1. Introduction
1.1 Problem statement and background
Consider the 2D incompressible MHD equations on $\mathbb {T}\times \mathbb {R}$:
Here $\mathbb {T}$ is the periodized interval $[0,1]$, $\tilde u=\tilde u(t, x, y): [0,\infty )\times \mathbb {T}\times \mathbb {R} \to \mathbb {R}^2$ denotes the velocity, $\tilde b=\tilde b(t,x,y): [0,\infty )\times \mathbb {T}\times \mathbb {R} \to \mathbb {R}^2$ denotes magnetic field, $\tilde p=\tilde p(t, x, y): [0,\infty )\times \mathbb {T}\times \mathbb {R} \to \mathbb {R}$ denotes pressure, $\nu$ denotes the inverse Reynolds number, and $\mu$ denotes the inverse magnetic Reynolds number. For a derivation of (1.1) and a general overview of MHD equations we refer the reader to [Reference Davidson17, Reference Lifschitz30].
Note that the shear profile $u_s=( e^{\nu t \partial _{yy}}U(y),0) ^\top$ is a solution of (1.1) in any uniform magnetic field $b_s=(\alpha,0 )^\top$, where $U(y)$ is a given smooth function, and $\alpha$ is a constant in $\mathbb {R}$. A natural question is to study the long time stability of the equilibrium state $(u_s, b_s)$. To this end, let us introduce the perturbations ${u}$ and ${b}$ by $\tilde {u}={u}+u_s$ and $\tilde {b}={b}+b_s$. Then $({u},{b})$ solves
where we have used the notations ${u}=( {u}_1,{u}_2)^\top,~{b}=( {b}_1,{b}_2)^\top$ and $\bar {U}=e^{\nu t \partial _{yy}}U(y)$. The corresponding perturbed vorticity and current density take the form of
Let $\psi$ and $\phi$ be the steam functions such that
Then we infer from (1.2) that the system of $( \omega,j)$ is of the following form
The mathematical theory of the stability of shear flows gained quite some attention in the last decade. The nonlinear stability of the 2D Couette flow on $\mathbb {T}\times \mathbb {R}$ in the Euler equation was first obtained by Bedrossian and Masmoudi in [Reference Bedrossian and Masmoudi6] when the initial perturbation is smoother than the Gevrey space of class 2. In particular, the inviscid damping, an important hydrodynamic phenomenon manifests itself as the algebraic decay of the velocity for the inviscid fluids, was rigorously justified at the nonlinear level in [Reference Bedrossian and Masmoudi6]. A generalization of the results in [Reference Bedrossian and Masmoudi6] to the finite channel $\mathbb {T}\times [0,1]$ for initial perturbation with compact support was given by Ionescu and Jia in [Reference Ionescu and Jia22]. The nonlinear inviscid damping for a class of monotone shear flows was proved in [Reference Ionescu and Jia23] and [Reference Masmoudi and Zhao33], independently. The stability results for the 2D Couette flow in inhomogeneous fluids can be found in [Reference Chen, Wei, Zhang and Zhang16, Reference Zhao51]. The instability result for the 2D Couette flow on $\mathbb {T}\times \mathbb {R}$ was shown by Deng and Masmoudi in [Reference Deng and Masmoudi18] when the initial perturbation is less smooth than the Gevrey space of class 2. For more general shear flows, the stability/instability results are more difficult to achieve due to the presence of nonlocal term. We refer to [Reference Chen, Wei and Zhang15, Reference Jia25, Reference Jia26, Reference Wei, Zhang and Zhao45, Reference Zillinger52] for the linear inviscid damping results for general monotone shear flows, and to [Reference Ionescu, Iyer and Jia24, Reference Wei, Zhang and Zhao46, Reference Wei, Zhang and Zhao47] for the linear stability results for non-monotone shear flows.
If the viscosity is taken into consideration, consider the toy model
it is not difficult to show that the non-zero mode $g_{\neq }$ undergo enhanced dissipation, namely for some $c>0$ there holds
The inviscid damping mentioned above and the enhanced dissipation demonstrated in (1.4) are two main stability mechanisms which are closely related to the stability threshold problem. That is, how small should initial perturbations be in terms of the viscous coefficient to ensure nonlinear stability? Since the early experiments of Reynolds [Reference Reynolds36], the the linear stability or instability of shear flows at high Reynolds number has been a classical problem in applied fluid mechanics [Reference Case10, Reference Kelvin27, Reference Orr35, Reference Schmid and Henningson39, Reference Yaglom48]. Significant progress has been made by Bedrossian, Germain and Masmoudi in [Reference Bedrossian, Germain and Masmoudi3–Reference Bedrossian, Germain and Masmoudi5] for the nonlinear stability of the Couette flow on $\mathbb {T}\times \mathbb {R}\times \mathbb {T}$ by using the time dependent Fourier multiplier method. Then Wei and Zhang [Reference Wei and Zhang44] proved the nonlinear stability of the Couette flow on $\mathbb {T}\times \mathbb {R}\times \mathbb {T}$ as long as the initial perturbation satisfying $\|u_{\mathrm {in}}\|_{H^2}\ll \mathrm {Re}^{-1}$ with $\mathrm {Re}$ the Reynolds number. For the 2D the domain $\mathbb {T}\times \mathbb {R}$, a series of results can be found in [Reference Bedrossian, Masmoudi and Vicol7, Reference Bedrossian, Vicol and Wang8, Reference Masmoudi and Zhao34]. Recently, Li, Masmoudi and the first author of this paper [Reference Li, Masmoudi and Zhao29] studied the relationship between the size and the regularity of the initial perturbation that ensures the nonlinear asymptotic stability. For the periodic finite channels, a delicate resolvent estimate method was developed in [Reference Chen, Li, Wei and Zhang12, Reference Chen, Wei and Zhang14]. In particular, Chen, Wei and Zhang showed that the size of the initial perturbation in [Reference Wei and Zhang44] still ensures the stability of the Couette flow in the presence of physical boundary in [Reference Chen, Wei and Zhang14].
In the presence of magnetic field, the behaviours of the shear flows become more complicated. On the one hand, it is classically known that a strong background magnetic field can have a stability effect on a conducting fluid, see [Reference Chandrasekhar11, Reference Shivamoggi and Debnath40, Reference Velikhov42], for instance. On the other hand, the magnetic field may destabilize the system [Reference Chen and Morrison13] even with shear flows (including Couette flow) that are asymptotically stable. Tataronis and Grossmam [Reference Tataronis and Grossmann41] predicted that the decaying of the vertical components of velocity and magnetic field due to the phase mixing. Ren and Zhao [Reference Ren and Zhao38] gave a rigorous mathematical proof under the assumption that the magnetic field is positive and strictly monotone. In [Reference Hirota, Tatsuno and Yoshida20], Hirota, Tatsuno and Yoshida investigated the linearized behaviour of the ideal MHD equation around Couette flow $(k_fy, 0)$ and linear magnetic field $(k_my,0)$. They predicted that if $|k_f|<|k_m|$, then the magnetic island appears in the final state, namely the linear asymptotic stability fails, and if $|k_m|<|k_f|$, then linear damping holds and the magnetic island will be destructed. Recently, the generation of magnetic island was rigorously proved by Zhai, Zhang and Zhao in [Reference Zhai, Zhang and Zhao49], and the rigorously mathematical proof for the destruction of magnetic fields was given by Ren, Wei and Zhang in [Reference Ren, Wei and Zhang37]. We refer to [Reference Knobel and Zillinger28, Reference Liu, Masmoudi, Zhai and Zhao32] for more recent results for the linear stability results on shear flows in magnetic field. For the nonlinear stability result on this direction, Liss [Reference Liss31] proved that for strong and suitably oriented background fields $\alpha (\sigma,0,1)^\top$, the Couette flow $(y, 0,0)^\top$ is asymptotically stable provided the initial perturbations $u_{\mathrm {in}}$ and $b_{\mathrm {in}}$ satisfy
with $N$ sufficiently large.
In this paper, under the same condition $\nu =\mu$ in [Reference Liss31], we study the nonlinear stability of the time dependent shear flow $u_s=( \bar {U}(t,y),0)$ in a uniform magnetic field $b_s=(\alpha,0)^\top$ on $\mathbb {T}\times \mathbb {R}$. More precisely, for $0<\nu =\mu \ll 1$, given an initial norm $X_i$ and a final norm $X_f$, our goal in this paper is determine a constant $\gamma =\gamma (X_i, X_f)\ge 0$ as small as possible, such that if the initial perturbations $\bar {u}_{\mathrm {in}}$ and $\bar {b}_{\mathrm {in}}$ satisfy
for $c_0$ sufficiently small (independent of $\nu$), then the solution of (1.1) is global in time and converges back to $( u_s,b_s )$ as $t \to \infty$ in the sense that
1.2 The main result
Our main result is stated as follows.
Theorem 1.1 Let $N>1,$ $\mu =\nu \in (0,1]$. There exist two sufficiently large constants $\alpha _0>0$ and $C \ge 1$ independent of $\nu,$ such that for all $|\alpha |\ge \alpha _0,$ if the shear flow $U=U(y)$ satisfies
with $\delta$ independent of $\nu$, and the initial perturbation obeys
for any fixed $\tilde \delta >0$, then the global in time solution $(\omega,j)$ to (1.3) obeys
and the enhanced dissipation estimate
Remark 1.2 The bounds (1.10) and (1.11) follow from (4.3), (4.5) in theorem 4.1 and lemma A.1 immediately. So the rest part of this paper aims to prove theorem 4.1.
Remark 1.3 It is not difficult to obtain the explicit enhanced dissipation decay $e^{-c\nu ^\frac {1}{3}t}$ for $(\omega _{\ne }, j_{\ne })$ like (1.4). In fact, we just need to change the multiplier $\tilde K$ (see (4.2)) slightly. Let ${\bf e}_{\ne }$ be a multiplier defined by
Set
Replacing the multiplier $\tilde K$ in (4.10) by $\mathsf {M}$, the proof of proposition 4.3 is still valid. In particular, one can use (3.21) to absorb the bad term $c\nu ^\frac {1}{3}\left \|\mathsf {M}Z^{\pm }_{\ne }\right \|_{L^2}^2$ arising from the time evolution of ${\bf e}_{\ne }$ for sufficiently small $c$ (see § 2.3 for the definition of $Z^{\pm }$). We refer to [Reference Zhang and Zi50] for more details of the use of the multiplier ${\bf e}_{\ne }$ to get the explicit enhanced dissipation decay $e^{-c\nu ^\frac {1}{3}t}$.
Remark 1.4 At first glance, it looks as if the loss $\nu ^{-\frac {1}{2}}$ on the right hand side of (1.11) might come from the diffusion $-\nu \Delta (\omega,j)$. As a matter of fact, roughly speaking, due to the linear growth $\langle t \rangle$ caused by the stretch $2\bar {U}'\partial _{xy}\phi$ (together with the oscillation stemming from the strong background magnetic field $(\alpha,0)^{\top }$) and the enhanced dissipation $e^{-c\nu ^\frac {1}{3}t}$ as mentioned in remark 1.3, $(\omega _{\ne }, j_{\ne })$ behaves like $\langle t\rangle e^{-c\nu ^\frac {1}{3}t}$. Clearly, $\left \|\langle t\rangle e^{-c\nu ^\frac {1}{3}t}\right \|_{L^2_t}\approx \nu ^{-\frac {1}{2}}$. This explains the loss $\nu ^{-\frac {1}{2}}$ on the right hand side of (1.11).
Remark 1.5 Compared with the Couette flow case $\bar {U}(t,y)=U(y)=y$, the extra exponent $\tilde \delta$ in (1.9) is derived from the linear stretch term $2\bar {U}'\partial _{xy}\phi$ in (1.3), which is absent in the 2D Navier–Stokes equations around shear flows near Couette (see [Reference Bedrossian, Vicol and Wang8]). It is open whether one can remove the extra exponent $\tilde \delta$ in (1.9).
Remark 1.6 The constant $C$ in theorem 1.1 depends on $\frac {1}{|\alpha |}$. As a matter of fact, if $\alpha =0$, the oscillations $-\alpha \partial _xj$ and $-\alpha \partial _x\omega$ disappear in (1.3). As a result, instead of the linear growth $\langle t\rangle$ in the case $|\alpha |\ge \alpha _0$, for $\alpha =0$ the linear stretch term $2\bar {U}'\partial _{xy}\phi$ in (1.3) will lead to $\langle t\rangle ^2$ growth even for $\bar {U}(t,y)=U(y)=y$, which costs more smallness of the initial perturbations in terms of some power of $\nu$ to ensure the stability.
1.3 Notations
(1) Throughout this paper, we use the standard notation
\[ \left\langle x\right\rangle =\sqrt{1+x^2} \]we write $\left \langle \nabla \right \rangle ^s$ for the operator with symbol\[ \left\langle \nabla\right\rangle^s=\left( 1+k^2+\eta^2\right) ^{\frac{s}{2}}. \](2) We use the notation $f\lesssim g$ to mean that there exist some constant $C>0$ such that $f\le Cg$. This constant $C$ may depend on $N$ and $\alpha$, but not on $\nu$.
(3) For a function $f(x,y)$, We denote the projection of $f$ onto the zero frequencies in $x$ by
\[ f_0(y)=\int_{\mathbb{T}}f(x,y){\rm d}x. \]Then we write\[ f_\ne(x,y)=f(x,y)-f_0(y) \]for the projection onto the nonzero frequencies in $x$.(4) The Fourier transform of function $f$ is denoted by
\[ \mathcal{F}(f)=\hat{f}(k,\eta)=\frac{1}{2\pi}\int_{\mathbb{T}\times \mathbb{R}}e^{{-}i(kx+\eta y)}f(x,y){\rm d}x{\rm d}y. \]The Fourier multiplier with symbol $m(t,k,\eta )$ is given by\[ mf=\mathcal{F}^{{-}1}\left( m(t,k,\eta) \mathcal{F} f\right) . \](5) For any $a \in \mathbb {R}$, we use the shorthand notation
(1.12)\begin{equation} O_a^t=e^{a \partial_x t} \end{equation}to denote the multiplier with symbol $e^{iakt}$. We then write $\partial _t O_a^t$ to denote the Fourier multiplier with symbol $iake^{iakt}$.(6) For $s\ge 0$, we define the Sobolev space $H^s$ by using the norm
\[ \left\| f\right\|_{H^s} := \left\| \left\langle \nabla\right\rangle^s f \right\|_{L^2}. \]The notation $L^pL^q=L^p_tL^q_{x,y}$ is used for the Banach space $L^p([0,T];L^q(\Omega ))$ with norm\[ \left\| f(t,x)\right\|_{L^pL^q}=\big\| \|f(t,\cdot)\|_{L^q} \big\|_{L^p}. \](7) For two real functions $f$ and $g$ , we write the associated inner product as
\[ \left\langle f,g \right\rangle=\int_{\mathbb{T}\times\mathbb{R}} f g {\rm d}x {\rm d}y, \]and denote\[ \langle f,g \rangle _{H^s}= \left\langle \langle\nabla\rangle^sf, \langle\nabla\rangle^sg \right\rangle. \]
2. Reformulations and key ideas of the proof
2.1 Elsässer variables under the restriction $\mu =\nu$
Since we focus on the case $\nu =\mu$ in this paper, the symmetry of the system (1.3) enables us to reformulate it by using the Elsässer variables
Then $w^{\pm }$ solves
Such kind of variables have played important roles in the study of large time behaviours of solutions to MHD equations in the absence of shear flows [Reference Bardos, Sulem and Sulem2, Reference Cai and Lei9, Reference He, Xu, Li and Yu19, Reference Wei and Zhang43], and the nonlinear stability result [Reference Liss31] where the three dimensional Couette flow $(y,0,0)^{\top }$ is taken into consideration as mentioned in § 1.
Similar to the 3D case, the background magnetic field $(\alpha, 0)^\top$ introduces oscillations (see the $-\alpha \partial _xw^\pm$ terms in (2.1)) that may stabilize the system. Following [Reference Liss31], we define the profiles
to hide the oscillations in the new unknowns $z^{\pm }$, where $O^t_{\pm \alpha }$ is defined in (1.12). Then $z^{\pm }$ solves
2.2 Change of coordinates
In this paper we use the coordinate transform introduced by Bedrossian, Vicol and Wang in [Reference Bedrossian, Vicol and Wang8] to unwind the decaying background shear flow $\bar {U}(t,y)$ in (2.1):
Denote the spatial derivatives of the shear flow in the new coordinates as follows
Note that by the chain rule, we have
For any function $\tilde {h}$ in the $(x,y)$ coordinates, the corresponding function $h$ in the $(X,Y)$ coordinates is given by
Then $\nabla \tilde h$ and $\Delta \tilde h$ can be rewritten in the new coordinate system (2.2) in terms of $h$ as follows (the notations $\nabla _t$ and $\Delta _t$ are introduced naturally):
and
where we have used the notations
and the modified Laplace operator is given by
Finally, using the fact $\partial _t \bar {U} =\nu \partial _y^2\bar {U}$ and the definitions of $a$ and $b$, we have
In particular,
which, together with the fact $\partial _t\bar {U}=\nu \partial _y^2\bar {U}$ imply that
2.3 Definition of $\Delta _t^{-1}$ and system (2.1) under new coordinates
In this subsection, we first give the definition of the inverse of $\Delta _t$ in the spirit of Antonelli, Dolce, and Marcati [Reference Antonelli, Dolce and Marcati1]. For the sake of completeness, we sketch the definitions below.
To begin with, assume that $a^2-1=(a^2-1)(t, Y)$ and $b=b(t, Y)$ are given functions in $L^\infty (\mathbb {R}^+; H^1(\mathbb {R}))$, then the fact that $\Delta _L^{-1}$ is well defined for $k\ne 0$ enables us to define an operator on $L^2(\mathbb {T}\times \mathbb {R})$:
with
for some constant $C_*\ge 1$, see proposition 4.1 in [Reference Antonelli, Dolce and Marcati1] for more details.
Definition 2.1 Assume that $(\|a^2-1\|_{L^\infty H^1}+\|b\|_{L^\infty H^1})\le \delta$ such that $C_*\delta <1$, then for $k\ne 0$, let us define
where
Remark 2.2 For the operator $\Lambda$, we also have the following useful identity
Now we are in a position to rewrite (2.1) under the coordinates defined by (2.2). To this end, let us denote
Note that
and
Then it follows from (2.1) that the equations of $Z^\pm$ take the form of
or
where
with
and
Here ‘$\mathrm {NLT}$’, ‘$\mathrm {NLS}$’ and ‘$\mathrm {LP}$’ stand for ‘nonlinear transport’, ‘nonlinear stretch’ and ‘linear perturbation’, respectively.
2.4 Toy model and key ideas
Compared with the Navier–Stokes equations around 2D shear flows near Couette [Reference Bedrossian, Vicol and Wang8], and the MHD equations around 3D Couette flows [Reference Liss31], we will encounter new difficulties for the MHD equations around 2D shear flows near Couette. In order to track the difficulties precisely, let us consider the toy model
or equivalently in Fourier variables
It is worth pointing out that if the shear flow is the Couette flow, i.e., $\bar {U}(t,y)=U(y)=y$, the operator $\Lambda$ in (2.20) will not appear. This scenario is reminiscent of the toy model introduced by Bedrossian, Germain and Masmoudi in [Reference Bedrossian, Germain and Masmoudi3] for the 3D Navier–Stokes equations near Couette:
To balance the interaction between $\frac {2k(\eta -kt)}{k^2+(\eta -kt)^2+l^2}\hat {g}$ and $-\nu (k^2+(\eta -kt)^2+l^2)\hat {g}$, the authors in [Reference Bedrossian, Germain and Masmoudi3] constructed a multiplier $m(t,k, \eta,l)$ satisfying
More precisely, the two dimensional analogue of the multiplier $m(t,k, \eta,l)$ is given as follows:
(1) if $k = 0: m(t,0,\eta ) = 1$;
(2) if $k \ne 0, \frac {\eta }{k} < -1000\nu ^{-\frac {1}{3}}: m(t,k,\eta ) = 1$;
(3) if $k \ne 0, -1000\nu ^{-\frac {1}{3}} < \frac {\eta }{k} < 0$:
• $m(t,k,\eta ) = \frac {k^2+\eta ^2}{k^2+(\eta -kt)^2}$ if $0 < t < \frac {\eta }{k}+1000\nu ^{-\frac {1}{3}}$,
• $m(t,k,\eta ) = \frac {k^2+\eta ^2}{k^2+(1000k\nu ^{-\frac {1}{3}})^2}$ if $t > \frac {\eta }{k}+1000\nu ^{-\frac {1}{3}}$;
(4) if $k \ne 0, \frac {\eta }{k} > 0$:
• $m(t,k,\eta ) = 1$ if $t < \frac {\eta }{k}$,
• $m(t,k,\eta ) = \frac {k^2}{k^2+(\eta -kt)^2}$ if $\frac {\eta }{k} < t < \frac {\eta }{k} + 1000\nu ^{-\frac {1}{3}}$,
• $m(t,k,\eta ) =\frac {k^2}{k^2+(1000k\nu ^{-\frac {1}{3}})^2}$ if $t > \frac {\eta }{k} + 1000\nu ^{-\frac {1}{3}}$.
Notice, in particular, that
and
In our case, for the Couette flow, it suffices to use $m^{\frac {1}{2}}(t,k,\eta )$ instead of $m(t,k,\eta )$ to suppress the potential growth caused by the linear stretch term $\frac {k(\eta -kt)}{k^2+(\eta -kt)^2}\widehat {f^{\pm }}$. Nevertheless, for general shear flows the presence of the operator $\Lambda$ may amplify the linear stretch effect since the norm of $\Lambda$ may be larger than 1 even though the shear flow $\bar {U}$ is close to Couette. Roughly speaking, one can regard the linear stretch term in (2.20) as $(1+O(\tilde \delta ))\frac {k(\eta -kt)}{k^2+(\eta -kt)^2}\widehat {f^{\pm }}$ with $\tilde \delta >0$. Then it is natural to modify the multiplier $m^{\frac {1}{2}}(t,k,\eta )$ so as to suppress the extra growth resulting from the operator $\Lambda$. To this end, our strategy is to replace $m^{\frac {1}{2}}(t,k,\eta )$ with $\tilde {m}^{\frac {1}{2}}(t,k,\eta )$, where $\tilde m(t,k,\eta )$ is defined by
with arbitrary $\tilde \delta >0$. Note that
Then (2.21) and (2.22) reduce to
and
For $(t,k,\eta )\in [0,\infty )\times \mathbb {Z}\backslash \{0\}\times \mathbb {R}$, let us define the following three disjoint sets
Then the effects of the linear stretch are summarized as follows:
(1) if $(t, k, \eta )\in D_{dam}$, then $k(\eta -kt)\ge 0$, and thus the linear stretch term behaves as a damping.
(2) if $(t, k, \eta )\in D_{dis}$, we have
\[ \left|t-\frac{\eta}{k}\right|\ge1000\nu^{-\frac {1}{3}}, \]and thus(2.27)\begin{equation} \frac{|k(\eta-kt)|}{k^2+(\eta-kt)^2}\le \frac{1}{1000^3} \nu\left(k^2+(\eta-kt)^2\right), \end{equation}which means that the linear stretch is dominated by the dissipation.(3) by the definition of $\tilde m$, there holds
(2.28)\begin{equation} \frac{|k(\eta-kt)|}{k^2+(\eta-kt)^2}{\bf 1}_{D_{mul}}(t, k,\eta)=\frac{1}{1+\frac 32\tilde\delta}\frac{-\partial_t(\tilde m^{1/2})(t, k,\eta)}{\tilde m^{1/2}(t, k, \eta)}, \end{equation}that is to say, the effect of the linear stretch is balanced by the evolution of ${\tilde m}^\frac {1}{2}$ on $D_{mul}$.
Remark 2.3 Clearly, the following decomposition of unity holds
Then it follows from (2.27) and (2.28) that
The treatment of the general shear flow $(\bar {U},0)^\top$ is much more complicated than that of the Couette flow $(y,0)^\top$. In fact, for the case $\bar {U}(t,y)=U(y)=y$, once the multiplier $m(t,k,\eta )$ is well defined as above, then it is straightforward to deal with the linear stretch term, see (3.24). In particular, the damping effect stemming from the linear stretch term when $(t,k,\eta )\in D_{dam}$ can be ignored. However, for the case that $U(y)$ is close to $y$, the appearance of the operator $\Lambda$ makes the damping effect of the linear stretch $\frac {k(\eta -kt)}{k^2+(\eta -kt)^2}\widehat {\Lambda f^\pm }$ unclear even though $(t,k,\eta )\in D_{dam}$. Our strategy is to isolate the damping effect by using (2.13):
The first term on the right hand side of (2.30) is the main part, and the second term is the perturbation. Some delicate commutator estimates will be performed to treat the perturbation. As a result, lots of errors appear. To close the estimates, the damping effect captured by the main part of the linear stretch in (2.30), together with the dissipation and other good terms, will be used to absorb the errors. See Step I of § 4.1 for more details.
In addition, the presence of the operator $\Lambda$ leads us to estimate the composition $B\circ \Lambda$ for some multiplier $B$ involved in this paper. The continuity of $B\circ \Lambda$ depends on the commutator estimates of $B$. That's why we give a collection of commutator estimates in Appendix B. In particular, some extra terms appear in the commutator estimate of $\sqrt {-\frac {\partial _t(\tilde m^\frac {1}{2})}{\tilde m^\frac {1}{2}}}\tilde K$, see lemma B.4. This motivates us to establish a composition inequality for the multipliers whose commutator estimates have extra errors with better commutator estimates, see lemma C.1 for the details.
On the other hand, it is worth pointing out that the above toy model (2.19) ignored the oscillation terms $O^t_{\pm 2\alpha }S\Lambda f^{\mp }$, which should be taken into consideration as well in the energy estimates. In order to take advantage of the oscillator $O^t_\alpha$, noting that
integrating by parts with respect to the time variable will be exploited as that in [Reference Liss31]. In this way, we will inevitably encounter the time derivative of the operator $\Lambda$. To achieve this, using again (2.13), we obtain an important relation
where $\partial _t\tilde \Lambda$ is derived from (2.10)
See the estimates of $\mathrm {OLS}_5$ in Step II of § 4.1 for more details.
Finally, we would like to remark that, unlike [Reference Antonelli, Dolce and Marcati1], the coefficients $a$ and $b$ hidden in the definition of $\Delta _t^{-1}$ are time dependent (see (2.10) and (2.12)), and $\partial _ta$ and $\partial _tb$ are involved when performing integrating by parts in time (see (2.33)). Accordingly, recalling (2.3) and (2.9), we find that some higher derivatives of $\bar {U}$ are actually involved. This partially explains the extra regularities required in (1.8).
3. Stability of the Couette flow
For comparison, we discuss in this section the case that the shear flow is the Couette flow. In fact, under the condition $\bar {U}(t,y)=U(y)=y$, the change of coordinates (2.2) reduces to
Without causing confusion, we continue to use the unknowns and notations introduced in § 2. Then it is easy to see that the system (2.15) now reads
The purpose of this section is to establish the following theorem.
Theorem 3.1 Let $\mu =\nu \in (0,1], N>1$. There exist a universal constant $\alpha _0>0$, and a positive constant $\delta$ depending only on $N$ and $\alpha$, such that if $|\alpha |\ge \alpha _0$ and
then the following estimates hold
and
Theorem 3.1 will be proved by using the Fourier multiplier method. In other words, the norms involved in the proof are defined based on special, time-dependent Fourier multipliers. Apart from the multiplier $m(t, k,\eta )$ defined in § 2, we also need two extra multipliers that are modified from the ones introduced in the study of the stability of the three dimensional Couette flow in [Reference Bedrossian, Germain and Masmoudi3]. More precisely, define
The multiplier $M_1$ is used to capture the inviscid damping effect in terms of the $L^2$ time integrability, and $M_2$ is designed to show the enhanced dissipation effect. Clearly, $M_1$ and $M_2$ can be given explicitly, and then one deduces that
For more properties of multiplier $M_1$ and $M_2$, one can refer to lemma 4.1 of [Reference Bedrossian, Vicol and Wang8].
Let us denote
Then from (2.21) and (3.9), we find that for any $f=f(X,Y)\in H^N$,
In addition, we need to estimate the interactions between the non-zero modes in the treatment of the nonlinear terms, so the following lemma is introduced.
Lemma 3.2 Let $N>1$. Then for all $f\in L^2H^N$ and $g$ such that $\nabla _Lg\in L^2H^N$, there holds
Proof. We first write
Thanks to (2.22), there hold
and
Combining these two estimates with (3.11) yields
To estimate $\partial _X\Delta _L^{-1}f_\neq \partial _Y^Lg_\neq$, in view of (2.21), (2.22) and the definition of $M$, we arrive at
and
Accordingly,
It follows from (3.15) and (3.18) that (3.12) holds. This completes the proof of lemma 3.2.
3.1 Proof of theorem 3.1
To prove theorem 3.1, it suffices to establish the following a priori estimates:
and
Indeed, (3.4) is a direct consequence of (3.11) and (3.19), and (3.5) is nothing but (3.20). To prove (3.6), by the definition of $M_2$, we have
It follows that
Thus, (3.6) follows from (3.11), (3.19) and (3.22) immediately.
Next we will prove (3.19) and (3.20) by using the standard continuity method. First of all, the local well-posedness of the 2D MHD equations in $H^N$ ensures that there exists a $T_0>0$, such that
and
Then define $T^* \le \infty$ to be the maximum of all time $T$ such that (3.19) and (3.20) hold on $[0,T]$. By the continuity, $T^* > T_0$.
We are left to prove that the constant 8 on the right of (3.19) and (3.20) can be replaced by 4, which implies that $T^* = \infty$. In fact, we have the following propositon.
Proposition 3.3 Let $\mu =\nu \in (0,1], N>1$. Assume that (3.19) and (3.20) hold on $[0, T^*]$. There exist a universal constant $\alpha _0>0$ and a positive constant $\delta$ depending only on $N$ and $\alpha$, such that if $|\alpha |\ge \alpha _0$ and (3.3) holds, then the same estimates in (3.19) and (3.20) hold with the occurrences of 8 on the right-hand side replaced by 4.
The proof of proposition 3.3 will be achieved in the following two subsections.
3.2 Improvement of (3.19)
Recalling the definition of the multiplier $K$ in (3.10), from (3.2), we derive the following energy identity:
By the definition of $m$, we have
Thanks to the fact (2.31), one can estimate $\mathrm {OLS}$ by integrating by parts in time:
where
Clearly,
To bound $\mathrm {OLS}_3$, in Fourier variables, we find that
On the support of $\dot {m}$, there holds
Thus,
Moreover,
It follows that
As for $\mathrm {OLS}_4$, in view of (3.2), we have
with
Integrating by parts, and using the fact $|S|\le \frac {1}{2}$, we have
Thanks to the fact $|S|\le \sqrt {-\frac {\dot {M}_1}{M_1}}$, one deduces that
Owing to the periodicity in $X$ variable, we find that
To bound $\mathrm {OLS}_4^{(4)}$, by (3.2), we get
with
By using lemma 3.2, we are led to
In view of (3.16), one deduces that
Recalling that $z^{\pm }=O^t_{\pm \alpha }w^{\pm }$, we then have $z^\pm _0=w^\pm _0=\omega _0\mp j_0$. Consequently,
and hence
Using (3.13), we arrive at
$\mathcal {NLS}1$ and $\mathcal {NLS}2$ can be treated in the same way, we only estimate $\mathcal {NLS}2$ now. To this end, we infer from (2.22) that
This, together with (2.21) and the obvious fact $|\partial _{XX}\Delta _L^{-1}|\le 1$, implies that
Note that $\mathrm {OLS}_5$ can be treated in the same manner as $\mathrm {OLS}_4$, and $\mathcal {NL}$ in (3.23) can be treated in the same way as $\mathrm {OLS}_4^{(4)}$. On the other hand, one can obtain an energy identity for $KZ^-$ similar to (3.23), and estimate the corresponding right hand side terms analogously. Putting the energy estimates for $KZ^+$ and $KZ^-$ together, we find that for sufficiently large $|\alpha |$, $\mathrm {LS}$, $\mathrm {OLS}_1$, $\mathrm {OLS}_3$, $\mathrm {OLS}_4^{(1)}$, $\mathrm {OLS}_4^{(2)}$ can be obsorbed by the left hand. In conclusion, there exist a universal constant $\alpha _0>0$, and a positive constant $C$ depending only on $N$ and $\alpha$, such that if $\left |\alpha \right | > \alpha _0$, we have
which suffices to improve (3.19) as long as $\epsilon \ll \nu ^\frac 56$.
3.3 Improvement of (3.20)
Since $u^2_0=b^2_0=0$, we derive from (1.2) for $\bar {U}(t,y)=U(y)=y$ that $(u^1_0, b^1_0)$ solves
Accordingly, in the coordinate system (3.1), we have
Then
Thanks to the divergence free condition, we have
Note that
and
Then in view of (3.14) and (3.16), we find that
and
Substituting (3.44)–(3.46) into (3.43), noting that the coordinate system (3.1) is the same as the original system in $y$ variable, and using the hypotheses (3.19) and (3.20), we are led to
which is sufficient to improve (3.20) provided $\epsilon \ll \nu ^{\frac {1}{2}}$. Combining (3.41) with (3.47), we complete the proof of proposition 3.3 and hence of theorem 3.1.
4. Stability of the shear flow close to Couette
In this section, we study the stability of the shear flow $(\bar {U}(t,y),0)^\top =(e^{\nu t \partial {yy}}U(y),0)^\top$, with $U(y)$ satisfying
The multiplier that will be used in this section is given by
where $\tilde m$ and $M$ are given in (2.23) and (3.10), respectively. To simplify the presentation, let us denote
The aim of this section is to establish the following theorem.
Theorem 4.1 Let $N>1$. Assume that the shear flow $( U(y),0)$ satisfies (4.1), and $\|(\omega _{\mathrm {in}},j_{\mathrm {in}})\|_{H^N}+\|({u}_{\mathrm {in}}, {b}_{\mathrm {in}})\|_{H^N}=\epsilon \le \delta \nu ^{\frac 56+\tilde \delta }$. Then there exist two positive constants $\alpha _0$ and $\delta _0$ independent of $\nu$, such that for all $|\alpha |\ge \alpha _0$ and $\delta \le \delta _0$, the solution to (1.2) and the profile $Z^\pm$ satisfy the global in time estimates
and
Similar to lemma 3.2, we give the following lemma to treat the non-zero frequency interactions of the nonlinear term.
Lemma 4.2 Let $N>1$. Assume that (4.1) holds with $\delta$ sufficiently small. Then for all $f\in L^2H^N$ and $g$ such that $\nabla _Lg\in L^2H^N$, there holds
Proof. In view of (2.25) and (2.26), there hold
and
Note that the condition (4.1) ensures that (A.3) holds. This, together with the commutator estimates (B.1), (B.9), enables us to use lemma C.1 and the fact
to obtain
Then (4.2) follows immediately.
4.1 Proof of theorem 4.1
The proof of theorem 4.1 is similar to that of theorem 3.1. By the definition of $M_2$, (3.22) still holds with $K$ replaced by $\tilde K$, it suffices to establish the following a priori estimates:
and
Let us define $T^*$ to be the end point of the largest interval $[0,T]$ such that (4.10) and (4.11) hold for all $0\leq t\leq T$. We are left to establish the following proposition.
Proposition 4.3 Assume that the conditions in theorem 4.1 hold, and that (4.10) and (4.11) hold on $[0, T^*]$. Then there exist two positive constants $\alpha _0$ and $\delta _0$ independent of $\nu$, such that for all $|\alpha |\ge \alpha _0$ and $\delta \le \delta _0$, the same estimates in (4.10) and (4.11) hold with the occurrences of 8 on the right-hand side replaced by 4.
Proof. We first improve (4.10). Similar to (3.23), from (2.16) and the definition of $\tilde K$ in (4.2), we have the following energy identity:
The improvement of (4.10) will be achieved by the following four steps.
Step I: estimates of $\mathrm {LS}$. We first split $\mathrm {LS}$ into two parts:
Thanks to (2.13), one can split $\mathrm {LS}^{dam}$ into two parts
By the definition of $\tilde \Lambda$ in (2.10), there holds
Combining this with (B.8), (B.11), and (A.3) yields
Using (B.8), (B.9), (B.11), lemma C.1 and (2.29), we find that
where $|S|$ denotes the multiplier with symbol $\frac {|k(\eta -kt)|}{k^2+(\eta -kt)^2}$. Substituting this into (4.14), and using (B.8), (B.9) and lemma C.1 again to bound the second term on the right hand side of (4.14), one deduces that
Next we turn to bound $\mathrm {LS}^*$. From (2.27) and (2.28), we infer that
By virtue of (B.8) and lemma C.1, we have
Now we are left to bound $\left \|\sqrt {-\frac {\partial _t(\tilde m^{1/2})}{\tilde m^{1/2}}}\tilde K\Lambda Z^+_{\neq }\right \|_{L^2L^2}$. In fact, thanks to (B.8), (B.10), and lemma C.1, we are led to
where we have used (4.18) to bound $\|\nabla _L\tilde K\Lambda Z^+_{\neq }\|_{L^2L^2}$, and used (B.8), (B.9) and lemma C.1 to bound $\left \|\sqrt {-\frac {\dot {M}_1}{M_1}}\tilde K\Lambda Z^+_{\neq }\right \|_{L^2L^2}$, respectively. It follows from (4.15) and (4.19) that
Substituting (4.18) and (4.20) into (4.17), and using Cauchy–Schwarz inequality, we arrive at
where
which is increasing in $\delta \in (0, \frac {1}{3C}]$, and $C_\delta \rightarrow 1$ as $\delta \rightarrow 0+$.
Step II: estimates of $\mathrm {OLS}$. Similar to (3.25), we write
where
We postpone the treatment of the nonlinear terms in $\mathrm {OLS}_4$ and $\mathrm {OLS}_5$ to Step IV, and focus on the linear terms here. The estimates for $\mathrm {OLS}_1$ and $\mathrm {OLS}_2$ are the same as (3.26), and thus omitted. To estimate $\mathrm {OLS}_3$, similar to (3.27)–(3.29), we have
and
Combining these calculations with (3.30), and using (B.8), (B.9) and lemma C.1, we are led to
To bound $\mathrm {OLS}_4$, by (2.16), we have
where
Integrating by parts, using the fact $|S|\le \frac{1}{2}$ and (4.18), we have
Thanks to the fact $|S|\le \sqrt {-\frac {\dot {M}_1}{M_1}}$, one deduces that
Owing to the periodicity in $X$ variable, we find that
By (2.18), we deduce that
with
Using $|S|\le \sqrt {-\frac {\dot {M}_1}{M_1}}$, (A.3) and lemma C.1, we have
Noting that
and
then we have
Integrating by parts and using (2.4), we find that
where we have used (A.3), (A.8) to bound $\left \|\tilde K(b\partial _Y^L)\right \|_{L^2L^2}$ and $\left \|\tilde K((a^2-1)\partial _Y^L)\right \|_{L^2L^2}$. The estimates of $\mathrm {OLS}_{4}^{(5)}$ will be postponed in Step IV.
Now we rewrite $\mathrm {OLS}_5$ as follows:
To estimate $\mathrm {OLS}_{5,1}$, instead of using (2.16), up to the nonlinear terms and some linear errors, we write $\partial _tZ^{-}$ in terms of $\nu \Delta _tZ^-$ by virtue of (2.6) and (2.15)
Then thanks to the relation (2.12), we have
Thus,
where
Similar to (4.32), we obtain
Similar to (4.26) and (4.29), we arrive at
Clearly, $\mathrm {OLS}_{5,1}^{(4)}$ can be bounded in the same way as (4.31)
The estimates of $\mathrm {OLS}_{5,1}^{(5)}$ will be postponed in Step IV.
To estimate $\mathrm {OLS}_{5,2}$, we split it into two parts according to (2.32) and (2.33):
where
From (A.8) and (3.29), we obtain
The rest part $\mathrm {OLS}_{5,2}^{(2)}$ can be bounded as follows. Using the definition of $\tilde \Lambda ^2_t$ in (2.33), lemmas A.2 and C.1, we are led to
Then it follows this and lemma C.1 that
Step III: estimates of $\mathcal {LP}$. By (2.18), we rewrite $\mathcal {LP}$ as
where
To estimate $\mathcal {LP}_1$, let us denote $Z_\ne :=Z^+_\ne -O^t_{2\alpha }Z^{-}_{\ne }$, and rewrite $\mathcal {LP}_1$ as
Swapping the positions of $\eta$ and $\xi$ in (A.11), then using the resulting inequality, (A.3), lemma C.1, and (4.15), we have
Now we turn to estimate $\mathcal {LP}_2$. Compared with (4.31), we need an extra commutator estimate of $\sqrt {\left |\partial _X\Delta _L^{-1}\right |}$. Indeed, using (A.8), (A.9) and lemma C.1, one easily deduces that
To bound $\mathcal {LP}_3$, it is natural to divide it into two parts as follows
Similar to (4.32), we obtain
For $\mathcal {LP}_{3,0}$, integrating by parts and using (2.4) and (A.3) yields
Step IV: nonlinear estimates. We first collect all the above nonlinear terms to be estimated:
Owing to lemma C.1 and the fact $|S|\le \frac{1}{2}$, we observe that the bounds of the three quantities are essentially the same. To avoid unnecessary repetition, we only sketch the treatment of $\mathcal {NL}$ by modifying the nonlinear estimates in the Couette case, see (3.35)–(3.40). In fact, recalling the definition of $\mathrm {NL}^\pm$ (2.17), we write
here we use the shorthand notation $\mathrm {NLS}^\pm :=\mathrm {NLS}1^\pm +\mathrm {NLS}2^\pm$. Noting that
by virtue of (4.6), (4.8), lemmas 4.2 and C.1, and the hypotheses (4.10) and (4.11), one deduces that
Similar to (3.40), using (2.21), (2.23), (3.39), (A.3), (A.1), and lemma C.1, we arrive at
Now collecting the above estimates in Step I–IV, we conclude that there exist positive constant $\alpha _0$ (sufficiently large) and $\delta _0$ (sufficiently small) independent of $\nu$, such that if $|\alpha |\ge \alpha _0$ and $\delta \le \delta _0$, there holds
where the constant $C$ depends only on $N$ and $\alpha$. It suffices to improve (4.10) as long as $\epsilon \ll \nu ^{\frac{5}{6}+\tilde \delta }$.
The improvement of (4.11) is similar to that of (3.20). Firstly, we write the equations of $U_0^1\mp B_0^1$:
Then we have the energy identity
Using the divergence free condition $\nabla _t\cdot (U\pm B)=0$, we find that
Recalling (2.4), it is easy to rewrite $\partial _Ya$ as follows
Combining this with (A.3) yields $\nu ^{\frac{1}{2}}\|\partial _Ya\|_{L^2H^N}\lesssim \delta$. On the other hand, similar to (3.45) and (3.46), we have
and
It follows that
Integrating by parts, and using (A.3), it is easy to see that
Substituting (4.48) and (4.49) into (4.47), and using the hypotheses (4.10) and (4.11), we have
for some constant $C$ independent of $\nu$. Combining (4.46) and (4.50), one deduces proposition 4.3 under the hypotheses $\epsilon \ll \nu ^{\frac{5}{6}+\tilde \delta }$ and hence of theorem 4.1.
Appendix A. Estimates for the coefficients $a$ and $b$
To begin with, we give a lemma to discuss the relation between the new coordinate system (2.2) and the original $( x,y)$. Please refer to [Reference Inci, Kappeler and Topalov21] and [Reference Bedrossian, Vicol and Wang8] for the proof.
Lemma A.1 Let $s'\ge 2,s'\geq s \geq 0, f\in H^s(\mathbb {R}),$ and $g \in H^{s'}(\mathbb {R})$ be such that $\left \| g\right \|_{H^{s'}} \leq \delta.$ Then, there holds
where the implicit constant obey $C_{s,s'}(\delta )\to 1$ as $\delta \to 0$.
From the properties of the heat equation and lemma A.1, we can deduce the energy estimates of the coefficients $a$ and $b$.
Lemma A.2 Let $s\ge 0$. Assume that $U(y)$ satisfies
then there holds
and
Proof. Note that $\partial _y^l\bar {U}, l\ge 0$ solves
Therefore, integrating by parts, we have
and
Consequently,
and
Combining these two estimates with (2.3), (2.9) and lemma A.1, we find that
and
This completes the proof of lemma A.2.
Appendix B. Commutator estimates
To deal with the nonlinear terms, we need the following lemmas to exchange frequencies. The first one can be regarded as an analogue of Lemma A.1 in [Reference Bedrossian, Germain and Masmoudi3]. We give a proof below for the sake of completeness.
Lemma B.1 The multiplier $\tilde m$ satisfies
Proof. We only consider the case $k \ne 0.$
Case 1: $\frac {\eta }{k} < -1000\nu ^{-\frac {1}{3}}, \tilde m(t,k,\eta ) = 1$. It suffices to estimate $\tilde m^{-1}(t,k,\xi )$.
Case 1.1: $-1000\nu ^{-\frac {1}{3}} < \frac {\xi }{k} < 0$. Now we have
Case 1.2: if $\frac {\xi }{k}>0, \frac {\xi }{k}< t < \frac {\xi }{k}+1000\nu ^{-\frac {1}{3}}$.
Case 1.3: if $\frac {\xi }{k}>0, t > \frac {\xi }{k}+1000\nu ^{-\frac {1}{3}}$.
Case 2: $-1000\nu ^{-\frac {1}{3}} < \frac {\eta }{k} < 0$.
Case 2.1: $0< t<\frac {\eta }{k}+1000\nu ^{-\frac {1}{3}}$, and $\tilde m(t,k,\eta )=( \frac {k^2+\eta ^2}{k^2+(\eta -kt)^2}) ^{1+\frac{3}{2}\tilde \delta }$.
• Case 2.1.1: $-1000\nu ^{-\frac {1}{3}}<\frac {\xi }{k}<0$.
(B.2)\begin{equation} \tilde m(t,k,\xi)\ge\left(\frac{k^2+\xi^2}{k^2+(\xi-kt)^2}\right)^{1+\frac{3}{2}\tilde{\delta}}, \quad \forall ~t\ge0. \end{equation}Therefore,\begin{align*} \frac{\tilde m(t, k,\eta)}{\tilde m(t, k, \xi)}& \le \left(\frac{1+\left(\frac{\eta}{k}\right)^2}{1+\left(\frac{\xi}{k}\right)^2}\cdot\frac{1+\left(\frac{\xi}{k}-t\right)^2}{1+\left(\frac{\eta}{k}-t\right)^2}\right)^{1+\frac{3}{2}\tilde\delta}\\ & \le \begin{cases} \left(\frac{1+\left(\frac{\xi}{k}-t\right)^2}{1+\left(\frac{\eta}{k}-t\right)^2}\right)^{1+\frac{3}{2}\tilde\delta}, \text{ if } \left|\frac{\eta}{k}\right|\le\left|\frac{\xi}{k}\right|\\ \left(\frac{1+\left(\frac{\eta}{k}\right)^2}{1+\left(\frac{\xi}{k}\right)^2}\right)^{1+\frac{3}{2}\tilde\delta}, \, \text{ if } \left|\frac{\eta}{k}\right|>\left|\frac{\xi}{k}\right| \end{cases}\\ & \lesssim\langle \eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \end{align*}• Case 2.1.2: $\frac {\xi }{k}>0, t> \frac {\xi }{k}$. In this case,
(B.3)\begin{equation} \tilde m(t,k,\xi)\ge\left(\frac{k^2}{k^2+(\xi-kt)^2}\right)^{1+\frac{3}{2}\tilde{\delta}}. \end{equation}Therefore,\begin{align*} \frac{\tilde m(t, k,\eta)}{\tilde m(t, k, \xi)}\le \left(\left( 1+\left(\frac{\eta}{k} \right)^2 \right) \cdot \frac{1+\left(\frac{\xi}{k}-t \right)^2 }{1+\left(\frac{\eta}{k}-t \right)^2}\right)^ {1+\frac{3}{2}\tilde\delta} & \le \left(1+\left( \frac{\eta}{k}-\frac{\xi}{k}\right)^2 \right) ^ {1+\frac{3}{2}\tilde\delta}\nonumber\\ & \lesssim \langle \eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \end{align*}where we have used $\frac {\eta }{k}<0<\frac {\xi }{k}< t$.
Case 2.2: $t>\frac {\eta }{k}+1000\nu ^{-\frac{1}{3}}$, and $\tilde m(t,k,\eta )=( \frac {k^2+\eta ^2}{k^2+(1000k\nu ^{-\frac{1}{3}})^2}) ^{1+\frac{3}{2}\tilde \delta }$.
• Case 2.2.1: $-1000\nu ^{-\frac{1}{3}}<\frac {\xi }{k}<0$. In this case,
(B.4)\begin{equation} \tilde m(t,k,\xi)\ge\left(\frac{k^2+\xi^2}{k^2+(1000k\nu^{-\frac{1}{3}})^2}\right)^{1+\frac{3}{2}\tilde{\delta}}, \end{equation}Thus,\[ \frac{\tilde m(t, k,\eta)}{\tilde m(t,k,\xi)}\le\left(\frac{k^2+\eta^2}{k^2+\xi^2}\right)^{1+\frac{3}{2}\tilde{\delta}}\lesssim\langle\eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \]• Case 2.2.2: $\frac {\xi }{k}>0, t> \frac {\xi }{k}$. In this case,
(B.5)\begin{equation} \tilde m(t,k,\xi)\ge\left(\frac{1}{1+(1000\nu^{-\frac{1}{3}})^2}\right)^{1+\frac{3}{2}\tilde{\delta}}, \end{equation}Thus,\[ \frac{\tilde m(t, k,\eta)}{\tilde m(t,k,\xi)}\le\left(1+\left( \frac{\eta}{k}\right)^2\right)^{1+\frac{3}{2}\tilde{\delta}}\le \left(1+\left( \frac{\eta}{k}-\frac{\xi}{k}\right) ^2\right)^{1+\frac{3}{2}\tilde{\delta}}\lesssim\langle\eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \]
Case 3: $\frac {\eta }{k}> 0$.
Case 3.1: $\frac {\eta }{k}< t<\frac {\eta }{k}+1000\nu ^{-\frac {1}{3}}$, and $\tilde m(t,k,\eta )=( \frac {k^2}{k^2+(\eta -kt)^2}) ^{1+\frac{3}{2}\tilde \delta }$.
• Case 3.1.1: $-1000\nu ^{-\frac {1}{3}}<\frac {\xi }{k}<0.$ In this case, (B.2) holds. Therefore,
\[ \frac{\tilde m(t, k,\eta)}{\tilde m(t,k,\xi)}\le\left( \frac{k^2}{k^2+\xi^2}\cdot \frac{k^2+(\xi-kt)^2}{k^2+(\eta-kt)^2}\right)^{1+\frac{3}{2}\tilde\delta} \lesssim\langle\eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \]• Case 3.1.2: $\frac {\xi }{k}>0$. Now (B.3) still holds. Thus,
\[ \frac{\tilde m(t, k,\eta)}{\tilde m(t,k,\xi)}\le\left( \frac{k^2+(\xi-kt)^2}{k^2+(\eta-kt)^2}\right)^{1+\frac{3}{2}\tilde\delta} \lesssim\langle\eta-\xi\rangle^{2(1+\frac{3}{2}\tilde\delta)}. \]
Case 3.2: $t>\frac {\eta }{k}+1000\nu ^{-\frac{1}{3}},$ and $\tilde m(t,k,\eta )=( \frac {1}{1+(1000\nu ^{-\frac{1}{3}})^2}) ^{1+\frac{3}{2}\tilde \delta }$.
• Case 3.2.1:$-1000\nu ^{-\frac {1}{3}}<\frac {\xi }{k}<0$. Using (B.4), we have
(B.6)\begin{equation} \frac{\tilde m(t, k,\eta)}{\tilde m(t,k,\xi)}\le1. \end{equation}• Case 3.2.2: $\frac {\xi }{k}>0,$ In this case, In this case, (B.5) holds. Consequently, (B.6) holds as well.
The proof of lemma B.1 is completed.
Recalling the definition of $\tilde K$ in (4.2), using (B.1) and the following fact
we get the following corollary immediately.
Corollary B.2 The multiplier $\tilde K$ satisfies
In view of the definitions of $M_1$ and $M_2$ in (3.7) and (3.8), respectively, using (B.7) again, one easily derives the commutator estimates for $\sqrt {-\frac {\dot {M}_i}{M_i}}, i=1, 2$.
Lemma B.3 For $k \ne 0$ there hold
Finally, the estimates of $\sqrt {-\frac {\partial _t(\tilde m^{1/2})}{\tilde m^{1/2}}}$ is given below.
Lemma B.4 For $k \ne 0$ there holds
Proof. Using (B.7), it is not difficult to verify that
From the above inequality and (2.29) with $\eta$ replaced by $\xi$, we are let to
Combining this with (2.28) yields
Then (B.10) is a consequence of (B.8) and (B.12). We complete the proof of lemma B.4.
Appendix C. Composition inequalities for the operator $\Lambda$ in $H^N$
Since $\Lambda$ is involved in the definition of $\Delta _t^{-1}$, recalling the relation (2.14), it is inevitable to deal with the composition $B_1\circ \Lambda$ for some multiplier $B_1$. Based on the commutator estimates of $B$, we establish the following composition inequality.
Lemma C.1 Let $l\in \mathbb {N}^+$, and $B_i, 1\le i\le l$ be Fourier multipliers, satisfying
for some constants $C_i>0$ and $\beta _i\ge 0, 1\le i\le l$. If $\|{a^2-1}\|_{L^\infty H^{N+\beta +1}}+\|{b}\|_{L^\infty H^{N+\beta +1}}\le \frac {\delta }{2}$ with $\beta =\max \left \{\beta _1,\cdots, \beta _l\right \}$, then we have
Proof. We first write $\Delta _L$ in terms of $\Delta _t$
Then
and thus
Thanks to (C.1), using Young's inequality for convolution, we find that
where we have used the fact $\Lambda =\Delta _L\Delta _t^{-1}$ in the last line above. Similarly, for $k\neq 0$, there holds
Thus, we have
Substituting (C.5) and (C.6) into (C.4), and using the restriction on $a$ and $b$, we arrive at
and hence (C.2) follows. This completes the proof of lemma C.1.
Corollary C.2 Let $B$ be Fourier multiplier satisfying
where $C$ and $\beta$ are two positive constants. If $\|{a^2-1}\|_{H^{N+\beta +2}}+\|{b}\|_{H^{N+\beta +2}}\le \frac {\delta }{2}$, then we have
and
where $\tilde \Lambda$ and $\tilde \Lambda _t^1$ are given in (2.10) and (2.33), respectively.
Proof. Recalling the definition of $\tilde \Lambda$ in (2.10), following the proof of (C.5) and (C.6) with $\Delta _t^{-1}$ replaced by $(-\Delta _L)^{-1}$, we get (C.7) immediately. Now we turn to prove (C.8). In fact, using (2.33) and lemma C.1, we are led to
In view of (3.16) and (3.29), similar to (C.5) and (C.6), and using the commutator estimate (B.9) and lemma C.1, we arrive at
Moreover, it follows from (C.7), (3.29), and lemma C.1 that
Substituting the above two inequalities into (C.9) yields (C.8). The proof is completed.
Acknowledgements
R. Zi is partially supported by NSF of China under Grants 12222105.