \\textit{hb}\\xspace and \\textit{kni}\\xspace expression boundaries\nare not pinned to the middle of the embryo, and their positions\nexhibit slow and large fluctuations, presumably because the energetic driving\nforce is small, and the diffusive influx of Hb\\xspace and Kni\\xspace from the\nregions near the poles is negligible. We will investigate this effect in\nmore detail in a forthcoming publication. \n\n\n\\mysubsubsection{Mutual repression inhibits boundary shifts}\nFig.\\xspace \\ref{Fig7}A shows the average gap-gene expression profiles for\n$A\\in\\lbrace 1,2,4\\rbrace$ and $D=1.0~\\unit{\\mu m^2\/s}$, which minimizes the\nboundary width $\\Delta x$ when $A=1$ (see Fig.\\xspace \\ref{Fig4}). While\nthe morphogen-activation thresholds shift beyond midembryo as $A$ is\nincreased beyond unity, leading to an overlap of the domains where the\ngap genes are activated by their respective morphogens (see inset),\nthe gap-gene expression boundaries overlap only marginally. This is\nquantified in panel B, which shows the Hb\\xspace boundary position $x_t$ as\na function of $A$ and as a function of $\\Delta x_A\\equiv\nx_{A,Kni}-x_{A,Hb}$, which is defined as the separation between the\npositions $x_{A,Kni}$ and $x_{A,Hb}$ where Kni\\xspace and Hb\\xspace are\nhalf-maximally activated by their respective morphogens; for $A=1$,\nwith adjoining morphogen activation regions, $\\Delta x_A=0$ and for\n$A>1$, with overlapping activation regions, $\\Delta x_A$ is\nnegative. Without mutual repression (red data), the Hb\\xspace boundary\nposition $x_t$ tracks the shift of the \\textit{hb}\\xspace activation threshold, as\nexpected. In contrast, with mutual repression (green data) the\nboundary does not move beyond the position for $A=1$ as $A$ is\nincreased. The same robustness was also observed for other values\nof the Hill coefficient of gap-gene activation (see Fig.\\xspace S7 in Text S1\\xspace).\n\n\\begin{figure}[!ht]\n\\begin{center}\n\\includegraphics[width=6in]{Figure7.pdf}\n\\end{center}\n\\caption{ {\\bf Mutual repression buffers against correlated variations\n in the activator levels.} \\subfig{A} Time- and\n circumference-averaged Hb\\xspace ($\\Avg{H}$, solid lines) and Kni\\xspace\n ($\\Avg{K}$, dashed lines) total copy-number profiles along the AP\n axis for various morphogen dosage factors $A$. Inset: the\n corresponding average occupancy of the promoter states with five\n bound morphogen molecules as a function of $x$. \\subfig{B} The average \n Hb\\xspace boundary position $x_t$ as a function of $\\Delta x_A$, the distance\n between the Hb\\xspace and Kni\\xspace boundaries without mutual repression, for\n the system with mutual repression (green,) and without it (red);\n $\\Delta x_A$ is varied by changing the morphogen dosage factor\n $A$. Note that mutual repression makes the gap-gene expression\n boundaries essentially insensitive to correlated changes in\n morphogen levels when $A>1$. \\subfig{C} AP profiles of the average\n standard deviation of the total Hb\\xspace ($\\sigma_H$, solid lines) and\n Kni\\xspace ($\\sigma_K$, dashed lines) copy numbers. Inset:\n $\\sigma_{H}(x)\/\\Avg{H}_{max}$ as a function of\n $\\Avg{H}(x)\/\\Avg{H}_{max}$, where $\\Avg{H}(x)$ is the average Hb\\xspace\n copy number at $x$ and $\\Avg{H}_{max}$ its maximum over $x$. The\n grey dashed line represents the Poissonian limit.\\subfig{D} The Hb\\xspace\n boundary width $\\Delta x$ as a function of $\\Delta x_A$ with (green)\n and without (red) mutual repression. For $A=4$, it was impossible to\n obtain a reliable error bar on $\\Delta x$, because of the weak\n pinning force on the \\textit{hb}\\xspace and \\textit{kni}\\xspace expression boundaries. }\n\\label{Fig7}\n\\end{figure}\n\n\\mysubsubsection{Mutual repression enhances robustness to\n embryo-to-embryo variations}\nThe fact that mutual repression can pin expression boundaries,\ndramatically enhances the robustness against embryo-to-embryo\nvariations in the morphogen levels. We did not sample inter-embryo\nvariations in $A$ explicitly, but made an estimate using $\\Delta x_t =\n(dx_t\/dA) \\Delta A$, where $d x_t\/dA$ was taken from Fig.\\xspace \\ref{Fig7}B.\nA correlated symmetric variation $\\delta_A\\equiv\\Delta A\/A=0.1$ of\nboth morphogen levels then would lead to $\\Delta\nx_t(\\delta_A)\\simeq0.82~\\unit{\\% EL}$ at $A=1$ and $\\Delta\nx_t(\\delta_A)\\simeq0.25~\\unit{\\% EL}$ at $A=2$. Without mutual repression\n$\\Delta x_{t,NR}(\\delta_A)\\simeq2.2~\\unit{\\% EL}$. This analysis thus\nsuggests that mutual repression reduces boundary variations due to\nfluctuations in the morphogen levels by almost a factor of 10 if the\nhalf-activation threshold is slightly posterior to midembryo\n(e.g. $A=2$). If, on average, $A=1$, then mutual repression\nstill reduces $\\Delta x_t$ by inhibiting posterior shifts in those\nembryos in which $A>1$. These results are consistent with those of\n\\cite{Howard2005,Vakulenko2009}.\n\n\\mysubsubsection{Overlap of morphogen activation domains does not\n corrupt robustness to intrinsic fluctuations}\nWhile mutual repression proves beneficial in buffering against\nembryo-to-embryo variations in morphogen levels, the question arises\nwhether overlapping morphogen-activation domains does not impair\nrobustness to intrinsic fluctuations arising from noisy gene\nexpression and diffusion of gap gene proteins. We found that this\ndepends on the Hill coefficient of gap-gene\nactivation, which depends on the number $n_{\\rm max}$ of\nmorphogen binding sites on the promoter. Fig.\\xspace \\ref{Fig7}C shows, for $n_{\\rm max}=5$, that even though\nmutual repression increases the noise in gap-gene expression away\nfrom the boundaries, it has little effect on the noise at the boundaries\nwhen $A\\leq 2$. For $A> 2$, the noise does increase significantly;\nin fact, it was impossible to obtain reliable error bars, because of\nthe weak pinning force of the \\textit{hb}\\xspace-\\textit{kni}\\xspace interface. Moreover,\noverlapping morphogen activation domains decrease the steepness of\nthe expression boundaries (panel A), and this increases the boundary\nwidth $\\Delta x$ (panel D). Indeed, when $n_{\\rm max}=5$,\nmutual repression can enhance the precision of gene-expression\nboundaries, but only if the activation domains are adjoining\n($A=1$), or have a marginal overlap ($1 c_k$}\n\\end{cases}\n\\label{eq:huber}\n\\end{equation}\n\\noindent\nIn Eq.~\\eqref{eq:huber}, the tuning factor $c_k$ is estimated inherently from the data and is set to $c_k=1.345 \\times \\sigma_k$. The factor of $1.345$ is chosen to provide approximately 95\\% asymptotic efficiency and $\\sigma_k$ is a robust measure of bit-wise variance of $r_{ij}^k$. Specifically, $\\sigma_k $ is estimated as $1.485$ times the median absolute deviation of $r_{ij}^k$ as empirically suggested in~\\cite{robusthuber}. This robust formulation provides immunity to outliers during training by clipping their gradients. For training with the aux-SI hashing arm, we employ a similar robust retrieval loss $J_{\\mathcal{S}}^{\\text{SI}}$ defined over single instances with bag-labels assigned to member instances.\n\nTo minimize loss of retrieval quality due to quantization, we use a differentiable quantization loss $J_{Q} = \\sum_{i=1}^M(\\text{log}\\ \\text{cosh}(\\mid\\mathbf{h_i}\\mid - \\ \\mathbf{1}))$ proposed in~\\cite{zhuloss}. This loss also counters the effect of using continuous relaxation in definition of $p_{ij}$ over using Hamming distance. \nAs a standard practice in deep learning, we also add an additional weight decay regularization term $R_{W}$, which is the Frobenius norm of the weights and biases, to regularize the cost function and avoid over-fitting. \n\n\n\n\\begin{wrapfigure}[15]{r}[0pt]{0pt}\n\\resizebox{0.4\\textwidth}{!}{\n\\begin{tikzpicture}\n\\begin{axis}[\n\txlabel= \\Large{t (Epochs)},ylabel= {\\Large{$\\lambda_{\\text{MI\/SI}}^t$}},ylabel shift = -1 pt,legend style={at={(0.46,0.5)},anchor=west,draw=none}]\n\\addplot[smooth,blue,ultra thick] coordinates {\n(0,\t0.500000000000000)\n(1.00671140939597,\t0.506688887887933)\n(2.01342281879195,\t0.513332732759786)\n(3.02013422818792,\t0.519931534615558)\n(4.02684563758389,\t0.526485293455250)\n(5.03355704697987,\t0.532994009278861)\n(6.04026845637584,\t0.539457682086392)\n(7.04697986577181,\t0.545876311877843)\n(8.05369127516779,\t0.552249898653214)\n(9.06040268456376,\t0.558578442412504)\n(10.0671140939597,\t0.564861943155714)\n(11.0738255033557,\t0.571100400882843)\n(45.3020134228188,\t0.756407369037431)\n(46.3087248322148,\t0.761069321201748)\n(47.3154362416107,\t0.765686230349984)\n(48.3221476510067,\t0.770258096482140)\n(49.3288590604027,\t0.774784919598216)\n(50.3355704697987,\t0.779266699698212)\n(51.3422818791946,\t0.783703436782127)\n(52.3489932885906,\t0.788095130849962)\n(53.3557046979866,\t0.792441781901716)\n(54.3624161073826,\t0.796743389937390)\n(55.3691275167785,\t0.800999954956984)\n(56.3758389261745,\t0.805211476960497)\n(57.3825503355705,\t0.809377955947930)\n(58.3892617449664,\t0.813499391919283)\n(59.3959731543624,\t0.817575784874555)\n(60.4026845637584,\t0.821607134813747)\n(61.4093959731544,\t0.825593441736859)\n(62.4161073825503,\t0.829534705643890)\n(90.6040268456376,\t0.921602630512139)\n(91.6107382550336,\t0.924237646952840)\n(92.6174496644295,\t0.926827620377461)\n(93.6241610738255,\t0.929372550786001)\n(94.6308724832215,\t0.931872438178460)\n(95.6375838926175,\t0.934327282554840)\n(96.6442953020134,\t0.936737083915139)\n(97.6510067114094,\t0.939101842259358)\n(98.6577181208054,\t0.941421557587496)\n(99.6644295302014,\t0.943696229899554)\n(109.731543624161,\t0.963965587135715)\n(110.738255033557,\t0.965744786270889)\n(111.744966442953,\t0.967478942389982)\n(112.751677852349,\t0.969168055492996)\n(113.758389261745,\t0.970812125579929)\n(114.765100671141,\t0.972411152650782)\n(115.771812080537,\t0.973965136705554)\n(116.778523489933,\t0.975474077744246)\n(117.785234899329,\t0.976937975766857)\n(138.926174496644,\t0.997274897527138)\n(139.932885906040,\t0.997747849195982)\n(140.939597315436,\t0.998175757848746)\n(141.946308724832,\t0.998558623485429)\n(142.953020134228,\t0.998896446106031)\n(143.959731543624,\t0.999189225710554)\n(144.966442953020,\t0.999436962298996)\n(145.973154362416,\t0.999639655871357)\n(146.979865771812,\t0.999797306427638)\n(147.986577181208,\t0.999909913967839)\n(148.993288590604,\t0.999977478491960)\n(150,\t1)\n};\n\\addlegendentry{\\Large{MIL}}\n\n\\addplot[smooth,color=red,dashed,ultra thick]\n\tcoordinates {\n(0,\t0.500000000000000)\n(1.00671140939597,\t0.493311112112067)\n(2.01342281879195,\t0.486667267240214)\n(3.02013422818792,\t0.480068465384442)\n(4.02684563758389,\t0.473514706544750)\n(5.03355704697987,\t0.467005990721139)\n(6.04026845637584,\t0.460542317913608)\n(7.04697986577181,\t0.454123688122157)\n(8.05369127516779,\t0.447750101346786)\n(9.06040268456376,\t0.441421557587496)\n(10.0671140939597,\t0.435138056844286)\n(11.0738255033557,\t0.428899599117157)\n(12.0805369127517,\t0.422706184406108)\n(13.0872483221477,\t0.416557812711139)\n(14.0939597315436,\t0.410454484032251)\n(15.1006711409396,\t0.404396198369443)\n(16.1073825503356,\t0.398382955722715)\n(17.1140939597315,\t0.392414756092068)\n(18.1208053691275,\t0.386491599477501)\n(19.1275167785235,\t0.380613485879014)\n(20.1342281879195,\t0.374780415296608)\n(21.1409395973154,\t0.368992387730282)\n(36.2416107382550,\t0.287577136165038)\n(37.2483221476510,\t0.282509796855998)\n(38.2550335570470,\t0.277487500563038)\n(39.2617449664430,\t0.272510247286158)\n(40.2684563758389,\t0.267578037025359)\n(41.2751677852349,\t0.262690869780641)\n(42.2818791946309,\t0.257848745552002)\n(43.2885906040268,\t0.253051664339444)\n(44.2953020134228,\t0.248299626142967)\n(45.3020134228188,\t0.243592630962569)\n(46.3087248322148,\t0.238930678798252)\n(47.3154362416107,\t0.234313769650016)\n(48.3221476510067,\t0.229741903517860)\n(64.4295302013423,\t0.162717895590289)\n(65.4362416107383,\t0.158911760731499)\n(66.4429530201342,\t0.155150668888789)\n(67.4496644295302,\t0.151434620062159)\n(68.4563758389262,\t0.147763614251610)\n(69.4630872483222,\t0.144137651457142)\n(70.4697986577181,\t0.140556731678753)\n(71.4765100671141,\t0.137020854916445)\n(72.4832214765101,\t0.133530021170218)\n(73.4899328859060,\t0.130084230440070)\n(74.4966442953020,\t0.126683482726003)\n(75.5033557046980,\t0.123327778028017)\n(76.5100671140940,\t0.120017116346111)\n(77.5167785234899,\t0.116751497680285)\n(78.5234899328859,\t0.113530922030539)\n(96.6442953020134,\t0.0632629160848610)\n(97.6510067114094,\t0.0608981577406423)\n(98.6577181208054,\t0.0585784424125040)\n(99.6644295302014,\t0.0563037701004459)\n(100.671140939597,\t0.0540741408044683)\n(101.677852348993,\t0.0518895545245710)\n(102.684563758389,\t0.0497500112607541)\n(103.691275167785,\t0.0476555110130175)\n(104.697986577181,\t0.0456060537813612)\n(105.704697986577,\t0.0436016395657852)\n(106.711409395973,\t0.0416422683662898)\n(107.718120805369,\t0.0397279401828747)\n(108.724832214765,\t0.0378586550155399)\n(109.731543624161,\t0.0360344128642853)\n(141.946308724832,\t0.00144137651457144)\n(142.953020134228,\t0.00110355389396877)\n(143.959731543624,\t0.000810774289446425)\n(144.966442953020,\t0.000563037701004410)\n(145.973154362416,\t0.000360344128642831)\n(146.979865771812,\t0.000202693572361579)\n(150,\t0)\n\n\t};\n\\addlegendentry{\\Large{SIL}}\n\\end{axis}\n\\end{tikzpicture}}\n\\vspace{-36pt}\n\\caption{\\scriptsize{Weight Trade-off.}}\\label{wrap-weights}\n\\end{wrapfigure}\n\n\\noindent\nThe following composite loss is used to train RMIH: \n\\begin{equation}\nJ = \\lambda_{\\text{MI}}^{t} J_{\\mathcal{S}}^{\\text{MI}}+ \\lambda_{\\text{SI}}^{t} J_{\\mathcal{S}}^{\\text{SI}}+ \\lambda_q J_{Q} + \\lambda_w R_{W}\n\\end{equation}\nwhere $ \\lambda_{\\text{MI}}^{t}$, $ \\lambda_{\\text{SI}}^{t}$, $\\lambda_q$ and $\\lambda_w$ are hyper-parameters that control the contribution of each of the loss terms. Specifically, $ \\lambda_{\\text{MI}}^{t}$ and $ \\lambda_{\\text{SI}}^{t}$ control the trade-off between the MI and SI hashing losses. The SI arm plays a significant role only in the early stages of training and can be traded off eventually to avoid sub-optimal MI hashing. For this we introduce a weight trade-off formulation that gradually down-regulates $ \\lambda_{\\text{SI}}^{t}$, while simultaneously up-regulating $ \\lambda_{\\text{MI}}^{t}$. Here, we use $\\lambda_{\\text{SI}}^{t} = 1 - 0.5\\left ( 1 - \\nicefrac{t}{t_{\\text{max}}} \\right )^{2}$ and $\\lambda_{\\text{MI}}^{t} = 1 - \\lambda_{\\text{SI}}^{t}$, where $t$ is the current epoch and $t_{\\text{max}}$ is the maximum number of epochs (see Fig.~\\ref{wrap-weights}). We train RMIH with mini-batch stochastic gradient descent (SGD) with momentum. Due to potential outliers that can occur at the beginning of training, we scale $c_{k}$ up by a factor of 7 for $t = 1$ to allow a stable state to be reached. \nSpecifically, the gradient of $J_{\\mathcal{S}}^{(\\cdot)}$ w.r.t. to $\\mathbf{h}_{i}$ is derived as: \n\\begin{align}\n\\frac{\\partial J_{\\mathcal{S}}^{(\\cdot)}}{\\partial \\mathbf{h}_{i}} = &\\left ( \\sum_{l:s_{li} > 0} p_{li}\\mathbf{\\mathcal{L'_H}}(\\mathbf{h}_l,\\mathbf{h}_i) - \\sum_{l \\neq i}\\left ( \\sum_{q:s_{lq} > 0} p_{lq} \\right ) p_{li}\\mathbf{\\mathcal{L'_H}}(\\mathbf{h}_l,\\mathbf{h}_i) \\right ) \\nonumber \\\\ \n- &\\left ( \\sum_{j:s_{ij} > 0} p_{ij}\\mathbf{\\mathcal{L'_H}}(\\mathbf{h}_i,\\mathbf{h}_j) - \\sum_{j:s_{ij} > 0} p_{ij}\\left ( \\sum_{z \\neq i} p_{iz}\\mathbf{\\mathcal{L'_H}}(\\mathbf{h}_i,\\mathbf{h}_z) \\right ) \\right ) \n\\label{eq:derJSbig}\n\\end{align}\n\\noindent\nwhere ${\\mathcal{L_H'}}(\\mathbf{h_i},\\mathbf{h_j}) = \\{ {\\rho_k'}(r_{ij}^k)\\}_{k=1}^{k}$. The derivative of the huber term ${\\rho_k}'(r_{ij}^k)$ can be computed as:\n\\begin{equation}\n{\\rho_k'}(r_{ij}^k) = \\begin{cases}\nr_{ij}^k, &\\text{if\\ $\\mid r_{ij}^k \\mid \\leqslant c_k$}\\\\\nc_k\\ \\text{sgn}(r_{ij}^k), &\\text{if $\\mid r_{ij}^k \\mid > c_k$}\n\\end{cases}\n\\end{equation}\n\nRegarding the quantization loss function, the derivative can be computed by $\\frac{\\partial J_{Q}}{\\partial \\mathbf{h}_{i}} = \\text{tanh}\\left ( \\left | \\mathbf{h}_{i} \\right | - \\mathbf{1} \\right )\\text{sgn}\\left ( \\mathbf{h}_{i} \\right )$. Having computed these gradients, we use backpropagation to compute the derivatives of the preceding layers.\n\n\n\\begin{figure}[t]\n\\centering\n\\begin{minipage}{\\textwidth}\n \\begin{minipage}[t]{0.6\\textwidth}\n\\centering\n\\includegraphics[width=\\textwidth]{RandomImages}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.36\\textwidth}\n\\vspace{-3.4cm}\n\\captionof{figure}{Select images from the \\textit{IPUHL} and \\textit{DDSM} datasets to showcase the degree of anatomical variability within and across the classes.}\n\\label{fig:randomimages}\n \\end{minipage}\n \\end{minipage}\n \\vspace{-0.7cm}\n\\end{figure} \n\n\n\n\n\\section{Experiments}\n\n\\textbf{Databases:} Clinical applicability of RMIH has been validated on two large scale datasets, namely, Digital Database for Screening Mammography (DDSM)~\\cite{mammogram1,mammogram2} and a retrospectively acquired histology dataset from the Indiana University Health Pathology Lab (\\textit{IUPHL})~\\cite{iupui1,iupui2}. The \\textit{DDSM} dataset comprises of 11,617 expert selected regions of interest (ROI) curated from 1861 patients. Multiple ROIs associated with a single breast from anatomical views constitute a bag (size: 1-12; median: 2), which has been annotated as normal, benign or malignant by expert radiologists.\nA bag labeled \\textit{malignant} could potentially contain multiple suspect normal and benign masses, which have not been individually identified. The \\textit{IUPHL} dataset is a collection of 653 ROIs from histology slides from 40 patients (20 with precancerous ductal hyperplasia (UDH) and rest with ductal carcinoma \\textit{in situ} (DCIS)) with ROI level annotations done by expert histopathologists. Due to high variability in sizes of these ROIs (upto 9K $\\times$ 8K pixels), we extract multiple patches (of size $1024\\times1024$) and populate a ROI-level bag (size: 1-15; median: 8). \nAs cellular and nuclei level characteristics are important to distinguishing DCIS from UDH, it is not recommended to rescale these images to standard input sizes used by CNNs (typically, 244 $\\times$ 224 in~\\cite{vgg, residualHe, goingdeeper}). Fig.~\\ref{fig:randomimages} illustrates select images from the two datasets to showcase anatomical variability within and across the constituent classes.\nFrom both the datasets, we use patient-level\nnon-overlapping splits to constitute the training (80\\%) and testing (20\\%) sets. \n\n\\textbf{Model Settings and Validations}: To validate proposed contributions, namely robustness within NCA loss and trade-off from the aux-SI arm, we perform ablative testing with combinations of their baseline variants by fine-tuning multiple network architectures. Additionally, we compare RMIH against four state-of-the art methods: ITQ~\\cite{ITQ}, KSH~\\cite{KSH}, SFLH~\\cite{lai2015} and DHN~\\cite{zhuloss}. For a fair comparison, we use R50 for both SFLH and DHN, since as discussed later it performs the best. Since SFLH and DHN were originally proposed for SI hashing, we introduce additional MI variants by hashing through the MIPool layer.\nFor ITQ and KSH, we further create two comparative settings: \\textbf{1)} Using IMIH~\\cite{bmih} that learns instance-level hash codes followed by bag-level distance computation and \\textbf{2)} Utilizing BMIH~\\cite{bmih} using bag-level kernalized representations followed by binarization.\nFor IMIH and SI variants of SFLH, DHN and RMIH, given two bags $B_p$ and $B_q$ with SI hash codes, say $\\mathcal{H}(B_q) = \\{h_{q1},\\dots,h_{qM}\\}$ and $\\mathcal{H}(B_p) = \\{h_{p1},\\dots,h_{pN}\\}$, the bag-level distance is computed as: \n\\begin{equation}\nd(B_p,B_q) = \\frac{1}{M}\\sum_{i=1}^{M}(\\min_{\\forall j }\\ \\text{Hamming}(h_{pi},h_{qj})).\n\\label{eq:bagDist}\n\\end{equation}\n\n\n\n\n\nAll images were resized to $224 \\times 224$ and training data were augmented\nto create equally balanced classes. $ \\lambda_{\\text{MI}}^{t}$ and $ \\lambda_{\\text{SI}}^{t}$ were set assuming $t_{\\text{max}}$ as 150 epoch; $\\lambda_q$ and $\\lambda_w$ were set at 0.05 and 0.001 respectively. The momentum term within SGD was set to 0.9 and batch size to 128 for \\textit{DDSM} and 32 for \\textit{IUPHL}. For efficient learning, we use an exponentially decaying learning rate initialized at 0.01. The RMIH framework was implemented in MatConvNet~\\cite{matconvnet}. We use standard retrieval quality metrics: nearest neighbor classification accuracy (nnCA) and precision-recall (PR) curves to perform the aforementioned comparisons. The results (nnCA)\nfrom ablative testing and comparative methods are tabulated in Table~\\ref{wrap-tab:ablative} and Table~\\ref{table:CM} respectively. Within Table~\\ref{table:CM}, methods were evaluated at two different code sizes (16 bits and 32 bits). We also present the PR curves of select bag-level methods (32 bits) in Fig. \\ref{fig:PRCurves}.\n\n\n\n\n\n\n\n\n\\section{Results and Discussion} \\label{results}\n\n\\begin{figure}[t]\n\\resizebox{\\textwidth}{!}{\\begin{tabular}{|c|c|C|C|c|c|c|c|c|c|} \\hline\n\\multicolumn{2}{|c|}{\\multirow{2}{*}{\\textbf{Method}}} & \\multicolumn{2}{c|}{\\thead{\\textbf{Variants}}} & \\multicolumn{3}{c|}{\\thead{\\textit{\\textbf{DDSM}}}} & \\multicolumn{3}{c|}{\\thead{\\textit{\\textbf{IUPHL}}}}\\\\ \\cline{3-10}\n\\multicolumn{2}{|c|}{\\multirow{2}{*}{}} & \\thead{R} &\\thead{T} &\\thead{\\ VGGF$\\ $} &\\thead{\\ R50$\\ $} &\\thead{\\ GN$\\ $ }&\\thead{\\ VGGF$\\ $} &\\thead{\\ R50$\\ $} &\\thead{\\ GN$\\ $} \\\\ \\hline\n\\multirow{5}{*}{\\makecell{{\\small \\textbf{Ablative}}\\\\ {\\small \\textbf{Testing}}}} & A & $\\circ$ & $\\circ$ & 68.65 & 72.76 & 71.70 & 83.85 & 85.42 & 82.29 \\\\ \\cline{2-10}\n& B & $\\circ$ & $\\bullet$ & 75.38 & 77.34 & 72.92 & 85.94 & 90.10 & 88.02 \\\\ \\cline{2-10}\n& C & $\\bullet$ & $\\circ$ & 70.65 & 76.63 & 70.02 & 83.33 & 85.94 & 86.46 \\\\ \\cline{2-10}\n& D & $\\circ$ & \\tiny{$\\blacksquare$} & 66.65 & 69.67 & 68.26 & 83.33 & 88.54 & 84.90 \\\\ \\cline{2-10}\n& E & $\\bullet$ & \\tiny{$\\blacksquare$} & 67.05 & 76.59 & 72.84 & 84.38 & 89.58 & 85.42 \\\\\\hline\n\\multicolumn{2}{|c|}{\\textbf{RMIH-mean}} & $\\bullet$ & $\\bullet$ & 78.67 & 82.31 & 76.83 & 87.50 & 89.58 & \\textbf{89.06} \\\\ \\hline\n\\multicolumn{2}{|c|}{\\textbf{RMIH-max}}& $\\bullet$ & $\\bullet$ & \\textbf{81.21} & \\textbf{85.68} & \\textbf{78.67} & \\textbf{91.67} & \\textbf{95.83} & 88.02 \\\\ \\hline\n\\multicolumn{2}{|c|}{\\textbf{RMIH}($\\lambda_q=0$)} & $\\bullet$ & $\\bullet$ & 75.34 & 79.88 & 73.06 & 87.50 & 89.58 & 88.51 \\\\ \\hline\n\\multicolumn{2}{|c|}{\\textbf{RMIH NB}} & $\\bullet$ & $\\bullet$ & 83.25 & 88.02 & 79.06 & 94.79 & 96.35 & 92.71 \\\\ \\hline\n\\multirow{4}{*}{\\textbf{Legend}} \n& \\multicolumn{3}{r|}{\\textbf{R(Robustness)}} & \\multicolumn{6}{l|}{ $\\circ=L_2,\\ \\bullet=L_{\\text{Huber}}$}\\\\ \n& \\multicolumn{3}{r|}{\\makecell[r]{\\textbf{T(Trade-off)}}} & \\multicolumn{6}{l|}{\\makecell[l]{$\\circ=\\text{\\small{Equal weights,}}\\ \\bullet=\\text{\\small{Decaying SIL weights,}}$ \\\\ \\scriptsize{$\\blacksquare$} \\small{$=\\text{No SIL branch}$}}}\\\\ \n& \\multicolumn{3}{r|}{\\textbf{Networks}} & \\multicolumn{6}{l|}{R50: ResNet50,\\ GN: GoogleNet} \\\\ \\hline\n\\end{tabular}}\n\\caption{Performance of ablative testing at code size of 16 bits. We report the nearest neighbor classification accuracy (nnCA) estimated over unseen test data. Letters A-E are introduced for easier comparisons, discussed in Section \\ref{results}.}\n\\label{wrap-tab:ablative}\n \\vspace{-0.5cm}\n\\end{figure}\n\n\n\\textbf{Effect of aux-SI Loss}: To justify using the aux-SI loss, we introduce a variant of RMIH without it (E in Table~\\ref{wrap-tab:ablative}), which\nleads to a significant decline of 3\\% to 14\\% \nin contrast to RMIH. This could be potentially attributed to the prevention of the gradient sparsification caused by the MIPool layer. From Table~\\ref{wrap-tab:ablative}, we observe a 3\\%-10\\% increase in performance, comparing cases with gradual decaying trade-off (B)\nagainst baseline setting ($\\lambda^{t}_{\\text{MI}} = \\lambda^{t}_{\\text{SI}} = 0.5$, A,C).\n\n\n\\textbf{Effect of Robustness}: For robust-NCA, we compared against the original NCA formulation proposed in~\\cite{torralba} (A,B,D in Table~\\ref{wrap-tab:ablative}). Robustness helps handle potentially noisy MI labels, inconsistencies within a bag (like non-informative patches) and the ambiguity in assigning SI labels. Comparing the effect of robustness \nfor baselines sans the SI hashing arm (D \\textit{vs.} E) \nwe observe marginally positive improvement across the architectures and datasets, with a substantial 7\\% in ResNet50 for \\textit{DDSM}. Robustness contributes more with the addition of the aux-SI hash arm (proposed \\textit{vs.} E)\nwith improved performance in the range of 4\\%-5\\% across all settings. This observation further validates our prior argument. \n\n\\textbf{Effect of Quantization}: To assess the effect of quantization, we define two baselines: (1) setting $\\lambda_{q} = 0$ and (2) using non-quantized hash codes for retrieval (RMIH - NB). The latter potentially acts as an upper bound for performance evaluation. From Table~\\ref{wrap-tab:ablative}, we observe a consistent increase in performance by margins of 3\\%-5\\% if RMIH is learnt with an explicit quantization loss to limit the associated error. It must also be noted that comparing with RMIH - NB, there is only a marginal fall in performance (2\\%-4\\%), which is desired. \n\n\nComparing max \\textit{vs.} mean MI Pool variants, we observe that max achieves marginally better performance, since it is more selective than mean, which is particularly important in cases of detecting malignancy. \n\n\\begin{figure}[t]\n\\includegraphics[width=\\textwidth]{RetrievalLarge_compressed}\n\\caption{Retrieval results for RMIH at code size 16 bits. +1 indicates retrieval from class consistent with query and -1 indicates otherwise.}\n\\label{fig:retrievedimages}\n\\vspace{-8pt}\n\\end{figure}\n\n\\begin{figure}[t]\n\\centering\n\\begin{minipage}{\\textwidth}\n \\begin{minipage}[t]{0.7\\textwidth}\n\\centering\n\\includegraphics[width=\\textwidth]{PR_Curves-cropped}\n \\end{minipage}\n \\hfill\n \\begin{minipage}[t]{0.26\\textwidth}\n\\vspace{-2.7cm}\n\\captionof{figure}{PR curves for \\textit{DDSM} and \\textit{IUPHL} datasets at code size of 32.}\n\\label{fig:PRCurves}\n \\end{minipage}\n \\end{minipage}\n \\vspace{-0.7cm}\n\\end{figure} \n\n\n\n\nAs a whole, the two-pronged proposed approach, including robustness and trade-off, along with quantization loss delivers the highest performance, proving that RMIH is able to learn effectively, despite the ambiguity induced by the SI hashing arm. Fig.~\\ref{fig:retrievedimages} demonstrates the retrieval performance of RMIH on the target databases. For \\textit{IUPHL}, the retrieved images are semantically similar to the query as consistent anatomical signatures\nare evident in the retrieved neighbors. For \\textit{DDSM}, in the cancer and normal cases the retrieved neighbors are consistent, however it is hard to distinguish between benign and malignant. The retrieval time for a single query for RMIH was observed at 31.62 ms (for \\textit{IUPHL}) and 17.48 ms (for \\textit{DDSM}) showing potential for fast and scalable search. \n\n\n\n\n\n\\begin{wraptable}[18]{L}{6.0cm}\n\\vspace{-0.4cm}\n \\centering\n \\resizebox{5.9cm}{!}{\\begin{tabular}{x{0.4cm}|x{2.2cm}|x{1cm}|x{0.5cm}|x{1cm}|x{1cm}|x{1.2cm}|x{1cm}|}\n\\multicolumn{2}{c|}{\\multirow{2}{*}{\\textbf{Method}}} & \\multirow{2}{*}{\\thead{\\textbf{A\/F}}} & \\multirow{2}{*}{\\thead{\\textbf{L}}} & \\multicolumn{2}{c|}{\\thead{\\textit{\\textbf{DDSM}}}} & \\multicolumn{2}{c|}{\\thead{\\textit{\\textbf{IUPHL}}}} \\tabularnewline \\cline{5-8} \n\\multicolumn{2}{c|}{} & & & 16-bit & 32-bit & 16-bit & 32-bit \\tabularnewline \\cline{1-8}\n\\multirow{8}{*}{\\rotatebox[origin=c]{90}{\\textbf{Shallow}}} & \\multirow{4}{*}{ITQ \\cite{ITQ}} & R50 & $\\circ$ & 66.35 & 67.71 & 78.58 & 80.28 \\tabularnewline \\cline{3-8}\n& & R50 & $\\bullet$ & 64.56 & 71.98 & 89.58 & 79.69 \\tabularnewline \\cline{3-8} \n& & G & $\\circ$ & 65.22 & 66.55 & 51.79 & 51.42 \\tabularnewline \\cline{3-8}\n& & G & $\\bullet$ & 59.73 & 61.03 & 57.29 & 58.85 \\tabularnewline \\cline{2-8}\n& \\multirow{4}{*}{KSH \\cite{KSH}} & R50 & $\\circ$ & 61.88 & 64.81 & 87.74 & 86.51 \\tabularnewline \\cline{3-8}\n& & R50 & $\\bullet$ & 59.81 & 72.17 & 70.83 & 80.21 \\tabularnewline \\cline{3-8}\n& & G & $\\circ$ & 60.50 & 61.91 & 57.36 & 57.83 \\tabularnewline \\cline{3-8}\n& & G & $\\bullet$ & 55.34 & 55.67 & 60.94 & 58.85 \\tabularnewline \\hline\n\\multirow{6}{*}{\\rotatebox[origin=c]{90}{\\textbf{Deep}}} & \\multirow{2}{*}{SFLH \\cite{lai2015}} & R50 & $\\circ$ & 73.54 & 77.46 & 83.33 & 85.94 \\tabularnewline \\cline{3-8}\n& & R50M & \\tiny{$\\blacksquare$} & 71.98 & 75.93 & 85.42 & 88.54 \\tabularnewline \\cline{2-8}\n& \\multirow{2}{*}{DHN \\cite{zhuloss}} & R50 & $\\circ$ & 65.64 & 74.79 & 82.29 & 86.46 \\tabularnewline \\cline{3-8}\n& & R50M & \\tiny{$\\blacksquare$} & 72.88 & 80.43 & 88.02 & 90.62 \\tabularnewline \\cline{2-8}\n& RMIH-SIL & R50 & $\\circ$ & 76.02 & 78.37 & 87.92 & 88.58 \\tabularnewline \\cline{2-8}\n& \\textbf{RMIH} & R50M & \\tiny{$\\blacksquare$} & \\textbf{85.68} & \\textbf{89.47} & \\textbf{95.83} & \\textbf{93.23} \\tabularnewline \\hline\n\\multirow{3}{*}{\\rotatebox[origin=c]{90}{\\textbf{Legend}}} & \\makecell*[r]{\\textbf{A\/F:}} & \\multicolumn{6}{l|}{\\makecell*[l]{\\small{\\textbf{A}: Architecture,\\ \\textbf{F}: Features} \\tabularnewline \\scriptsize{R50: ResNet50, R50M: ResNet50$+$MIPool, G: GIST}}} \\tabularnewline \\cline{2-8}\n& \\makecell*[r]{\\textbf{L:}} & \\multicolumn{6}{l|}{\\makecell*[l]{$\\circ=\\text{IMIH}$,\\ $\\bullet=\\text{BMIH}$,\\ \\footnotesize{$\\ \\blacksquare$\\ }=\\small{\\ End-to-end}}} \\tabularnewline \\hline\n\\end{tabular}}\n\\caption{Results of comparison with state-of-the art hashing methods. }\\label{table:CM}\n\\end{wraptable}\n\n\\noindent\n\\textbf{Comparative Methods}\nIn the contrastive experiments against ITQ and KSH, hand-crafted GIST \\cite{gist} features underperformed significantly, while the improvement with the R50 features ranged from 5\\%-30\\%. However, RMIH still performed 10\\%-25\\% better, proving that even if deep learnt features severely boost the performance, the shallow methods cannot fully breach the gap to the deep ones.\nComparing the SI with the MI variations of DHN, SFLH and RMIH, it is observed that the performance improved in the range of 3\\%-11\\%, suggesting that end-to-end learning of MI hash codes is preferred over two-stage hashing \\textit{i.e.} hashing at SI level and comparing at bag level with Eq.~\\eqref{eq:bagDist}. However, RMIH fares comparably better than both the SI and MI versions of SFLH and DHN, owing to the robustness of the proposed retrieval loss function to potentially noisy labels and inconsistent instances within bags (also evident from PR curves in Fig. \\ref{fig:PRCurves}). In all fairness, the concepts of training with aux-SI hashing arm with gradual trade-off and robustness could be potentially adapted to SFLH and DHN to improve their MI hashing performance. \nAs also seen from the associated PR curves in Fig. \\ref{fig:PRCurves}, the performance gap between shallow and deep hashing methods remains significant despite using R50 features. Comparative results strongly support our premise that end-to-end learning of MI hash codes is preferred over conventional two-stage approaches. \n\n\n\n\n\n\n\\section{Conclusion}\nIn this paper, for the first time, we proposed an end-to-end deep robust hashing framework, termed RMIH, for retrieval under a multiple instance setting. We incorporate the MIPool layer to aggregate representations across instances to generate a bag-level discriminative hash code. We introduce the notion of robustness into our supervised retrieval loss and improve the trainability of RMIH by utilizing an aux-SI hashing arm regulated by a trade-off. Extensive validations and ablative testing on two public breast cancer datasets\ndemonstrate the superiority of RMIH and its potential for future extension to other MI applications. \n\\vspace{-0.15cm}\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{ Introduction}\n\\medskip\n\n\nThe mean curvature flow (abbreviated as MCF) of a submanifold $M \\subset \\mathbb{R}^{N}$\nover an interval $I$ is a map\n$ f: I \\times M \\longrightarrow \\mathbb{R}^{N} $\nsuch that for all $t \\in I$ and $x \\in M$, $\\frac{\\partial}{\\partial t} f(t, x)$ is equal\nto the mean curvature vector of $M(t)= f(t, M)$ at the point\n$x(t) = f(t, x)$.\nMean curvature flows of convex hypersurfaces have been extensively studied in the literature\n(cf. \\cite{GH}, \\cite{Hu}).\nAn exposition of the work in this area was given in the book \\cite{Zhu}.\nComparatively, the behavior of mean curvature flows of submanifolds with higher codimension\nis less understood (cf. \\cite{W}). This is partly due to the lack of understanding of collapsing\nand the formation of singularities of the flow equations in the higher codimensional case.\n\n\nA submanifold $M$ of a Riemannian manifold is {\\it isoparametric} if\nits normal bundle is flat and principal curvatures along any parallel normal\nvector field are constant. The codimension of $M$ is called the {\\it rank} of $M$.\n An isoparametric submanifold $M$ in $\\mathbb{R} ^N$ is {\\it full} if it is not contained in any proper\n hyperplane, and\nis {\\it irreducible\\\/} if it is not a product of two isoparametric submanifolds.\nWe refer to \\cite{Terng} for the basic properties and structure\ntheories for isoparametric submanifolds.\nPrincipal orbits of isotropy representations of symmetric spaces are isoparametric, they are the only compact homogeneous isoparametric submanifolds in Euclidean space\n(cf. \\cite{PT}), and are called {\\it generalized flag manifolds}.\nThere are also infinite families of non-homogeneous isoparametric submanifolds which arise from\nrepresentations of Clifford algebras (cf. \\cite{FKM}).\nAll these non-homogeneous examples have rank 2. A theorem of Thorbergsson \\cite{Th} asserts\nthat compact full irreducible isoparametric submanifolds with rank bigger than 2 are always\nhomogeneous.\n\nA complete isoparametric submanifold of $\\mathbb{R}^{N}$ can be\ndecomposed as the product of a compact, irreducible, isoparametric\nsubmanifold and a subspace of $\\mathbb{R}^{N}$. Since mean\ncurvature flows with affine subspaces of $\\mathbb{R}^{N}$ as\ninitial data is trivial and the mean curvature flow starting from\na product submanifold stays as product, we will only consider\ncompact, full, irreducible isoparametric submanifolds.\n\nLet $M$ be an isoparametric submanifold of $\\mathbb{R} ^N$, and $\\xi$ a parallel normal vector field\non $M$. Then\n$M_\\xi=\\{x+ \\xi(x)\\ \\vert\\ x\\in M\\}$\nis again a smooth submanifold (may have higher codimension), and the map $M\\to M_\\xi$ is either a\ndiffeomorphism or a fibration with a generalized flag manifold as fiber.\nThe family of these parallel sets forms a singular foliation of the ambient Euclidean space $\\mathbb{R} ^N$.\nTop dimensional leaves are all isoparametric in $\\mathbb{R}^{N}$, and they are called {\\it parallel\nisoparametric submanifolds}.\nLower dimensional leaves are no longer isoparametric, and they are called\n{\\it focal submanifolds\\\/}.\n\n\nWe show that if $f:M\\times [0, T)\\to \\mathbb{R} ^N$ is a solution of the MCF in $\\mathbb{R} ^N$ with\n$f(\\cdot, 0)$ isoparametric then $f(\\cdot, t)$ is isoparametric for all $t\\in [0,T)$, i.e.,\nthe MCF preserves isoparametric condition. This reduces the MCF to a system of ordinary\ndifferential equations. There is a Weyl group $W$ associated to each isoparametric submanifold $M$ that acts on the normal plane $p+ \\nu_p M$. The ODE given by the mean curvature flow with initial data an isoparametric submanifold is given by a vector field $H$ smoothly defined on the interior of the Weyl chamber $C$ of $W$ but blows up at the boundary of $C$. However, we can use generators of $W$-invariant polynomials to change coordinate so that the vector field $H$ becomes a polynomial vector field and its flows can be solved explicitly.\n\nEvery compact isoparametric submanifold is contained in a sphere. This sphere is\nalso foliated by parallel isoparametric submanifolds and focal submanifolds.\nEach isoparametric foliation contains a unique isoparametric submanifold which is a minimal\nsubmanifold of this sphere. The mean curvature flow in $\\mathbb{R} ^N$ with initial data a minimal submanifold\nin $S^{N-1}$ behaves like the mean curvature flow of a sphere, i.e. it just shrinks homothetically\nalong the radial direction and collapses to a point in finite time\n(cf. Lemma \\ref{miniso}). If $M$ is an isoparametric submanifold in $\\mathbb{R} ^N$\nwhich is not minimal in the sphere, then its mean curvature flow will converge to a focal\nsubmanifold $F$ of positive dimension (cf. Corollary \\ref{cor:miniso}).\nIn fact, $M$ is a fibration over $F$ with each fiber a homogeneous isoparametric submanifold of a\nlower dimensional Euclidean space.\nEach fiber of this fibration collapses\nto a point under the mean curvature flow in a finite time.\n\nWe summarize some of the main results of this paper in the following Theorem (cf. Theorem \\ref{thm:finiteconv}, Theorem \\ref{thm:focalconv}, and Proposition \\ref{bm}):\n\n\\begin{thm}\\label{bo}\nThe mean curvature flow in $\\mathbb{R} ^N$ with initial data a compact isoparametric submanifold\n\\begin{enumerate}\n\\item converges to a focal submanifold in finite time $T$,\n\\item if the fibration from the initial isoparametric submanifold to the limiting focal\nsubmanifold is a sphere fibration (this is the generic case), then the mean curvature flow\n$M(t)$ has type I singularity, i.e., there is a constant $c_0$ such that\n$||{\\rm II\\\/}(t)||_\\infty^2(T-t)\\leq c_0$ for all $t\\in [0,T)$, where $||{\\rm II\\\/}(t)||_\\infty$ is\nthe sup norm of the second fundamental form of $M(t)$,\n\\item every focal submanifold is the limit of\nthe mean curvature flow with some parallel isoparametric submanifold as initial data,\n\\item if $M_1$ and $M_2$ are distinct parallel full isoparametric submanifolds in\n$\\mathbb{R} ^N$ that lie in the same sphere. Then the mean curvature flows in $\\mathbb{R} ^N$ with initial data\n$M_1$ and $M_2$ collapse to two distinct focal submanifolds.\n\\end{enumerate}\n\\end{thm}\n\n The mean curvature flow in $S^{N-1}$ with initial data an isoparametric submanifold behaves very similarly to\nthe Euclidean mean curvature flow. In particular we have the following theorem:\n\n\\begin{thm} \\label{thm:MCFisopS}\nLet $M$ be an isoparametric submanifold of $S^{N-1}$. Then the mean curvature flow in\n$S^{N-1}$ with $M$ as initial data\n\\begin{enumerate}\n\\item is constant if $M$ is minimal in $S^{N-1}$, or\n\\item converges to a focal submanifold of positive dimension in finite\ntime if $M$ is not minimal.\n\\end{enumerate}\n\\end{thm}\n\nAn isometric action of $G$ on a Riemannian manifold $N$ is {\\it polar\\\/} if there exists a closed embedded submanifold $\\Sigma$ of $N$ that meets all $G$-orbits and meets orthogonally. Such $\\Sigma$ is called a {\\it section\\\/} of the polar action. Principal orbits of a polar action in $\\mathbb{R} ^n$ and $S^n$ are isoparametric (cf. \\cite{PT}). We prove that if the $G$-action on $N$ is polar then the mean curvature flow preserves $G$-orbits and the flow becomes an ordinary differential equation on the section $\\Sigma$. We expect that methods developed in this paper can be applied to study mean curvature flows for orbits of polar actions with flat sections in symmetric spaces.\n\nThis paper is organized as follows: We give a brief review of properties of isoparametric submanifolds that are needed in section \\ref{sec:prelim}, present proofs of results stated in Theorem \\ref{bo} in section \\ref{sec:Basic}, construct explicit solutions of the MCF in $\\mathbb{R} ^N$ with initial data an isoparametric submanifold in section \\ref{sec:solution}. Since focal submanifolds are smooth manifolds, we can consider their mean curvature\nflow. Most properties of the mean curvature flows for isoparametric submanifolds\nalso hold for focal submanifolds. This will be briefly discussed in section \\ref{sec:focal}.\nWe describe MCF in spheres with initial data an isoparametric submanifold in spheres in section \\ref{sec:sphere}, and in the last section we discuss MCF in a Riemannian manifold $N$ with initial data a principal orbit of a polar action on $N$.\n\n\nThe authors like to thank Mu-Tao Wang, Yng-Ing Lee, and Mao-Pei Tsui for discussions on types of singularities of the mean curvature flows.\n\n\\bigskip\n\\section{Preliminaries}\n\\label{sec:prelim}\n\nGeometric and topological properties of isoparametric submanifolds can be found in\n\\cite{Terng}. In this section we\nbriefly review the properties which will be used in this paper.\nLet $M \\subset \\mathbb{R}^{N}$ be a full compact\nisoparametric submanifold of rank $k$.\n\n\\subsection{Curvature spheres and curvature normals}\n\n\\hspace{100pt}\\\\\nThe tangent bundle of $M$ can be decomposed into\northogonal sums of {\\it curvature distributions} $\\{ E_{i} \\mid i=1, \\cdots, g\\}$\nfor some integer $g>0$. At each point of $M$, $E_{i}$ is a common eigenspace\nof the shape operators of $M$ at that point.\n There are parallel normal vector fields ${\\bf n}_{i}$\nsuch that the shape operator $A_{\\xi}$ has the property\n\\[ A_{\\xi} |_{E_{i}} = \\langle \\xi, {\\bf n}_{i}\\rangle {\\rm Id\\\/}_{E_i} \\]\nfor all normal vector $\\xi$.\nEach vector field ${\\bf n}_{i}$ is called the {\\it curvature normal} of $E_{i}$. The\nrank of $E_{i}$ is called the {\\it multiplicity} of ${\\bf n}_{i}$,\nwhich will be denoted by $m_{i}$.\nEach $E_{i}$ is an integrable distribution whose leaves are $m_{i}$-dimensional\nround spheres with radius $1\/\\|{\\bf n}_{i}\\|$. Such spheres are called\n{\\it curvature spheres}.\n\n\n\n\\subsection{Weyl group associated to $M$}\\label{ae}\n\n\\hspace{100pt}\\\\\nFor each $i \\in \\{1, \\cdots, g\\}$, let $\\sigma_{i}(x)$ be the antipodal point in the\n$i$-th curvature sphere passing through $x$. Then $\\sigma_{i}$ is an involution on $M$.\nThe group $W$ generated by $\\sigma_{1}, \\cdots, \\sigma_{g}$ is a crystallographic {\\it Coxeter group}.\nIt is known that $M$ is irreducible if and only if $W$ is irreducible. For each $x \\in M$,\n$W$ also acts as a reflection group\non the affine normal space $x+\\nu_{x}M$\ngenerated by reflections along hyperplanes\n\\[ L_{i} := \\{ x+\\xi \\mid \\xi \\in \\nu_{x}M, \\,\\,\\, 1- \\langle \\xi, {\\bf n}_{i}(x)\\rangle = 0 \\} \\]\nfor $i=1, \\cdots, g$.\n\nThe intersection $\\bigcap_{i=1}^{g} L_{i}$ consists of a single constant point\nwhich is denoted by $a$. Then $M$ is contained in a sphere which is centered at $a$.\nWithout loss of generality, we always {\\bf assume} that $a=0$, i.e., $M$ is contained in a sphere\ncentered at the origin of $\\mathbb{R}^{N}$. This condition is equivalent to\n\\begin{equation} \\label{eqn:centerO}\n \\langle -x, \\, {\\bf n}_{i}(x)\\rangle \\,\\, = \\,\\, 1\n\\end{equation}\nfor all $x \\in M$ and $i=1, \\cdots, g$ (cf. \\cite[Corrolary 1.17]{Terng}).\n\n\n\\subsection{Parallel submanifold}\n\n\\hspace{100pt} \\\\\n For any parallel normal vector field $\\xi$ on $M$, define\n\\[ M_{\\xi}:= \\{ x+\\xi(x) \\mid x \\in M \\}. \\]\nIf\n\\begin{equation} \\label{eqn:noncollcond}\n 1- \\langle\\xi(x), {\\bf n}_{i}(x)\\rangle \\neq 0\n\\end{equation}\n for $i=1, \\cdots, g$, then\n$M_{\\xi}$ is again an isoparametric submanifold with the same dimension as $M$.\n$M_{\\xi}$ is called the {\\it parallel isoparametric submanifold} of $M$ defined by $\\xi$.\nThe curvature normals of $M_{\\xi}$ at the point $x+ \\xi(x)$ are given\nby\n\\[ \\frac{{\\bf n}_{i}(x) }{1- \\langle\\xi(x), {\\bf n}_{i}(x)\\rangle} \\]\nwith same multiplicities $m_{i}$ for $i=1, \\cdots, g$.\nThe mean curvature vector of $M_{\\xi}$ at $x+ \\xi(x)$ is given\nby\n\\begin{equation} \\label{eqn:MCV}\n H(x+\\xi(x)) = \\sum_{i=1}^{g} \\frac{m_{i}{\\bf n}_{i}(x) }{1- \\langle\\xi(x), {\\bf n}_{i}(x)\\rangle}.\n\\end{equation}\n\nWhen condition \\eqref{eqn:noncollcond} fails, $M_{\\xi}$ is still\na smooth submanifold of $\\mathbb{R}^{N}$, but it is no longer\nisoparametric. This submanifold is called\na focal submanifold of $M$. The dimension of $M_{\\xi}$ is strictly\nsmaller than that of $M$. The map\n\\[ \\begin{array}{crcl}\n\\pi:& M & \\longrightarrow & M_{\\xi} \\\\\n & x & \\mapsto & x+\\xi(x)\n\\end{array} \\]\nis a fibration over $M_{\\xi}$ with each fiber an isoparametric submanifold in the normal space\nof $M_{\\xi}$ at $\\pi(x)$.\n In fact, fix $x_0\\in M$, let $C$ denote the Weyl chamber of $W$ on $x_0+ \\nu_{x_0}M$\n containing $x_0$, i.e.,\n $$C=\\{x_0+\\xi\\ \\vert\\ \\xi\\in \\nu_{x_0}M, \\, \\langle \\xi, n_i\\rangle <1\\}.$$\nIf $y_0=x_0+\\xi(x_0)$ lies in the boundary of $C$ and $y_0\\not=0$, then the fiber of the\nfibration $M\\to M_\\xi$ is a generalized flag manifold with Weyl group $W_{y_0}$,\nthe isotropy subgroup of $W$ at $y_0$.\n\n For any parallel normal vector field $\\xi$ on $M$, the intersection of $M_{\\xi}$ with\n$x + \\nu_{x}M$ is an orbit of $W$. In particular, if $M_{\\xi}$\nis a parallel isoparametric submanifold, then it intersects each open Weyl chamber of\n$W$ exactly once. Moreover, $M_{\\xi}$ is a focal submanifold if and only if\n$ x+\\xi(x)$ is contained in $\\bigcup_{i=1}^{g} L_{i}$.\n\n\n\\subsection{Isoparametric map and $W$-invariant polynomials}\n\n\\hspace{100pt} \\\\\nGiven a $W$-invariant polynomial $f$ on $V=x_0+\\nu_{x_0}M$, there is a unique extension\n$\\Psi(f)$ on $\\mathbb{R} ^N$ such that $\\Psi(f)$ is constant along any parallel submanifold $M_\\xi$\nand $\\Psi(f)\\ \\vert\\ _{V}= f$. Moreover, $\\Psi(f)$ is also a polynomial.\n\nLet $\\tilde \\triangle$ and $\\triangle$ denote the Laplacian in $\\mathbb{R} ^N$ and $V$ respectively.\nThen by Lemma 3.2 of \\cite{Terng},\n\\begin{equation}\\label{eqn:mcfpoly}\nF(x)=\\tilde \\triangle \\Psi (f)(x)- \\Psi(\\triangle f)(x)= \\sum_{i=1}^g \\frac{m_i \\langle \\nabla f(x), {\\bf n}_i\\rangle}{\\langle x, {\\bf n}_i\\rangle}.\n\\end{equation}\nis a polynomial on $\\mathbb{R} ^N$ and is constant along parallel submanifolds of $M$.\nMoreover, if $f$ is a homogeneous $W$-invariant polynomial of degree $m$ on $V$,\nthen $F$ is a homogeneous polynomial of degree $m-2$ on $\\mathbb{R} ^N$.\n\n\\bigskip\n\\section{Mean curvature flows for general isoparametric submanifolds}\n\\label{sec:Basic}\n\nLet $M$ be an isoparametric submanifold of $\\mathbb{R} ^N$, fix $x_0\\in M$, and $W$ the Coxeter group associated to $M$.\nWe prove that the MCF stays isoparametric and the MCF equation becomes a flow equation of a vector field $H$ defined in the interior of the Weyl chamber of $W$ containing $x_0$ in $x_0+ \\nu_{x_0}M$ and the vector field $H$ tends to infinity at the boundary of the Weyl chamber. We prove that solutions of the ODE $x'= H(x)$ only exists for finite time. To see the finer structure of the behavior of the blow-up of MCF, we use $W$-invariant polynomials to construct a new coordinate system for the Weyl chamber so that the corresponding vector field $H$ becomes a polynomial vector field. We then analyse the behavior of flows of this polynomial vector field to obtain informations on the collapsing of the MCF.\n\nFix $x_{0} \\in M$. Since $\\nu M$\nis globally flat, we can identify a vector $v\\in \\nu_{x_0}M$ and the unique parallel normal field $\\hat v$\nalong $M$ defined by $\\hat v(x_0)=v$.\nLet ${\\bf n}_{i}$ be curvature normals of $M$ with multiplicity $m_{i}$ for $i=1, \\ldots, g$.\nWe may view ${\\bf n}_{i}$ either as a global parallel normal vector field along $M$ or\nan element in $\\nu_{x_0}M$. The precise meaning should be clear from the context.\n\nLet $\\xi(t) \\in \\nu_{x_0}M$ be a one parameter family of normal vectors satisfying the flow equation\n\\begin{equation} \\label{eqn:mcf}\n \\dot{\\xi}(t) = \\sum_{i=1}^{g} \\frac{m_{i}{\\bf n}_i}{1- \\langle\\xi(t), {\\bf n}_{i}\\rangle}, \\qquad\n \\xi(0) = 0.\n\\end{equation}\nIt follows from \\eqref{eqn:MCV} that $\\xi$ is a solution of \\eqref{eqn:mcf} if and only if\n the one parameter family of parallel submanifolds $M(t) := M_{\\xi(t)}$ satisfy the\n{\\it mean curvature flow} equation with $M(0) = M$.\nIn other words, the MCF preserves the isoparametric condition:\n\n\\begin{prop}\nIf $f:M\\times [0,T)\\to \\mathbb{R} ^N$ satisfies the mean curvature flow in $\\mathbb{R} ^N$ and $f(\\cdot, 0)$ is isoparametric,\nthen $f(\\cdot, t)$ is isoparametric for all $t\\in [0,T)$.\n\\end{prop}\n\nEquation \\eqref{eqn:mcf} does not make sense if $\\langle\\xi(t), {\\bf n}_{i}\\rangle = 1$ for some $i$.\nWe will only study the flow equation under the condition\n\\[ \\langle\\xi(t), {\\bf n}_{i}\\rangle \\,\\, < \\,\\, 1 \\]\nfor all $i=1, \\ldots, g$. In other words, we require that $x_{0} + \\xi(t)$ stays in the same\nWeyl chamber as $x_{0}$ for all $t$. Under this condition, all $M(t)$ are still isoparametric.\n\nNote that \\eqref{eqn:mcf} is a system of non-linear ODE given by a vector field defined on the\nWeyl chamber $C$ containing $x_0$ and the vector field blows up along the boundary of $C$.\nThe study of MCF with isoparametric submanifolds as initial data reduces to the study of\nsolutions of this ODE system.\n\n\\begin{thm} \\label{thm:radialMCV}\nLet $M\\subset S^{N-1}(r_0)$ be an $n$-dimensional isoparametric submanifold in $\\mathbb{R} ^N$, and $x_0\\in M$.\nIf $\\xi(t)$ satisfies the mean curvature flow equation \\eqref{eqn:mcf}, then $x(t)=x_0+\\xi(t)$\nsatisfies\n\\begin{equation}\\label{aa}\nx'(t)= -\\sum_{i=1}^g \\frac{m_i {\\bf n}_i}{\\langle x(t), {\\bf n}_i\\rangle},\n\\end{equation}\nwith $x(0)=x_0$.\nLet $H(t)$ be the mean curvature vector of $M(t)=M_{\\xi(t)}$ at the point $x(t)$. Then\n\\begin{enumerate}\n\\item[(a)] $\\langle x(t), \\, H(t)\\rangle = - n$,\n\\item[(b)] $|| x(t) ||^{2} = || x(0)||^{2} - 2nt$.\n\\end{enumerate}\n\\end{thm}\n\n\\begin{proof}\nBy equation \\eqref{eqn:centerO},\n$$ \\langle x(t), \\, {\\bf n}_{i}\\rangle =\\langle x(0), {\\bf n}_{i}\\rangle + \\langle \\xi(t), {\\bf n}_{i}\\rangle \\nonumber = -1+ \\langle \\xi(t), {\\bf n}_{i}\\rangle$$\nfor all $i=1, \\cdots, g$.\nSince\n\\begin{equation} \\label{eqn:MCVcenterO}\nH(t) = - \\sum_{i=1}^{g} \\frac{m_{i}{\\bf n}_i}{\\langle x(t), {\\bf n}_{i}\\rangle},\n\\end{equation}\nwe have $\\langle x(t), \\, H(t)\\rangle = - \\sum_{i=1}^{g} m_{i} = -n$.\nThis proves part (a). Part (b) follows from integrating the following formula\n\\[ \\frac{d}{dt} \\|x(t)\\|^{2}\n\\,\\, = \\,\\, 2 \\langle x(t), \\, x^{\\prime}(t)\\rangle\n\\,\\,=\\,\\, 2 \\langle x(t), \\, H(t)\\rangle\n\\,\\,= \\,\\, - 2n.\\]\n\\end{proof}\n\nHence we have\n\\begin{cor} \\label{cor:finiteInt}\nThe mean curvature flow in $\\mathbb{R} ^N$ with initial data an isoparametric submanifold in $S^{N-1}(r_0)$ exists only for finite time with maximal interval $[0, T)$, where\n$0k\/2$.\nTherefore $f(t)$ is a polynomial in $t$.\n\nLet $y(t)= (y_1(t), \\ldots, y_k(t))= P(x(t))$, and\n $$F_i(x)= \\sum_{i=1}^g m_{i} \\, \\frac{\\langle \\nabla P_i(x), {\\bf n}_i\\rangle}{\\langle x, {\\bf n}_i\\rangle}.$$\nBy equation \\eqref{eqn:mcfpoly}, $F_i$ is a $W$-invariant homogeneous polynomial on\n$V$ of degree $s_i-2$. Since $\\mathbb{R} [V]^W=\\mathbb{R} [P_1, \\ldots, P_k]$,\n $$F_i= - \\eta_i(P_1, \\ldots, P_{i-1})$$\nfor some polynomial $\\eta_i$. But we have shown above that $y_{i}'(t)= - F_i(x(t))$, so\n$$y_i'(t)= - F_{i}(x(t)) = \\eta_i(y_1(t), \\ldots, y_{i-1}(t)).$$\n\nThis shows that $y(t)$ is an integral curve of the polynomial vector field $\\eta$ on $\\mathbb{R} ^k$.\nSince $y_1(t)= ||x(0)||^2 -2nt$, solution $y$ can be solved explicitly by integrations.\n \\end{proof}\n\nThe MCF equation \\eqref{aa} is given by the vector field\n $$H(x)= -\\sum_{i=1}^g \\frac{m_i {\\bf n}_i}{\\langle x, {\\bf n}_i\\rangle},$$ which is smoothly defined\n on the Weyl chamber $C$ of $x_0+\\nu_{x_0}M$ and blows up at the boundary $\\partial C$.\n If we use generators of $W$-invariant polynomials on $x_0+\\nu_{x_0}M$ to change coordinates\n to $P$ as in Theorem \\ref{ba}, then the vector field $H$ becomes the polynomial vector field $\\eta$ on $P(C)$.\n Moreover, the flow of $\\eta$ can be solved explicitly and globally. Then apply $P^{-1}$\n to flows of $\\eta$ to get flows of \\eqref{aa}.\n\n\\begin{thm} \\label{thm:finiteconv}\nFor any compact isoparametric submanifold $M$ in $\\mathbb{R} ^N$, the mean curvature flow always converges\nto a focal submanifold at a finite time. Moreover, if $M_{1}$ and $M_{2}$ are parallel full\nisoparametric submanifolds which\n are contained in the same sphere, then mean curvature flows with initial data $M_1$ and $M_2$\n never intersect and they\n converge to two distinct focal submanifolds.\n\\end{thm}\n\n\\begin{proof}\n We use the same notation as Theorem \\ref{ba}.\nLet $x:[0, T)\\to C$ be the maximal interval for a solution of the\nmean curvature flow equation \\eqref{aa}. Note that\n$P_i(t)=P_{i}(x(t))$ are well defined since $P_{i}(t)$ are\npolynomials in $t$. Therefore the mean curvature flow of $x_{0}\n\\in M$ must converge to $P^{-1}(P_{1}(T), \\cdots, P_{k}(T))$ which\nlies on the boundary of $\\overline{C}$. The mean curvature flow of\n$M$ then converges to the focal submanifold passing through this\npoint.\n\nWe may assume that $x_i(0)$ lies in the unit sphere. Let $T_i$ denote the maximum time for the solution $x_i(t)$. If $T_1\\not= T_2$, then $||x_i(t)||^2= 1- 2n t$, so $\\lim_{t\\to T_1^-}||x_1(t)||^2\\not=\\lim_{t\\to T_2^-} ||x_2(t)||^2$. If $T_1=T_2= T$, then since $x_i(t)$ are solutions of \\eqref{aa} and $\\langle x_i(t), n_j\\rangle <0$, we have\n\\begin{equation}\\label{bp}\n\\frac{1}{2} \\frac{d}{dt}\\, \\big| ||x_1(t)-x_2(t)||^2= \\sum_{i=1}^g m_i\\, \\frac{\\langle x_1(t)-x_2(t), {\\bf n}_i\\rangle^2}{\\langle x_1(t), n_i\\rangle \\langle x_2(t), {\\bf n}_i\\rangle} \\geq 0.\n\\end{equation}\nThis implies that $||x_1(t)-x_2(t)||^2$ increases in $t\\in [0, T)$, hence\n$$\\lim_{t\\to T^-} x_1(t)\\not= \\lim_{t\\to T^-} x_2(t).$$\n\\end{proof}\n\n\n\\begin{thm}\\label{thm:focalconv}\nEvery focal submanifold is a limit of the mean curvature flow of certain isoparametric submanifold.\n\\end{thm}\n\nWe need a couple Lemmas.\nFirst a simple Lemma on scaling and the proof is obvious:\n\n\\begin{lem}\\label{ac}\nIf $f:M\\times [0,T)\\to \\mathbb{R} ^N$ is a solution to the mean curvature flow with $f(x,0)= f_0(x)$,\nthen given any $r\\not=0$, $\\tilde f(x,t)= rf(x, r^{-2}t)$ is a solution with $\\tilde f(x,0)= rf_0(x)$.\n\\end{lem}\n\n\\begin{lem} \\label{miniso}\nLet $f:M^{n} \\longrightarrow S^{N-1}(r_{0})$ be an immersed minimal submanifold of\na sphere with radius $r_{0}$.\nFor any $x \\in M$, the solution to the mean curvature flow equation in $\\mathbb{R}^{N}$\nwith initial data $M$ is given by\n$$F(x, t) = \\sqrt{1 - (2nt\/r_{0}^{2}) } \\, f(x).$$\nIn particularly, the mean curvature flow of $M$ shrinks to a point homothetically\n in finite time\n$T_{0} \\,\\, = \\,\\, r_{0}^{2}\/(2n)$.\n\\end{lem}\n\n\\begin{proof}\nFor minimal submanifolds of the sphere $S^{N-1}(r)$ with radius $r$,\nthe mean curvature vector at a point $x$ is $- n x\/r^{2}$.\nLet $F(x, t)= r(t) f(x)$ for $x \\in M$ with $r(t) \\geq 0$. Then the mean curvature\nvector field of $F(\\cdot, t)$ at point $x$ is given by\n$- \\frac{n}{ r_{0}^{2} r(t)} \\, f(x)$.\nSo $F(x, t)$ satisfies the mean curvature flow equation for $f$ if and only if\n\\[ r^{\\prime}(t) = - n\/( r_{0}^{2} r(t)) \\hspace{20pt}\n {\\rm and} \\hspace{20pt} r(0)=1.\\]\nIt follows that $ r(t) = \\sqrt{1 - (2nt\/r_{0}^{2})}$.\n\\end{proof}\n\n\\medskip\n\\noindent {\\bf Proof of Theorem \\ref{thm:focalconv}}\n\nIn each isoparametric family, there exists a unique isoparametric submanifold $M\\subset S^{N-1}(1)$,\nwhich is minimal in $S^{N-1}(1)$. Let $x_0\\in M$. The mean curvature flow for minimal\n submanifold in spheres can be solved explicitly as in Lemma \\ref{miniso}, i.e.,\n $x(t)= \\sqrt{1-2nt}\\, x_0$ is a solution of \\eqref{aa} and $x(t)\\in C$ for all\n $t\\in [0,\\frac{1}{2n})$.\n\n Recall that integral curves of $H(x)= -\\sum_{i=1}^g \\frac{m_i{\\bf n}_i}{\\langle x, {\\bf n}_i\\rangle}$ map to\n integral curves of the polynomial vector field $\\eta$ under the homeomorphism $P$ defined in\n Theorem \\ref{ba}. Since the integral curve starting from $x_0$ lies in $C$, the flow of $\\eta$\n starting at $P(x_0)$ lies in $P(C)$. But $-\\eta$ is a polynomial vector field and the one-parameter\n subgroup $\\phi_t$ generated by $-\\eta$ is a globally defined polynomial map. So there exists $\\delta>0$\n and an open subset $\\mathcal U$ of $P(\\bar C)$ such that $\\mathcal U$ contains the origin and\n $\\phi_t(z)\\in P(C)$ for $t\\in (0,\\delta)$ and\n $z\\in {\\mathcal U}$. This shows that the flow of $-\\eta$ starting at the boundary of $\\mathcal U$\n points inward in $P(C)$. Hence any boundary point of $P^{-1}({\\mathcal U})$ is a limit of some\n MCF with initial data in $C$. It follows from Lemma \\ref{ac} that any focal submanifold can be a\n limit of some MCF with some initial isoparametric submanifold.\n $\\Box$\n\n\nAs consequence of Theorem \\ref{thm:finiteconv} and Lemma \\ref{miniso}, we have\n\n\\begin{cor} \\label{cor:miniso}\nLet $M$ be an isoparametric submanifold.\nThe mean curvature flow of $M$ converges to a point if and only\nif it is minimal in the sphere containing it.\n\\end{cor}\n\nBelow we describe the rate of collapsing of the MCF for isoparametric submanifolds. Recall that a MCF, $M_t$, collapses at time $T<\\infty$ is said to have {\\it type I singularity\\\/} (cf. \\cite{W}) if there is a constant $c_0$ such that\n$$||{\\rm II\\\/}(t)||_\\infty^2(T-t)\\leq c_0$$\nfor all $t\\in [0,T)$, where $||{\\rm II\\\/}(t)||_\\infty$ is the sup norm of the second fundamental form for $M_t$.\n\n\\begin{prop}\\label{bm}\nLet $x(t)$ be a solution of the MCF \\eqref{aa}, and $T$ is the maxmial time. Then\n\\begin{enumerate}\n\\item $x(T):=\\lim_{t\\to T^-} x(t)$ exists and belong to the boundary $\\partial C$ of the Weyl chamber $C$,\n\\item $\\lim_{t\\to T^-} \\frac{||x(t)-x(T)||^2}{T-t} = 2m$, where $m=\\dim(M_{x(0)})-\\dim(M_{x(T)})$,\n\\item if $x(T)$ lies in a highest dimensional stratum of $\\partial C$, then the MCF has type I singularity.\n\\end{enumerate}\n\\end{prop}\n\n\\begin{proof}\nWe have proved (1) in Theorem \\ref{thm:finiteconv}. Statement (2) follows from the L'Hopital law:\n\\begin{align*}\n& \\lim_{t\\to T^-} \\frac{||x(t)-x(T)||^2}{T-t}\n=\\lim_{t\\to T^{-}} \\frac{2 \\langle x(t)-x(T), \\, x'(t)\\rangle}{-1} \\\\\n&= 2\\lim_{t\\to T^{-1}} \\sum_{i=1}^{g} \\langle x(t)-x(T), \\, \\frac{m_i {\\bf n}_i}{\\langle x(t), \\, {\\bf n}_i\\rangle}\\rangle\\\\\n&=2 \\lim_{t\\to T^{-1}} \\sum_{i\\not\\in I} m_i \\frac{\\langle x(t)-x(T), \\, {\\bf n}_i\\rangle}{\\langle x(t), \\, {\\bf n}_i\\rangle}\n + 2\\sum_{i\\in I} m_i\\frac{\\langle x(t), \\, {\\bf n}_i\\rangle }{\\langle x(t), \\, {\\bf n}_i\\rangle} = 2 \\sum_{i\\in I} m_i,\n\\end{align*}\nwhich is the dimension of the fiber of $M_{x(0)}\\to M_{x(T)}$.\nHere $I=\\{1\\leq i\\leq g\\ \\vert\\ \\langle x(T), \\, {\\bf n}_i\\rangle=0\\}$.\n\nWe now prove statement (3).\nIf $x(T)$ lies in a highest dimensional stratum of $\\partial C$, then there exists a unique $i$ such that $x(T)$ lies\nin the hyperplane defined by ${\\bf n}_i$, i.e., $\\langle x(T), {\\bf n}_i\\rangle=0$. We may assume $i=1$.\nNote that the norm square of the second fundamental form of $M_{x(t)}$ satisfies\n\\begin{align*}\n& ||{\\rm II\\\/}(x(t))||^2(T-t) \\leq \\sum_{i=1}^g \\frac{m_i ||{\\bf n}_i||^2}{\\langle x(t), {\\bf n}_i\\rangle^2}(T-t) \\\\\n&= \\frac{m_1||{\\bf n}_1||^2(T-t)}{\\langle x(t), {\\bf n}_1\\rangle^2}+\\sum_{i=2}^g \\frac{m_i||{\\bf n}_i||^2}{\\langle x(t), {\\bf n}_i\\rangle^2}(T-t).\n\\end{align*}\nAs $t\\to T^-$, the second term tends to zero because $\\langle x(T), {\\bf n}_i\\rangle\\not=0$ for\nall $i\\geq 2$, and by the l'Hopital law the first term has the same limit as\n$$\\frac{-m_1||{\\bf n}_1||^2}{-2\\langle x(t), {\\bf n}_1\\rangle \\sum_{i=1}^g m_i \\frac{\\langle{\\bf n}_i, {\\bf n}_1\\rangle}{\\langle x(t), {\\bf n}_i\\rangle}}.$$\nBut the denominator tends to $-2m_1||{\\bf n}_1||^2$, so the limit is $1\/2$.\n\\end{proof}\n\nWe remark that there is an open dense subset ${\\mathcal{O}}$ of the Weyl chamber $C$ such that the solution $x(t)$ of \\eqref{aa} with $x(0)\\in {\\mathcal{O}}$ converges to a point in a highest dimensional stratum of $\\partial C$.\n\n\n\\bigskip\n\\section{Solutions to the mean curvature flow equation}\n\\label{sec:solution}\n\nIn this section, we use Theorem \\ref{ba} to construct explicit solutions of the MCF \\eqref{aa} by\nselecting a set of generators $P_1, \\ldots, P_k$ for the $W$-invariant polynomials and calculating\nflows of the polynomial vector field $\\eta$.\n\nWe use the root system of the Coxeter group given in \\cite{GB}.\nLet $M$ be a compact, irreducible isoparametric submanifold in $\\mathbb{R} ^N$, $W$ its Weyl group, and\n${\\bf n}_i$ its curvature normals.\nLet $\\Pi$ denote a set of simple roots of $W$, and $\\Delta_{+}$ the set of positive roots\ndefined by $\\Pi$. Then $\\{\\mathbb{R} {\\bf n}_i\\ \\vert\\ 1\\leq i\\leq g\\}$ is equal to $\\{\\mathbb{R} \\alpha\\ \\vert\\ \\alpha\\in \\Delta_+\\}$. So\nthe Weyl chamber $C$ containing $x_0$ is precisely given by\n\\[ C = \\{ x \\in V \\mid\n -\\langle x, \\alpha\\rangle > 0 \\,\\,\\, {\\rm for \\,\\,\\, all} \\,\\,\\, \\alpha \\in \\Pi \\}. \\]\nThe closure of $C$ is\n\\[ \\overline{C} = \\{ x \\in V \\mid\n -\\langle x, \\alpha\\rangle \\geq 0 \\,\\,\\, {\\rm for \\,\\,\\, all} \\,\\,\\, \\alpha \\in \\Pi \\}, \\]\nand the MCF \\eqref{aa} becomes\n\\begin{equation}\\label{ab}\nx'(t)=-\\sum_{\\alpha\\in \\Delta^+} \\frac{m_\\alpha }{\\langle x(t),\\alpha\\rangle}\\,\\, \\alpha\n\\end{equation}\nwhere $m_{\\alpha}$ is the multiplicity of the curvature normal which is parallel to $\\alpha$.\nSince \\eqref{ab} is invariant under re-scaling of each $\\alpha$, we may\nnormalize roots of the Coxeter group to be of {\\it unit length}.\n\nIf $M = M_{1} \\times M_{2}$ with $M_{i}$ an isoparametric submanifold of $\\mathbb{R}^{N_{i}}$\nfor $i=1, 2$, then the Weyl group of $M$ is the product of the Weyl groups of $M_1$ and $M_2$\nand the mean curvature flow of $M$ is the product\nof the mean curvature flows of $M_{1}$ and $M_{2}$.\n So without loss of generality,\nwe may assume that $M$ is an irreducible isoparametric submanifold.\nIn the rest of this section, we work out explicit solutions for mean curvature flow equations\nfor compact isoparametric submanifolds whose Coxeter group are $A_{k}$, $B_{k}$,\n$D_{k}$ and $G_{2}$.\n\n\\begin{eg} {\\bf The $A_{k}$ case}\\par\n\n\\smallskip\n\n\nSuppose that $k \\geq 2$.\nLet $\\{ e_{1}, \\cdots e_{k+1} \\}$ be the standard orthonormal basis of $\\mathbb{R}^{k+1}$\nand $(x_{1}, \\cdots , x_{k+1})$ the corresponding coordinate.\nThe set\n$\\frac{1}{\\sqrt{2}} \\left(e_{i} - e_{i+1} \\right)$ with $1 \\leq i \\leq k$ is a simple root\nsystem of $A_k$, and\nthe set of positive roots is\n$\\frac{1}{\\sqrt{2}} \\left(e_{i} - e_{j} \\right)$ with $1 \\leq iq} \\frac{1}{x_{i_{q}}(x_{i_{p}}-x_{i_{q}})}\n + \\sum_{pq} \\frac{1}{x_{i_{p}}-x_{i_{q}}} \\left(\\frac{1}{x_{i_{q}}} - \\frac{1}{x_{i_{p}}} \\right) \\\\\n&=& \\sum_{p>q} \\frac{1}{x_{i_{p}} x_{i_{q}}}.\n\\end{eqnarray*}\nHence by equation \\eqref{eqn:dersrA}\n\\begin{eqnarray*}\n \\frac{r!}{m} \\frac{d}{dt} y_{r}\n&=& \\sum_{i_{1} \\neq \\cdots \\neq i_{r}} \\,\\, \\, \\sum_{1 \\leq p, q \\leq r, \\, p > q}\n x_{i_{1}} \\cdots \\,\\, \\widehat{x_{i_{q}}} \\,\\, \\cdots \\,\\,\n \\widehat{x_{i_{p}}} \\,\\,\\cdots x_{i_{r}} \\\\\n&=& \\frac{1}{2} r(r-1) (k-r+3)(k-r+2) (r-2)! y_{r-2}\n\\end{eqnarray*}\nThis proves the claim.\n\nThe explicit formula for $y_{r}(t)$ can be obtained from \\eqref{ak1}\nrecursively, and it is a polynomial\nin $t$ and initial conditions $x_{1}(0), \\cdots, x_{k+1}(0)$.\nFor each $t$, we can obtain $x_{1}(t), \\cdots, x_{k+1}(t)$ as the $k+1$ solutions\nof the following polynomial equation in $z$:\n\\begin{equation} \\label{eqn:rootssymAk}\n \\sum_{r=0}^{k+1} \\,\\, (-1)^{k+1-r} \\,\\, y_{k+1-r}(t) \\, z^{r} = 0,\n\\end{equation}\nwith the property\n\\[ x_{1}(t) < x_{2}(t) < \\cdots < x_{k+1}(t). \\]\n\\end{eg}\n\n\\begin{eg} \\label{sec:Bk} {\\bf The $B_{k}$ case}\\par\n\n\\smallskip\nSuppose that $k \\geq 2$.\nLet $\\{ e_{1}, \\cdots e_{k} \\}$ be the standard orthonormal basis of $\\mathbb{R}^{k}$\nand $(x_{1}, \\cdots , x_{k})$ the corresponding coordinate.\nWe identify $\\mathbb{R}^{k}$ with a normal space of an isoparametric submanifold of type $B_{k}$.\nThe set $e_{k}$ and\n$\\frac{1}{\\sqrt{2}} \\left(e_{i} - e_{i+1} \\right)$ with $1 \\leq i \\leq k-1$ is a simple root system of $B_k$, and the set of positive roots are $e_{i}$ with $1 \\leq i \\leq k$ and\n$\\frac{1}{\\sqrt{2}} \\left(e_{i} \\pm e_{j} \\right)$ with $1 \\leq i p }\n \\frac{( y_{i_{1}} \\cdots \\widehat{y_{i_{l}}} \\cdots y_{i_{j}})y_{i_{l}}}{y_{i_{l}}-y_{p}}\n + \\sum_{ i_{1} \\neq \\cdots \\neq i_{j}, i_{l} < p}\n \\frac{ (y_{i_{1}} \\cdots \\widehat{ y_{i_{l}}} \\cdots y_{i_{j}})y_{i_{l}}}{y_{i_{l}}-y_{p}}.\n\\]\nSwitching the indices $i_{l}$ and $p$ in the second term and adding to the first term, we have\n\\begin{eqnarray*}\n\\sum_{i_{1} \\neq \\cdots \\neq i_{j} \\neq p}\n \\frac{y_{i_{1}} \\cdots y_{i_{j}}}{y_{i_{l}}-y_{p}}\n&=& \\sum_{\\begin{array}{c} i_{1} \\neq \\cdots \\neq i_{j} \\neq p \\\\ i_{l} > p \\end{array}}\n y_{i_{1}} \\cdots \\widehat{y_{i_{l}}} \\cdots y_{i_{j}}\n\\end{eqnarray*}\nTherefore the second term on the right hand side of equation \\eqref{eqn:dersigma} is\n\\[ -2 m_{1}(k-j+1)(k-j) s_{j-1}.\\]\nThis proves the claim.\n\nNote that explicit formula for $\\zeta_{i}(t)$ can be obtained from \\eqref{bk1} recursively, and\nit is always a\ndegree $i$ polynomial\nin $t$ and initial conditions $y_{1}(0), \\cdots, y_{k}(0)$.\n Let $y_{1}(t), \\cdots y_{k}(t)$\nbe the $k$ roots of\n\\begin{equation} \\label{eqn:rootssymBk}\n \\sum_{r=0}^{k} \\,\\, (-1)^{k-r} \\,\\, s_{k-r}(t) \\, z^{r} = 0,\n\\end{equation}\nwith the property $y_{1}(t) > y_{2}(t) > \\cdots > y_{k}(t) > 0$.\nThen $x_{i}(t) = - \\sqrt{y_{i}(t)}$ for $i=1, \\cdots k$.\n\\end{eg}\n\n\\begin{eg} \\label{sec:Dk} {The $D_{k}$ case}\\par\n\n\\smallskip\nSuppose that $k \\geq 4$.\nLet $\\{ e_{1}, \\cdots e_{k} \\}$ be the standard orthonormal basis of $\\mathbb{R}^{k}$\nand $(x_{1}, \\cdots , x_{k})$ the corresponding coordinate.\nWe identify $\\mathbb{R}^{k}$ with a normal space of an isoparametric submanifold of type $D_{k}$.\nThe set of simple roots are $\\frac{1}{\\sqrt{2}} \\left(e_{k-1} + e_{k} \\right)$ and\n$\\frac{1}{\\sqrt{2}} \\left(e_{i} - e_{i+1} \\right)$ with $1 \\leq i \\leq k-1$, and the set\nof positive roots is\n$\\{\\frac{1}{\\sqrt{2}} \\left(e_{i} \\pm e_{j} \\right)\\ \\vert\\ 1 \\leq i0\\},$$ then $\\triangle^+(x_1)=\\triangle^+(x_2)$ if and only if\n$x_1, x_2$ lie in the same stratum $\\sigma$, and will be denoted by $\\triangle^+(\\sigma)$,\n\\item $\\sigma$ is the Weyl chamber of $W_x$ for $x\\in \\sigma$ and $\\sigma$ is an\nopen simplicial cone in the following linear subspace\n$$V(\\sigma)=\\{x\\in p+\\nu_pM_0\\ \\vert\\ \\langle x, \\alpha\\rangle\n=0, \\,\\, {\\rm for\\, all\\,} \\, \\alpha\\in \\triangle^+\\setminus \\triangle^+(\\sigma)\\},$$\n\\end{enumerate}\n\nLet $\\sigma$ be a stratum in $\\partial C$, $x_0\\in \\sigma$, and $M$ the focal submanifold of\n$M_0$ through $x_0$.\nBy \\cite[Theorem 4.1]{Terng},\nthe mean curvature vector field of $M$ at $x_0$ is given by\n\\begin{equation} \\label{eqn:focalMCV}\nH(x_0) = - \\sum_{\\alpha \\in \\Delta^{+}(\\sigma)} \\frac{m_{\\alpha}}{\\langle x_0, \\alpha\\rangle} \\,\\, \\alpha\n\\end{equation}\nwhere $m_{\\alpha}$ are multiplicities of curvature spheres of $M_{0}$.\nMoreover\n\\begin{equation} \\label{eqn:focalxH}\n \\langle x_0, \\alpha\\rangle \\,\\, =\\,\\, \\langle H(x_0), \\alpha\\rangle \\,\\,= \\,\\,0\n\\end{equation}\nfor all $\\alpha \\in \\Delta^{+} \\setminus \\Delta^{+}(\\sigma)$.\nThe mean curvature flow equation of $M$ is the following ODE on $\\sigma$:\n\\begin{equation} \\label{eqn:focalMCF}\n\\frac{dx}{dt} = - \\sum_{\\alpha \\in \\Delta^{+}(\\sigma)}\n\\frac{m_{\\alpha}}{\\langle x, \\alpha\\rangle} \\,\\, \\alpha.\n\\end{equation}\n\nThe analogue of Theorem \\ref{thm:radialMCV} also holds for this case. In particular,\nif $x(t)$ satisfies the flow equation \\eqref{eqn:focalMCF} then\n\\begin{equation}\n \\|x(t)\\|^{2} = \\|x(0)\\|^{2} - 2nt\n\\end{equation}\nwhere $n = \\sum_{\\alpha \\in \\Delta^{+}(\\sigma)} m_{\\alpha}$ is the dimension of $M$.\nTherefore we have\n\\begin{thm}\nThe maximal interval for the solution of the mean curvature flow equation for any focal submanifold\nis finite.\n\\end{thm}\n\nSuppose $x(t)$ and $y(t)$ satisfy equation \\eqref{eqn:focalMCF} on $\\sigma$ and $x(0)\\neq y(0)$. Use the same computation for \\eqref{bp} to get\n\\begin{align}\\label{bd}\n\\frac{1}{2} \\frac{d}{dt} \\|x(t) - y(t) \\|^{2}\n&= \\langle x(t) - y(t), \\,\\, x^{\\prime}(t) - y^{\\prime}(t)\\rangle \\nonumber \\\\\n&= \\sum_{\\alpha\\in \\triangle_+(\\sigma)}\n m_{\\alpha}\\frac{\\langle x(t)-y(t), \\alpha \\rangle^2}{\\langle x(t), \\alpha \\rangle \\langle y(t), \\alpha \\rangle}\\, > 0.\n \\end{align}\nThen proofs given in section \\ref{sec:Basic} works, so we have\n\n\\begin{thm} \\label{bc}\nLet $M^n\\subset S^{n+k-1}$ be a compact isoparametric submanifold in $\\mathbb{R} ^{n+k}$, $W$ its Weyl group,\n$C$ the Weyl chamber in $x_0+\\nu(M)_{x_0}$ containing $x_0\\in M$, and $M_y$ the submanifold parallel\nto $M$ through $y$. If $\\sigma \\subset \\overline{C}$ is a stratum, then\n\\begin{enumerate}\n\\item there is a unique $y_\\sigma\\in \\sigma$ such that the focal submanifold $M_{y_\\sigma}$ is minimal\nin $S^{n+k-1}$, and the MCF in $\\mathbb{R} ^{n+k}$ with initial data $M_{y_{\\sigma}}$ homothetically shrinks\nto a point,\n\\item if $y_0\\in \\sigma \\bigcap S^{n+k-1}-\\{y_\\sigma\\}$, then the MCF in $\\mathbb{R} ^{n+k}$ with $M_{y_0}$\nas initial data blows up in finite time $T < \\frac{1}{2n}$, $x(t)\\in \\sigma$ for all $t\\in [0, T)$,\nand\n$\\lim_{t\\to T^-} x(t)\\,\\in \\partial \\sigma$, in particular, the limit is a focal submanifold with\nlower dimension,\n\\item if $y_1, y_2\\in \\sigma \\bigcap S^{n+k-1}$ are distinct, then\nthe MCF in $\\mathbb{R} ^{n+k}$ with initial data $M_{y_1}$ and $M_{y_2}$ converge to\ndistinct focal submanifolds of lower dimensions.\n\\end{enumerate}\n\\end{thm}\n\n\\bigskip\n\n\\section{Mean curvature flows for isoparametric submanifolds in spheres}\n\\label{sec:sphere}\n\nIf $M^{n}\\subset S^{n+k-1}$ is an isoparametric submanifold in $\\mathbb{R} ^{n+k}$, then\n$M$ is also isoparametric\nin $\\mathbb{R}^{n+k}$. So basic structure theory for isoparametric submanifolds in Euclidean\nspaces also applies to $M$.\nFor $x \\in M$, let $H(x)$ and $H_{E}(x)$ be the mean curvature vector fields of $M$ at $x$\nas a submanifold of $S^{n+k-1}$ and $\\mathbb{R}^{n+k}$ respectively.\nThen $H(x)$ is the orthogonal projection of $H_{E}(x)$ to $T_{x}S^{n+k-1}$. More precisely\n\\[ H(x) = H_{E}(x) + n x \\]\nfor all $x \\in M$. In particular, $H$ is again a parallel normal vector field along $M$.\nThe mean curvature flow of $M$ as a submanifold of $S^{n+k-1}$ behaves similarly\nto its flow as a submanifold of $\\mathbb{R}^{n+k}$. With slight modifications, most results\nfor mean curvature flows for isoparametric submanifolds in the Euclidean spaces also hold\nfor isoparametric submanifolds in spheres. We only need to explain how to deal with\nthe arguments in the Euclidean case which can not be applied directly to the spherical case.\n\nFix $x_{0} \\in M$ and let $V = x_0+ \\nu_{x_{0}}M$ be\nthe normal space of $M$ as a submanifold of $\\mathbb{R}^{N}$ at the point $x_{0}$, $W$ its\nCoxeter group, and $C\\subset V$ the Weyl chamber containing $x_0$.\nThe mean curvature flow of $M$ in $S^{n+k-1}$ is uniquely determined by the flow of $x_{0}$ in\n$S := C \\bigcap S^{k-1}$:\n\\begin{equation} \\label{be}\nx'(t)= -\\sum_{\\alpha\\in \\triangle_+} \\frac{m_\\alpha \\alpha}{\\langle x(t),\\alpha\\rangle} \\, + n x(t).\n\\end{equation}\nThe set $S$ is a geodesic $(k-1)$-simplex on $S^{k-1}$.\nLet $x(t) \\in S$ be a solution to equation \\eqref{be} with initial condition $x_{0}$.\nThen\n\\[ y(t) = \\sqrt{1-2nt} \\,\\,\\,\\, x\\left(- \\frac{1}{2n} \\log(1-2nt)\\right) \\]\nsatisfies the Euclidean mean curvature flow equation \\eqref{ab} with initial condition\n$y(0) = x_{0}$.\nLet $[0, T_{x})$ and $[0, T_{y})$ be the maximal intervals for the domains of $x(t)$ and $y(t)$\nrespectively. Then\n\\[ T_{x} = - \\frac{1}{2n} \\log(1-2n T_{y}) \\]\nand\n\\[ \\lim_{t \\rightarrow T_{y}^{-}} \\,\\, y(t)\n \\,\\,=\\,\\, \\sqrt{1-2n T_{y}}\\,\\, \\lim_{t \\rightarrow T_{x}^{-}} \\,\\, x(t). \\]\nNote that by Theorems \\ref{thm:radialMCV} and Corollary \\ref{cor:miniso},\n$T_{y} \\leq \\frac{1}{2n}$ and the equality holds if and only if\nthe isoparametric submanifold $M_{0}$ passing $x_{0}$ is minimal in the sphere $S^{n+k-1}$.\nSo by Theorem \\ref{thm:finiteconv}, if $M_{0}$ is not minimal in the sphere, then\n$x(t)$ converges to a focal submanifold at a finite time $T_{x}$.\nThis proves Theorem \\ref{thm:MCFisopS}.\n\n\nIf $x_{1}(t) \\in S$ and $x_{2}(t) \\in S$ satisfy the spherical mean curvature flow equation\n\\eqref{be}, then\n\\begin{eqnarray*}\n\\frac{1}{2} \\frac{d}{dt} \\|x_{1}(t) - x_{2}(t)\\|^{2}\n&=& \\langle x_{1}-x_{2}, \\left(H_{E}(x_{1}) + n x_{1} \\right)\n- \\left(H_{E}(x_{2}) + n x_{2} \\right)\\rangle \\\\\n&=& n \\|x_{1}-x_{2}\\|^{2} + \\langle x_{1}-x_{2}, H_{E}(x_{1}) - H_{E}(x_{2})\\rangle.\n\\end{eqnarray*}\nBy \\eqref{bp}, $\\langle x_{1}-x_{2}, H_{E}(x_{1}) - H_{E}(x_{2})\\rangle \\,\\,\\geq \\,\\, 0$.\nTherefore\n\\begin{equation} \\label{eqn:departS}\n \\frac{d}{dt} \\|x_{1}(t) - x_{2}(t)\\|^{2} \\,\\,\\geq \\,\\, 2n \\|x_{1}(t)-x_{2}(t)\\|^{2}.\n\\end{equation}\nWe use \\eqref{eqn:departS} to give an estimate of the maximal interval $[0, T)$\nfor the spherical mean curvature flow $x(t)$.\nLet $p_{0}$ be the unique point in $S$ such that the isoparametric submanifold\npassing $p_{0}$ is minimal in the sphere $S^{n+k-1}$.\nSet $x_{1}(t) = x(t)$ and $x_{2}(t) = p_{0}$ in equation \\eqref{eqn:departS}. Since $x_2(t)$ exists for all $t>0$, we obtain\n\\[ \\|x(t) - p_{0} \\| \\geq e^{nt} \\|x(0) - p_{0} \\| \\]\n for all $t$ as long as $x(t)\\in S$.\nLet $D$ be the diameter of $S$, then $D \\leq 2$ and\n\\[ T \\leq \\frac{1}{n} \\log \\frac{D}{\\|x(0) - p_{0} \\|}.\\]\n\nNow we discuss the behavior of invariant polynomials under the spherical mean curvature flow.\nLet $x(t) \\in S$ be the mean curvature flow of $x_{0}$.\nFor any function $f$ on $V$, let $f(t) = f(x(t))$. Then\n\\[ f^{\\prime}(t) \\,\\,=\\,\\, \\langle \\nabla f(x(t)), H(x(t))\\rangle \\,\\,=\\,\\,\n \\langle \\nabla f(x(t)), H_{E}(x(t))\\rangle + n \\langle \\nabla f(x(t)), x(t)\\rangle. \\]\nIf $f$ is a homogenous polynomial of degree $k$ which is invariant under the action\nof the Coxeter group $W$, then as in the proof of Theorem \\ref{thm:finiteconv},\n\\begin{equation} \\label{eqn:invpolES}\n f^{\\prime}(t) \\,\\,=\\,\\, - F(x(t)) + nk f(t)\n\\end{equation}\nwhere $F$ is defined by equation \\eqref{eqn:mcfpoly}\nand it is an invariant polynomial of degree $k-2$.\nIf we have computed $F(t) := F(x(t))$, then we can solve $f(t)$ from equation \\eqref{eqn:invpolES}\nand obtain\n\\begin{equation} \\label{eqn:intinvpolS}\n f(t) = - e^{knt} \\int e^{-knt} F(t) \\, dt.\n\\end{equation}\nNote that there is no homogeneous invariant polynomial of degree 1. By induction on the degree,\nwe obtain the following\n\\begin{thm}\nIf $x(t)$ satisfies the spherical mean curvature flow equation \\eqref{be} and $f$ is a\n$W$-invariant polynomial, then $f(t) = f(x(t))= c_1 e^{knt} +h(t)$ for some constant $c_1$ and polynomial $h$.\n\\end{thm}\n\nIn particular $f(t)$ is well defined for all $t \\in \\mathbb{R}$.\nIn section \\ref{sec:solution}, we have given explicit formulas for\n$F_i$ for invariant homogeneous polynomials $P_i$ for isoparametric submanifolds.\nWe can use these formula and \\eqref{eqn:intinvpolS} to construct explicit\nsolutions to the spherical mean curvature flow equation for isoparametric submanifolds in spheres.\n\n\\begin{eg} {\\bf Phase portrait for rank $2$ cases}\\par\n\nLet $M^n\\subset S^{n+1}\\subset \\mathbb{R} ^{n+2}$ be an isoparametric hypersurface with $g$ distinct principal curvatures. Then the Weyl group associated to $M$ as a rank $2$ isoparametric submanifold in $\\mathbb{R} ^{n+2}$ is the dihedral group of $2g$ elements. Let $C$ denote the Weyl chamber containing $x_0\\in M$, and $D$ the intersection of $C$ and the normal circle at $x_0$ in $S^{n+1}$. Let $p_1, p_2$ denote the end points of $D$. The arc $D=\\widehat{p_1p_2}$ has length $\\pi\/g$. For $y\\in \\bar C$, let $M_y$ denote the submanifold through $y$ that is parallel to $M$ (a leaf of the isoparametric foliation). There exists a unique $p_0\\in D$ such that $M_{p_0}$ is minimal in $S^{n+1}$.\n\\begin{enumerate}\n\\item The spherical MCF \\eqref{be} has three orbits: a stationary point $p_0$, the orbit $\\widehat{p_0p_1}$ with one end tends to $p_0$ and the other end tends to $p_1$, and the orbit $\\widehat{p_0p_2}$ with one end tends to $p_0$ and the other end tends to $p_2$.\n\\item The MCF \\eqref{aa} in $\\mathbb{R} ^{n+2}$ starting at $M_y$ degenerates homothetically to one point (the origin) if $y=p_0$, to $M_{rp_2}$ for some $0