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

Temperature patches for the subcritical Boussinesq–Navier–Stokes System with no diffusion

Calvin Khor111Corresponding author.   and Xiaojing Xu
Abstract

In this paper, we prove that temperature patch solutions to the subcritical Boussinesq–Navier–Stokes System with no diffusion preserve the Hölder regularity of their boundary for all time, which generalises the previously known result by F. Gancedo and E. García-Juárez [Annals of PDE, 3(2):14, 2017] to the full range of subcritical viscosity.

Keywords Boussinesq–Navier–Stokes System, temperature patch, subcritical dissipation, global existence.
MSC Classification 35R05 (Primary) 35Q35, 35F25, 35A01, 35A02 (Secondary)

1 Introduction

This paper studies temperature patch solutions of the following initial-value problem for the subcritical Boussinesq–Navier–Stokes equations, α(12,1)\alpha\in(\frac{1}{2},1):

{ut+uu+Λ2αu=p+θe2,(t,x)+×2,u=0,θt+uθ=0,u|t=0=u0,θ|t=0=θ0.\displaystyle\left\{\begin{array}[]{rcll}u_{t}+u\cdot\nabla u+\Lambda^{2\alpha}u&=&\nabla p+\theta e_{2},&(t,x)\in\mathbb{R}_{+}\times\mathbb{R}^{2},\\ \nabla\cdot u&=&0,\\ \theta_{t}+u\cdot\nabla\theta&=&0,\\ u|_{t=0}&=&u_{0},\\ \theta|_{t=0}&=&\theta_{0}.\end{array}\right. (1.6)

Here, Λ2α=(Δ)α\Lambda^{2\alpha}=(-\Delta)^{\alpha} is a fractional Laplacian on 2\mathbb{R}^{2} defined initially for Schwartz functions as a Fourier multiplier Λ2αf^(ξ)=|2πξ|2αf^(ξ)\widehat{\Lambda^{2\alpha}f}(\xi)=|2\pi\xi|^{2\alpha}\hat{f}(\xi). Also, e2:=(0,1)T2e_{2}:=(0,1)^{T}\in\mathbb{R}^{2}, u=u(t,x)=(u1(t,x),u2(t,x))T2u=u(t,x)=(u_{1}(t,x),u_{2}(t,x))^{T}\in\mathbb{R}^{2} is the velocity of the fluid, p=p(t,x)p=p(t,x)\in\mathbb{R} is the pressure of the fluid, θ=θ(t,x)\theta=\theta(t,x)\in\mathbb{R} is the temperature of the fluid, and u0,θ0u_{0},\theta_{0} are the initial data of the system.

This active scalar transport system for θ\theta arises as a natural generalisation of the Boussinesq–Navier–Stokes system, where the dissipation is the full Laplacian (corresponding to α=1\alpha=1). Details of the physics described by the Boussinesq–Navier–Stokes system can be found in [44]. Actually, (1.6) is a special case of a two parameter family of equations where there is also a diffusion term Λ2βθ\Lambda^{2\beta}\theta in the temperature equation; see for instance the papers [54], [43], [50], [53], and citations within. Only local results for the zero viscosity and zero diffusion case are known, such as those in [6].

In addition, some authors consider the opposite situation, where α=0\alpha=0 and β>0\beta>0 (called the Euler–Boussinesq system) such as [30], [29], and [52], but to the best of the authors’ knowledge, there are no other papers currently available studying temperature patches for (1.6) with α(12,1)\alpha\in(\frac{1}{2},1).

We consider solutions in the following sense:

Definition 1.1.

We say that (u,θ)Lloc2([0,]×2;2)×Lloc2([0,]×2)(u,\theta)\in L^{2}_{\text{loc}}([0,\infty]\times\mathbb{R}^{2};\mathbb{R}^{2})\times L^{2}_{\text{loc}}([0,\infty]\times\mathbb{R}^{2}) with u=0\nabla\cdot u=0 is a solution to (1.6) if for every divergence-free ϕCc((0,T)×2;2)\phi\in C^{\infty}_{c}((0,T)\times\mathbb{R}^{2};\mathbb{R}^{2}), and for every ψCc((0,T)×2),\psi\in C^{\infty}_{c}((0,T)\times\mathbb{R}^{2}),

02uϕt+uu:ϕuΛ2αϕθϕe2dxdt=0,\displaystyle\int_{0}^{\infty}\int_{\mathbb{R}^{2}}u\cdot\phi_{t}+u\otimes u:\nabla\phi-u\cdot\Lambda^{2\alpha}\phi-\theta\phi\cdot e_{2}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0, (1.7)
02θψt+θuψdxdt=0.\displaystyle\int_{0}^{\infty}\int_{\mathbb{R}^{2}}\theta\psi_{t}+\theta u\cdot\nabla\psi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0. (1.8)

By a ‘temperature patch solution’, we mean a solution of (1.6) where θ\theta is an indicator function for each time t0t\geq 0. These are similar to the sharp front solutions of the SQG equation and their variants which have been and are extensively studied [47], [46], [20], [19], [18], [21], [13], [5], [7], [38], [39], [34], [35], [32], [33], and the more classical theory of vortex patches for the 2D Euler equation as laid out in [41], and also [10], [11], and [37]. The Boussinesq system also supports initial data of ‘Yudovich’ type. We mention a recent paper [42] working on local-in-time Yudovich solutions to the full inviscid equation. The case of critical diffusion was studied in [55]. Similar problems can be and have been studied for other systems with a transport equation with no diffusion [12], [26] [23], and multiple patch solutions for SQG have also been studied [36], [25]. Finally we mention some recent work [24] on piecewise Hölder solutions to the 3D analogue of the system (1.6) with α=1\alpha=1.

Global unique classical solutions to (1.6) with α=1\alpha=1 were shown to exist in HkH^{k} spaces in [4] and [31]; global (weak) solutions in (L2B,11)×B2,10(L^{2}\cap B^{-1}_{\infty,1})\times B^{0}_{2,1} were studied in [2], and global (weak) solutions in L2×L2L^{2}\times L^{2} were proven to exist in [27] with uniqueness proven in [14]. In the case of temperature patches, Abidi and Hmidi [1] proved the persistence of C1C^{1} regularity of the boundary, and Danchin and Zhang [15] improved this result to C1+εC^{1+\varepsilon} regularity. In Gancedo and García–Juárez [22], a second proof of Danchin and Zhang’s result was given, and they improved the result to C2+εC^{2+\varepsilon} persistence, in particular implying that the curvature of the patch remains bounded.

In [28], the critical equation α=1/2\alpha=1/2 has been studied in a low regularity scenario. However, the well-posedness result there requires initial data θ0B,10Cb0\theta_{0}\in B^{0}_{\infty,1}\hookrightarrow C^{0}_{b} which falls short of allowing LL^{\infty} data, of which temperature patches are a special case. Our work shows that this seems to be a special feature of the critical α=1/2\alpha=1/2 case, as LL^{\infty} data is allowed for all α(1/2,1].\alpha\in(1/2,1]. This is a curious property, as the the transport evolution of θ\theta gives that weak solutions cannot increase LpL^{p} norms of θ\theta. At the same time, there was some foreshadowing, as the method of Gancedo and García-Juárez which relies on the explicit formula for the heat kernel can only be replicated for the critical case α=1/2\alpha=1/2, and etΛfe^{-t\Lambda}\nabla f has roughly the same regularity as ff (as opposed to etΔfe^{t\Delta}\nabla f being better behaved than ff: see [22] for details.)

Our main technical result is the following existence and uniqueness result:

Theorem 1.2.

Suppose that α(12,1)\alpha\in(\frac{1}{2},1), u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p} for some p(2,)p\in(2,\infty), u0=0\nabla\cdot u_{0}=0, and θ0L1L\theta_{0}\in L^{1}\cap L^{\infty}. Then there is a unique global solution (u,θ)(u,\theta) in the sense of Definition 1.1 to (1.6) such that for each ρ[1,min(2,p/2))\rho\in[1,\min(2,p/2)) and for each α[0,α)\alpha^{\prime}\in[0,\alpha),

uLtW1,pLtρC2α, and θLtL1L.\displaystyle u\in L^{\infty}_{t}W^{1,p}\cap L^{\rho}_{t}C^{2\alpha^{\prime}},\ \text{ and }\ \theta\in L^{\infty}_{t}L^{1}\cap L^{\infty}. (1.9)

Using Theorem 1.2, for any α<α\alpha^{\prime}<\alpha, we can control the C2αC^{2\alpha^{\prime}} boundary regularity of temperature patch solutions to (1.6):

Theorem 1.3.

Suppose that α(12,1)\alpha\in(\frac{1}{2},1), u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p} for some p(2,]p\in(2,\infty], u0=0\nabla\cdot u_{0}=0, and θ0=𝟏D0\theta_{0}=\mathbf{1}_{D_{0}} be an indicator function of a simply connected set D0D_{0} with D0C2αD_{0}\in C^{2\alpha^{\prime}} for some α<α\alpha^{\prime}<\alpha. Then the global solution θ\theta to (1.6) remains an indicator, θ(t)=𝟏D(t)\theta(t)=\mathbf{1}_{D(t)} where D(t)LtC2α\partial D(t)\in L^{\infty}_{t}C^{2\alpha^{\prime}} for all t0t\geq 0.

The main method of proof is the use of the special structure of the equation by the study of Γ:=ω2αθ\Gamma:=\omega-\mathcal{R}_{2\alpha}\theta, where 2α\mathcal{R}_{2\alpha} is a smoothing operator. It turns out that it is easier to study this combination of terms than the vorticity ω=u\omega=\nabla^{\perp}\cdot u by itself. This is the method used in [28], but since α>1/2\alpha>1/2 the proof is more streamlined.

Theorem 1.2 raises the following interesting questions: (a) Is it possible to control the curvature for α(1/2,1)\alpha\in(1/2,1), as it is possible in α=1\alpha=1? (b) can the critical equation α=1\alpha=1 support unique temperature patch solutions, and what regularity of their boundary is preserved?

The remainder of the paper is organised as follows. In Section 2, we list the notation that we use for function spaces and inequalities. In Section 3, we explain how introducing the term Γ\Gamma leads to better estimates. In Section 4, we give some easy a priori estimates from the equation obtained by classical means. In Section 5, we use the equation for Γ\Gamma to derive better a priori estimates for the vorticity (and hence the velocity). In Section 6, we use the Osgood Lemma to prove uniqueness of solutions. In Section 7, we use the quantitative bound from the Osgood Lemma to show existence of solutions, and also present the proof for conservation of Hölder regularity of temperature patch boundaries.

2 Preliminaries

We follow the notation of [28] as follows.

  • We write ϕk(t)\phi_{k}(t) to denote any function of the form

    ϕk(t)=C0expexpexpk times(C1t).\phi_{k}(t)=C_{0}\underbrace{\exp\exp\dots\exp}_{k\text{ times}}(C_{1}t).
  • We write ABA\lesssim B to mean that |A|CB|A|\leq CB for some constant CC that does not depend on A,B,A,B, or any other variable under consideration. We write Aϕ1,,ϕNBA\lesssim_{\phi_{1},\dots,\phi_{N}}B to emphasise that the implicit constant CC depends on the NN quantities ϕ1,,ϕN\phi_{1},\dots,\phi_{N}.

  • When (X,X)(X,\|\bullet\|_{X}) is a Banach space, we will write for p[1,]p\in[1,\infty]

    LtpXL^{p}_{t}X

    to denote the Bochner space Lp(0,t;X)L^{p}(0,t;X); a function f=f(t)f=f(t) with values in XX is in LtpXL^{p}_{t}X if

    fLtpX:=f(t)XLtp<.\displaystyle\|f\|_{L^{p}_{t}X}:=\left\|\|f(t)\|_{X}\right\|_{L^{p}_{t}}<\infty. (2.1)

    In a norm, we also abusively adopt the notation that a subscripted variable qq, like qr\ell^{r}_{q} below, means that the norm is to be taken with respect to qq. This should not cause confusion with the above convention for time integrals. See for instance [17] or [48] for details on Bochner spaces.

2.1 Littlewood–Paley Decomposition and Function spaces

We fix our notation for Littlewood–Paley decompositions here; see [3], [8], [45], [51], or [49] for more details about Littlewood–Paley, Besov spaces, and Tilde spaces.

There exist two radial non-negative functions χ𝒟(d)\chi\in\mathcal{D}\left(\mathbb{R}^{d}\right) and φ𝒟(d\{0})\varphi\in\mathcal{D}(\mathbb{R}^{d}\backslash\{0\}) such that

  1. (i)

    χ(ξ)+q0φ(2qξ)=1;q1,suppχsuppφ(2q)=\chi(\xi)+\sum_{q\geqslant 0}\varphi\left(2^{-q}\xi\right)=1;\forall q\geqslant 1,\operatorname{supp}\chi\cap\operatorname{supp}\varphi\left(2^{-q}\right)=\emptyset,

  2. (ii)

    suppφ(2j)suppφ(2k)=,\operatorname{supp}\varphi\left(2^{-j}\bullet\right)\cap\operatorname{supp}\varphi\left(2^{-k}\bullet\right)=\emptyset, if |jk|2|j-k|\geqslant 2.

For every v𝒮(d)v\in\mathcal{S}^{\prime}\left(\mathbb{R}^{d}\right) we define Fourier multipliers Δq,Sq\Delta_{q},S_{q} by

Δ1v=χ(D)v;q,Δqv=φ(2qD)v; and Sqv=1pq1Δpv.\Delta_{-1}v=\chi(\mathrm{D})v;\ \forall q\in\mathbb{N},\ \Delta_{q}v=\varphi\left(2^{-q}\mathrm{D}\right)v;\ \text{ and }\ S_{q}v=\!\!\sum_{-1\leq p\leq q-1}\!\Delta_{p}v.
Lemma 2.1 (Bernstein Inequalities).

There exists aa constant CC such that for q,k,1abq,k\in\mathbb{N},1\leqslant a\leqslant b and for fLa(d)f\in L^{a}(\mathbb{R}^{d})

sup|α|=kαSqfLb\displaystyle\sup_{|\alpha|=k}\left\|\partial^{\alpha}S_{q}f\right\|_{L^{b}} Ck2q(k+d(1a1b))SqfLa, and\displaystyle\leqslant C^{k}2^{q\left(k+d\left(\frac{1}{a}-\frac{1}{b}\right)\right)}\left\|S_{q}f\right\|_{L^{a}},\text{ and} (2.2)
Ck2qkΔqfLa\displaystyle C^{-k}2^{qk}\left\|\Delta_{q}f\right\|_{L^{a}} sup|α|=kαΔqfLaCk2qkΔqfLa.\displaystyle\leq\sup_{|\alpha|=k}\left\|\partial^{\alpha}\Delta_{q}f\right\|_{L^{a}}\leq C^{k}2^{qk}\left\|\Delta_{q}f\right\|_{L^{a}}. (2.3)

2.1.1 Besov and Tilde Spaces

The Besov space Bp,qs(2)B^{s}_{p,q}(\mathbb{R}^{2}) for p,q[1,],sp,q\in[1,\infty],\ s\in\mathbb{R} is the space of distributions uu such that

uBp,rs:=2qsΔquLpqr<.\|u\|_{B_{p,r}^{s}}:=\left\|2^{qs}\left\|\Delta_{q}u\right\|_{L^{p}}\right\|_{\ell^{r}_{q}}<\infty.

In a norm, a subscripted variable qq like qr\ell^{r}_{q} above means that the norm is to be taken with respect to qq. Since we will only consider functions on 2\mathbb{R}^{2}, we will only write Bp,qsB^{s}_{p,q} (and similarly LtpBp,qsL_{t}^{p}B^{s}_{p,q} instead of LtpBp,qs(2)L_{t}^{p}B^{s}_{p,q}(\mathbb{R}^{2})). Also write Hs:=B2,2sH^{s}:=B^{s}_{2,2}. The Besov spaces trivially satisfy

s1s2\displaystyle s_{1}\geq s_{2} Bp,rs1Bp,r~s2,p,r,r~[1,],\displaystyle\implies B^{s_{1}}_{p,r}\hookrightarrow B^{s_{2}}_{p,\tilde{r}},\quad\forall p,r,\tilde{r}\in[1,\infty], (2.4)
r1r2\displaystyle r_{1}\leq r_{2} Bp,r1sBp,r2s,s,p[1,],\displaystyle\implies B^{s}_{p,r_{1}}\hookrightarrow B^{s}_{p,r_{2}},\quad\forall s\in\mathbb{R},\,p\in[1,\infty], (2.5)

and there is also the analogue of Sobolev embedding in 2D:

Bp,rs0+sBp,rs0,1p=1ps2,s0,p,r[1,].\displaystyle B^{s_{0}+s}_{p,r}\hookrightarrow B^{s_{0}}_{p*,r},\quad\frac{1}{p_{*}}=\frac{1}{p}-\frac{s}{2},\quad\forall s_{0}\in\mathbb{R},p,r\in[1,\infty]. (2.6)

In addition, we have the embeddings for any p[1,]p\in[1,\infty],

Bp,10LpBp,0,k0Bp,1kWk,pBp,k.B^{0}_{p,1}\hookrightarrow L^{p}\hookrightarrow B^{0}_{p,\infty},\quad k\in\mathbb{N}_{0}\implies B^{k}_{p,1}\hookrightarrow W^{k,p}\hookrightarrow B^{k}_{p,\infty}.

We say that a function u=u(x,t)u=u(x,t) belongs to the ‘Tilde space’ L~tpBp,qs\widetilde{L}^{p}_{t}B^{s}_{p,q} if

uL~tpBp,rs:=2sqΔqu(x,t)LxpLtpqr.\|u\|_{\widetilde{L}^{p}_{t}B^{s}_{p,r}}:=\left\|2^{sq}\left\|\left\|\Delta_{q}u(x,t)\right\|_{L^{p}_{x}}\right\|_{L^{p}_{t}}\right\|_{\ell^{r}_{q}}.

In comparison with

uLtpBp,rs=u(Bp,rs)xLtp=2sqΔqu(x,t)LxpqrLtp,\|u\|_{L^{p}_{t}B^{s}_{p,r}}=\left\|\left\|u\right\|_{(B^{s}_{p,r})_{x}}\right\|_{L^{p}_{t}}=\left\|\left\|2^{sq}\left\|\Delta_{q}u(x,t)\right\|_{L^{p}_{x}}\right\|_{\ell^{r}_{q}}\right\|_{L^{p}_{t}},

From the generalised Minkowski inequality fLxpLyqfLyqLxp\big{\|}\|f\|_{L^{p}_{x}}\big{\|}_{L^{q}_{y}}\leq\big{\|}\|f\|_{L^{q}_{y}}\big{\|}_{L^{p}_{x}} if qpq\geq p and embeddings between Besov spaces, we have the following relations:

LTρBp,rsL~TρBp,rsLTρBp,rsε,\displaystyle L_{T}^{\rho}B_{p,r}^{s}\hookrightarrow\tilde{L}_{T}^{\rho}B_{p,r}^{s}\hookrightarrow L_{T}^{\rho}B_{p,r}^{s-\varepsilon},\ if rρ,\displaystyle\text{ if }r\geq\rho, (2.7)
LTρBp,rs+εL~TρBp,rsLTρBp,rs,\displaystyle L_{T}^{\rho}B_{p,r}^{s+\varepsilon}\hookrightarrow\tilde{L}_{T}^{\rho}B_{p,r}^{s}\hookrightarrow L_{T}^{\rho}B_{p,r}^{s},\ if ρr,\displaystyle\text{ if }\rho\geq r, (2.8)

where ε>0\varepsilon>0 is arbitrarily small. In particular L~TrBp,rs=LTrBp,rs\tilde{L}_{T}^{r}B_{p,r}^{s}=L_{T}^{r}B_{p,r}^{s}.

3 Study of 2α\mathcal{R}_{2\alpha} and some commutators

3.1 Introduction of the Γ=ω2αθ\Gamma=\omega-\mathcal{R}_{2\alpha}\theta term

We use the technique of [28], [52], [43], and other papers of introducing an auxillary quantity Γ\Gamma determined from the equation, that has better regularity properties than the original functions under consideration. The ω:=u\omega:=\nabla^{\perp}\cdot u equation is obtained by applying the :=(21)\nabla^{\perp}\cdot:=\binom{-\partial_{2}}{\partial_{1}}\cdot operator to the first equation of (1.6). By writing 1θ=Λ2αΘ\partial_{1}\theta=\Lambda^{2\alpha}\Theta, i.e. Θ=2αθ:=Λ12α1θ\Theta=\mathcal{R}_{2\alpha}\theta:=\Lambda^{1-2\alpha}\mathcal{R}_{1}\theta, the ω\omega equation takes the form

(t+u+Λ2α)ωΛ2αΘ=0.(\partial_{t}+u\cdot\nabla+\Lambda^{2\alpha})\omega-\Lambda^{2\alpha}\Theta=0.

By adding (t+u)Θ-(\partial_{t}+u\cdot\nabla)\Theta to both sides of the equation, we obtain an equation for Γ=ωΘ\Gamma=\omega-\Theta which has a commutator:

(t+u+Λ2α)Γ=[2α,u]θ.\displaystyle(\partial_{t}+u\cdot\nabla+\Lambda^{2\alpha})\Gamma=[\mathcal{R}_{2\alpha},u\cdot\nabla]\theta. (3.1)

This structure will be key in deriving the a priori estimates for uu. An important lemma for the study of the critical dissipation case is Lemma 3.3 in Hmidi–Keraani–Rousset’s paper [28]. The generalisation of Lemma 3.3 (ii) is the following result of Wu–Xue [52]:

Proposition 3.1 (Proposition 4.2 of [52]).

Let β[1,2),(p,r)[2,]×[1,),u\beta\in[1,2),(p,r)\in[2,\infty]\times[1,\infty),u be a smooth divergence-free vector field of n\mathbb{R}^{n} (n2)(n\geq 2) with vorticity ω\omega and θ\theta be a smooth scalar function. Then we have that for every s(β2,β),s\in(\beta-2,\beta),

[β,u]θBp,rss,βuLp(θB,rs+1β+θL2).\left\|\left[\mathcal{R}_{\beta},u\cdot\nabla\right]\theta\right\|_{B_{p,r}^{s}}\lesssim_{s,\beta}\|\nabla u\|_{L^{p}}\left(\|\theta\|_{B_{\infty,r}^{s+1-\beta}}+\|\theta\|_{L^{2}}\right).

Besides, if p=,p=\infty, we also have

[β,u]θB,rss,β(ωL+uL2)θB,rs+1β/2+uL2θL2.\left\|\left[\mathcal{R}_{\beta},u\cdot\nabla\right]\theta\right\|_{B_{\infty,r}^{s}}\lesssim_{s,\beta}\left(\|\omega\|_{L^{\infty}}+\|u\|_{L^{2}}\right)\|\theta\|_{B_{\infty,r}^{s+1-\beta/2}}+\|u\|_{L^{2}}\|\theta\|_{L^{2}}.

We will similarly generalise Lemma 3.3 (i) of [28] by using Bony’s decomposition, as follows.

Theorem 3.2.

For any s(0,2α)s\in(0,2\alpha) and any smooth v,θv,\theta with v=0\nabla\cdot v=0,

[2α,v]θHssvL2θB,2s2α+vL2θL2.\|[\mathcal{R}_{2\alpha},v]\theta\|_{H^{s}}\lesssim_{s}\|\nabla v\|_{L^{2}}\|\theta\|_{B^{s-2\alpha}_{\infty,2}}+\|v\|_{L^{2}}\|\theta\|_{L^{2}}.

We prove this result using the following lemma and the equivalent of Hmidi–Keraani–Rousset’s Proposition 3.1 (Wu–Xue’s Proposition 4.1). This lemma is also quoted in [29]. We give its short proof here, which is a variant of Lemma 2.97 of [3].

Lemma 3.3 (Lemma 3.2 of [28]).

Let p[1,]p\in[1,\infty], and let f,gf,g, and hh be three functions such that fLp\nabla f\in L^{p}, gLg\in L^{\infty}, and xhL1xh\in L^{1}. Then

h(fg)f(hg)Lpxh(x)Lx1fLpgL.\displaystyle\|h*(fg)-f(h*g)\|_{L^{p}}\leq\|xh(x)\|_{L^{1}_{x}}\|\nabla f\|_{L^{p}}\|g\|_{L^{\infty}}. (3.2)
Proof.

By a direct computation and the fundamental theorem of calculus, setting F:=h(fg)f(hg)F:=h*(fg)-f(h*g),

F(x)\displaystyle F(x) =nh(xy)(f(y)f(x))g(y)dy\displaystyle=\int_{\mathbb{R}^{n}}h(x-y)\big{(}f(y)-f(x)\big{)}g(y)\mathop{}\!\mathrm{d}y (3.3)
=01nh(xy)f((1τ)x+τy)(yx)g(y)dydτ\displaystyle=\int_{0}^{1}\int_{\mathbb{R}^{n}}h(x-y)\nabla f((1-\tau)x+\tau y)\cdot(y-x)g(y)\mathop{}\!\mathrm{d}y\mathop{}\!\mathrm{d}\tau (3.4)
=01nh(z)f(xτz)zg(xz)dzdτ.\displaystyle=-\int_{0}^{1}\int_{\mathbb{R}^{n}}h(z)\nabla f(x-\tau z)\cdot zg(x-z)\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}\tau. (3.5)

Therefore, Hölder’s inequality and translation invariance of the Lebesgue measure gives

FLp\displaystyle\|F\|_{L^{p}} 01n|zh(z)|f(xτz)g(xz)Lxpdzdτ\displaystyle\leq\int_{0}^{1}\int_{\mathbb{R}^{n}}|zh(z)|\|\nabla f(x-\tau z)g(x-z)\|_{L^{p}_{x}}\mathop{}\!\mathrm{d}z\mathop{}\!\mathrm{d}\tau (3.6)
n|zh(z)|01f(xτz)Lxqdτg(xz)Lxrdz\displaystyle\leq\int_{\mathbb{R}^{n}}|zh(z)|\int_{0}^{1}\|\nabla f(x-\tau z)\|_{L^{q}_{x}}\mathop{}\!\mathrm{d}\tau\|g(x-z)\|_{L^{r}_{x}}\mathop{}\!\mathrm{d}z (3.7)
zh(z)Lz1fLqgLr,\displaystyle\leq\|zh(z)\|_{L^{1}_{z}}\|\nabla f\|_{L^{q}}\|g\|_{L^{r}}, (3.8)

for any 1q+1r=1p\frac{1}{q}+\frac{1}{r}=\frac{1}{p}. In particular, setting q=p,r=q=p,\ r=\infty, we obtain the claimed result. ∎

Proposition 3.4 (Proposition 4.1 of [52]).

Let jj\in\mathbb{N}, α(12,1)\alpha\in(\frac{1}{2},1), and 2α=Λ12α\mathcal{R}_{2\alpha}=\Lambda^{1-2\alpha}\mathcal{R}. Then:

  1. 1.

    For every p(1,)p\in(1,\infty) and q>pq>p defined by 1q=1p2α1n\frac{1}{q}=\frac{1}{p}-\frac{2\alpha-1}{n}, 2α\mathcal{R}_{2\alpha} is a bounded map LpLqL^{p}\to L^{q}.

  2. 2.

    Let χ𝒟(n)\chi\in\mathcal{D}(\mathbb{R}^{n}). Then for each (p,s)[1,]×(2α1,)(p,s)\in[1,\infty]\times(2\alpha-1,\infty) and fLp(n)f\in L^{p}(\mathbb{R}^{n}),

    ||sχ(2j||)2αfLp2j(s+12α)fLp.\displaystyle\||\nabla|^{s}\chi(2^{-j}|\nabla|)\mathcal{R}_{2\alpha}f\|_{L^{p}}\leq 2^{j(s+1-2\alpha)}\|f\|_{L^{p}}. (3.9)

    Moreover, ||sχ(2j||)|\nabla|^{s}\chi(2^{-j}|\nabla|) is a convolution operator with kernel KK satisfying

    |K(x)|1(1+|x|)n+s+1β.|K(x)|\lesssim\frac{1}{(1+|x|)^{n+s+1-\beta}}.
  3. 3.

    Let 𝒪\mathcal{O} be an annulus centered at the origin. Then there exists ϕ𝒮(n)\phi\in\mathcal{S}(\mathbb{R}^{n}) with spectrum supported away from 0 such that for every ff with spectrum in 2j𝒪2^{j}\mathcal{O},

    2αf=2j(n+12α)ϕ(2j)f.\displaystyle\mathcal{R}_{2\alpha}f=2^{j(n+1-2\alpha)}\phi(2^{j}\bullet)*f. (3.10)
Proof of Theorem 3.2.

We use Bony’s decomposition,

[2α,v]θ\displaystyle[\mathcal{R}_{2\alpha},v]\theta =I+II+III,\displaystyle=\mathrm{I}+\mathrm{II}+\mathrm{III}, (3.11)
I\displaystyle\mathrm{I} =q[2α,Sq1v]Δqθ=qIq,\displaystyle=\sum_{q\in\mathbb{N}}[\mathcal{R}_{2\alpha},S_{q-1}v]\Delta_{q}\theta=\sum_{q\in\mathbb{N}}\mathrm{I}_{q}, (3.12)
II\displaystyle\mathrm{II} =q[2α,Δq1v]Sq1θ=qIIq,\displaystyle=\sum_{q\in\mathbb{N}}[\mathcal{R}_{2\alpha},\Delta_{q-1}v]S_{q-1}\theta=\sum_{q\in\mathbb{N}}\mathrm{II}_{q}, (3.13)
III\displaystyle\mathrm{III} =q1[2α,Δqv]Δ~qθ=q1IIIq,\displaystyle=\sum_{q\geq-1}[\mathcal{R}_{2\alpha},\Delta_{q}v]\widetilde{\Delta}_{q}\theta=\sum_{q\geq-1}\mathrm{III}_{q}, (3.14)

where Δ~q:=Δq1+Δq+Δq+1\widetilde{\Delta}_{q}:=\Delta_{q-1}+\Delta_{q}+\Delta_{q+1} (and Δ2:=0\Delta_{-2}:=0). The terms I\mathrm{I} and II\mathrm{II} are low-high and high-low interactions; the term III\mathrm{III} are the interactions at similar frequencies (low-low and high-high).

Estimation of I\mathrm{I}

We have

Iq=2(12α)q(hq(Sq1vΔqθ)(Sq1v)(hqΔqθ)),\displaystyle I_{q}=2^{(1-2\alpha)q}\big{(}h_{q}*(S_{q-1}v\Delta_{q}\theta)-(S_{q-1}v)(h_{q}*\Delta_{q}\theta)\big{)}, (3.15)

for some hq=2dqh(2qx)h_{q}=2^{dq}h(2^{q}x) coming from the third part of Proposition 3.4. By Lemma 3.3,

IqL2\displaystyle\|\mathrm{I}_{q}\|_{L^{2}} 2(12α)qxhqL1Sq1uL2ΔqθL\displaystyle\lesssim 2^{(1-2\alpha)q}\|xh_{q}\|_{L^{1}}\|\nabla S_{q-1}u\|_{L^{2}}\|\Delta_{q}\theta\|_{L^{\infty}} (3.16)
2(2α1)q2quL2ΔqθL\displaystyle\lesssim 2^{-(2\alpha-1)q}2^{-q}\|\nabla u\|_{L^{2}}\|\Delta_{q}\theta\|_{L^{\infty}} (3.17)
=22αquL2ΔqθL.\displaystyle=2^{-2\alpha q}\|\nabla u\|_{L^{2}}\|\Delta_{q}\theta\|_{L^{\infty}}. (3.18)

Therefore, we have (there are no low frequency terms in I\mathrm{I})

IHs2\displaystyle\|{\mathrm{I}}\|_{H^{s}}^{2} q022qsIqL22\displaystyle\sim\sum_{q\geq 0}2^{2qs}\|\mathrm{I}_{q}\|_{L^{2}}^{2} (3.19)
uL22q022q(s2α)ΔqθL2\displaystyle\lesssim\|\nabla u\|_{L^{2}}^{2}\sum_{q\geq 0}2^{2q(s-2\alpha)}\|\Delta_{q}\theta\|_{L^{\infty}}^{2} (3.20)
uL22θB,2s2α2.\displaystyle\leq\|\nabla u\|^{2}_{L^{2}}\|\theta\|^{2}_{B^{s-2\alpha}_{\infty,2}}. (3.21)

Estimation of II\mathrm{II}

Similarly to I\mathrm{I}, we write

IIq=2(12α)q(hq(ΔquSq1θ)(Δqv)(hqSq1θ))\displaystyle\mathrm{II}_{q}=2^{(1-2\alpha)q}\big{(}h_{q}*(\Delta_{q}uS_{q-1}\theta)-(\Delta_{q}v)(h_{q}*S_{q-1}\theta)\big{)} (3.22)

Lemma 3.3 this time gives

IIqL2\displaystyle\|\mathrm{II}_{q}\|_{L^{2}} 2(12α)qxhqL1ΔquL2Sq1θL\displaystyle\lesssim 2^{(1-2\alpha)q}\|xh_{q}\|_{L^{1}}\|\nabla\Delta_{q}u\|_{L^{2}}\|S_{q-1}\theta\|_{L^{\infty}} (3.23)
22αquL2jq2ΔjθL.\displaystyle\lesssim 2^{-2\alpha q}\|\nabla u\|_{L^{2}}\sum_{j\leq q-2}\|\Delta_{j}\theta\|_{L^{\infty}}. (3.24)

We want to use IIHs2q022qsIIqL22\|\mathrm{II}\|_{H^{s}}^{2}\sim\sum_{q\geq 0}2^{2qs}\|\mathrm{II}_{q}\|_{L^{2}}^{2} again; this time we will use the discrete Young’s inequality. Multiplying by 2sq2^{sq}, we note that

2sqIIqL2\displaystyle 2^{sq}\|\mathrm{II}_{q}\|_{L^{2}} =2(2αs)quL2jq2ΔjθL\displaystyle=2^{-(2\alpha-s)q}\|\nabla u\|_{L^{2}}\sum_{j\leq q-2}\|\Delta_{j}\theta\|_{L^{\infty}} (3.25)
uL2(2(2αs)(2(2αs)ΔθL))(q),\displaystyle\lesssim\|\nabla u\|_{L^{2}}\left(2^{-(2\alpha-s)\bullet}\star(2^{-(2\alpha-s)\bullet}\|\Delta_{\bullet}\theta\|_{L^{\infty}})\right)(q), (3.26)

where \star denotes the discrete convolution on 1\mathbb{Z}_{\geq-1},

ab(q)=q1+q2=qaq1bq2.a\star b(q)=\sum_{q_{1}+q_{2}=q}a_{q_{1}}b_{q_{2}}.

Applying the discrete Young’s inequality (note that 2(2αs)1s<2α2^{-(2\alpha-s)\bullet}\in\ell^{1}\iff s<2\alpha) with parameters 1+12=11+121+\frac{1}{2}=\frac{1}{1}+\frac{1}{2},

IIHs\displaystyle\|\mathrm{II}\|_{H^{s}} uL22(2αs)2(2αs)ΔθL(q)2(dq)\displaystyle\lesssim\|\nabla u\|_{L^{2}}\Big{\|}2^{-(2\alpha-s)\bullet}\star 2^{-(2\alpha-s)\bullet}\|\Delta_{\bullet}\theta\|_{L^{\infty}}(q)\Big{\|}_{\ell^{2}(dq)} (3.27)
2αsuL2θB,2(2αs).\displaystyle\lesssim_{2\alpha-s}\|\nabla u\|_{L^{2}}\|\theta\|_{B^{-(2\alpha-s)}_{\infty,2}}. (3.28)

Estimation of III\mathrm{III}

We further split III\mathrm{III} into high-high and low-low interactions,

III\displaystyle\mathrm{III} =J1+J2,\displaystyle=J_{1}+J_{2}, (3.29)
J1\displaystyle J_{1} =q1[2α,Δqv]Δ~qθ,\displaystyle=\sum_{q\geq 1}[\mathcal{R}_{2\alpha},\Delta_{q}v]\widetilde{\Delta}_{q}\theta, (3.30)
J2\displaystyle J_{2} =q0[2α,Δqv]Δ~qθ.\displaystyle=\sum_{q\leq 0}[\mathcal{R}_{2\alpha},\Delta_{q}v]\widetilde{\Delta}_{q}\theta. (3.31)

J1J_{1} with no low frequency terms is dealt with as before, giving

q1[2α,Δqv]Δ~qθL222αqvL2Δ~qθL,\displaystyle q\geq 1\implies\|[\mathcal{R}_{2\alpha},\Delta_{q}v]\widetilde{\Delta}_{q}\theta\|_{L^{2}}\lesssim 2^{-2\alpha q}\|\nabla v\|_{L^{2}}\|\widetilde{\Delta}_{q}\theta\|_{L^{\infty}}, (3.32)

so that

ΔjJ1L2\displaystyle\|\Delta_{j}J_{1}\|_{L^{2}} vL2q:qj422αqΔ~qθL,\displaystyle\lesssim\|\nabla v\|_{L^{2}}\sum_{q:q\geq j-4}2^{-2\alpha q}\|\widetilde{\Delta}_{q}\theta\|_{L^{\infty}}, (3.33)
2jsΔjJ1L2\displaystyle 2^{js}\|\Delta_{j}J_{1}\|_{L^{2}} vL2qj42js22αqΔ~qθL\displaystyle\lesssim\|\nabla v\|_{L^{2}}\sum_{q\geq j-4}2^{js}2^{-2\alpha q}\|\widetilde{\Delta}_{q}\theta\|_{L^{\infty}} (3.34)
=vL2qj42(jq)s2(2αs)qΔ~qθL.\displaystyle=\|\nabla v\|_{L^{2}}\sum_{q\geq j-4}2^{(j-q)s}2^{-(2\alpha-s)q}\|\widetilde{\Delta}_{q}\theta\|_{L^{\infty}}. (3.35)

Under the assumption that s>0s>0, we have by the discrete Young’s inequality again,

J1HsvL2θB,2(2αs).\displaystyle\|J_{1}\|_{H^{s}}\lesssim\|\nabla v\|_{L^{2}}\|\theta\|_{B^{-(2\alpha-s)}_{\infty,2}}. (3.36)

For J2J_{2}, the lowest frequency terms don’t have a corresponding annulus for us to apply Lemma 3.3, so the Riesz transform has to be dealt with in a different way. We proceed without using the commutator structure, instead relying on Bernstein inequalities and L2LpL^{2}\to L^{p} boundedness of 2α\mathcal{R}_{2\alpha} for p(2,)p\in(2,\infty) defined by 1p=122α1d\frac{1}{p}=\frac{1}{2}-\frac{2\alpha-1}{d} (as in Proposition 3.4):

[2α,Δqv]Δ~qθL2\displaystyle\|[\mathcal{R}_{2\alpha},\Delta_{q}v]\widetilde{\Delta}_{q}\theta\|_{L^{2}} ΔqvL2(ΔqθL+2αΔqθL)\displaystyle\lesssim\|\Delta_{q}v\|_{L^{2}}\big{(}\|\Delta_{q}\theta\|_{L^{\infty}}+\|\mathcal{R}_{2\alpha}\Delta_{q}\theta\|_{L^{\infty}}\big{)} (3.37)
=ΔqvL2(ΔqθL+Δq2αθL)\displaystyle=\|\Delta_{q}v\|_{L^{2}}\big{(}\|\Delta_{q}\theta\|_{L^{\infty}}+\|\Delta_{q}\mathcal{R}_{2\alpha}\theta\|_{L^{\infty}}\big{)} (3.38)
vL2(θL2+2αθLp)\displaystyle\lesssim\|v\|_{L^{2}}(\|\theta\|_{L^{2}}+\|\mathcal{R}_{2\alpha}\theta\|_{L^{p}}) (3.39)
vL2θL2.\displaystyle\lesssim\|v\|_{L^{2}}\|\theta\|_{L^{2}}. (3.40)

(Note that all constants from Bernstein inequalities are left implicit because J2J_{2} is a finite sum.) This completes the estimation of III\mathrm{III} and the Theorem is proved. ∎

4 Basic A Priori Estimates

Here we collect some estimates for smooth solutions of (1.6). Since θ\theta solves a transport equation with no diffusion, and by testing the uu equation with itself, we obtain

θ(t)Lp\displaystyle\|\theta(t)\|_{L^{p}} θ0Lp,p[1,],\displaystyle\leq\|\theta_{0}\|_{L^{p}},\qquad p\in[1,\infty], (4.1)
12ddtu(t)L22+Λαu(t)L22\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|u(t)\|_{L^{2}}^{2}+\|\Lambda^{\alpha}u(t)\|_{L^{2}}^{2} θ0L2uL2.\displaystyle\leq\|\theta_{0}\|_{L^{2}}\|u\|_{L^{2}}. (4.2)

These imply

ddtu(t)L2\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|u(t)\|_{L^{2}} θ0L2,\displaystyle\leq\|\theta_{0}\|_{L^{2}}, (4.3)
12ddtu(t)L22+Λαu(t)L22\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|u(t)\|_{L^{2}}^{2}+\|\Lambda^{\alpha}u(t)\|_{L^{2}}^{2} θ0L2(u0L2+θ0L2t),\displaystyle\leq\|\theta_{0}\|_{L^{2}}(\|u_{0}\|_{L^{2}}+\|\theta_{0}\|_{L^{2}}t), (4.4)

and therefore:

Proposition 4.1.

Suppose (u,θ)(u,\theta) are smooth solutions of (1.6) with initial data u0L2u_{0}\in L^{2} and θ0L2Lp\theta_{0}\in L^{2}\cap L^{p}. Then

θ(t)Lp\displaystyle\|\theta(t)\|_{L^{p}} θ0Lp,\displaystyle\leq\|\theta_{0}\|_{L^{p}}, (4.5)
u(t)L22+0tΛαu(s)L22ds\displaystyle\|u(t)\|_{L^{2}}^{2}+\int_{0}^{t}\|\Lambda^{\alpha}u(s)\|_{L^{2}}^{2}\mathop{}\!\mathrm{d}s u0,θ01+t2.\displaystyle\lesssim_{u_{0},\theta_{0}}1+t^{2}. (4.6)
Proposition 4.2.

Suppose (u,θ)(u,\theta) are smooth solutions of (1.6) with initial data u0H1u_{0}\in H^{1} and θL2L\theta\in L^{2}\cap L^{\infty}. Then for ω=u\omega=\nabla^{\perp}\cdot u,

ω(t)L22+0tω(s)2αθ(s)H˙α2dsΦ1(t).\displaystyle\|\omega(t)\|_{L^{2}}^{2}+\int_{0}^{t}\|\omega(s)-\mathcal{R}_{2\alpha}\theta(s)\|_{\dot{H}^{\alpha}}^{2}\mathop{}\!\mathrm{d}s\leq\Phi_{1}(t). (4.7)
Proof.

As mentioned, we set Γ=ω2αθ\Gamma=\omega-\mathcal{R}_{2\alpha}\theta. It solves the equation

(t+u+Λ2α)Γ=[2α,u]θ\displaystyle(\partial_{t}+u\cdot\nabla+\Lambda^{2\alpha})\Gamma=[\mathcal{R}_{2\alpha},u\cdot\nabla]\theta (4.8)

Taking the L2L^{2} inner product with Γ\Gamma and using that u=0\nabla\cdot u=0 in the form of the identity [2α,u]θ=([2α,u]θ)[\mathcal{R}_{2\alpha},u\cdot\nabla]\theta=\nabla\cdot([\mathcal{R}_{2\alpha},u]\theta),

12ddtΓL22+ΓH˙α2\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|\Gamma\|_{L^{2}}^{2}+\|\Gamma\|_{\dot{H}^{\alpha}}^{2} =2([2α,v]θ)Γ\displaystyle=\int_{\mathbb{R}^{2}}\nabla\cdot([\mathcal{R}_{2\alpha},v]\theta)\Gamma (4.9)
[2α,v]θH˙1αΓH˙α.\displaystyle\leq\|[\mathcal{R}_{2\alpha},v]\theta\|_{\dot{H}^{1-\alpha}}\|\Gamma\|_{\dot{H}^{\alpha}}. (4.10)

By Theorem 3.2, Proposition 4.1, and Proposition 3.4,

[2α,u]θH˙1α\displaystyle\|[\mathcal{R}_{2\alpha},u]\theta\|_{\dot{H}^{1-\alpha}} uL2θB,213α+uL2θL2\displaystyle\lesssim\|\nabla u\|_{L^{2}}\|\theta\|_{B^{1-3\alpha}_{\infty,2}}+\|u\|_{L^{2}}\|\theta\|_{L^{2}} (4.11)
ωL2θL+uL2θL2\displaystyle\lesssim\|\omega\|_{L^{2}}\|\theta\|_{L^{\infty}}+\|u\|_{L^{2}}\|\theta\|_{L^{2}} (4.12)
ωL2+1+t\displaystyle\lesssim\|\omega\|_{L^{2}}+1+t (4.13)
ΓL2+2αθL2+1+t\displaystyle\lesssim\|\Gamma\|_{L^{2}}+\|\mathcal{R}_{2\alpha}\theta\|_{L^{2}}+1+t (4.14)
ΓL2+1+t.\displaystyle\lesssim\|\Gamma\|_{L^{2}}+1+t. (4.15)

Therefore,

12ddtΓL22+ΓH˙α2=2([2α,v]θ)Γ(ΓL2+1+t)ΓH˙α.\displaystyle\frac{1}{2}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|\Gamma\|_{L^{2}}^{2}+\|\Gamma\|_{\dot{H}^{\alpha}}^{2}=\int_{\mathbb{R}^{2}}\nabla\cdot([\mathcal{R}_{2\alpha},v]\theta)\Gamma\lesssim(\|\Gamma\|_{L^{2}}+1+t)\|\Gamma\|_{\dot{H}^{\alpha}}. (4.16)

Young’s inequality for products gives

ddtΓ(t)L22+ΓH˙α2ΓL22+1+t2.\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\|\Gamma(t)\|_{L^{2}}^{2}+\|\Gamma\|_{\dot{H}^{\alpha}}^{2}\lesssim\|\Gamma\|_{L^{2}}^{2}+1+t^{2}. (4.17)

An application of Grownwall’s inequality finishes the proof. ∎

5 A Priori Estimates for the Vorticity

Proposition 5.1.

Let (u,θ)(u,\theta) be a smooth solution of (1.6), and let u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p}, with p(2,)p\in(2,\infty), u0=0\nabla\cdot u_{0}=0, and θ0L2L\theta_{0}\in L^{2}\cap L^{\infty}. Then for ω=u\omega=\nabla^{\perp}\cdot u,

ω(t)LpΦ1(t).\displaystyle\|\omega(t)\|_{L^{p}}\leq\Phi_{1}(t). (5.1)
Proof.

We will again use the equation for Γ\Gamma,

tΓ+uΓ+Λ2αΓ=[2α,u]θ\partial_{t}\Gamma+u\cdot\nabla\Gamma+\Lambda^{2\alpha}\Gamma=[\mathcal{R}_{2\alpha},u\cdot\nabla]\theta

From the LpL^{p} estimates in Corollary 3.6 of Cordoba–Cordoba (they prove the estimate for the homogeneous equation Γt+uΓ+Λ2αΓ=0\Gamma_{t}+u\cdot\nabla\Gamma+\Lambda^{2\alpha}\Gamma=0; what we need follows by Duhamel’s principle),

Γ(t)LpΓ0Lp+0t[2α,u]θLpds.\|\Gamma(t)\|_{L^{p}}\leq\|\Gamma_{0}\|_{L^{p}}+\int_{0}^{t}\|[\mathcal{R}_{2\alpha},u\cdot\nabla]\theta\|_{L^{p}}\mathop{}\!\mathrm{d}s.

By Proposition 3.1,

[2α,u]θBp,10uLp(θB,112α+θL2)ωLp(θ0L2+θ0L).\left\|\left[\mathcal{R}_{2\alpha},u\cdot\nabla\right]\theta\right\|_{B_{p,1}^{0}}\lesssim\|\nabla u\|_{L^{p}}\left(\|\theta\|_{B_{\infty,1}^{1-2\alpha}}+\|\theta\|_{L^{2}}\right)\lesssim\|\omega\|_{L^{p}}(\|\theta_{0}\|_{L^{2}}+\|\theta_{0}\|_{L^{\infty}}).

Therefore, using Γ=ω2αθ\Gamma=\omega-\mathcal{R}_{2\alpha}\theta and the boundedness of 2α\mathcal{R}_{2\alpha},

ω(t)Lpθ0,ω01+0tω(s)Lpds,\|\omega(t)\|_{L^{p}}\lesssim_{\theta_{0},\omega_{0}}1+\int_{0}^{t}\|\omega(s)\|_{L^{p}}\mathop{}\!\mathrm{d}s,

and Gronwall’s inequality completes the proof. ∎

The following result is the analogue of Theorem 6.3 in [28].

Theorem 5.2.

Let (u,θ)(u,\theta) solve (1.6) with initial data u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p}, u=0\nabla\cdot u=0, and θ0L2L\theta_{0}\in L^{2}\cap L^{\infty} with p(2,)p\in(2,\infty). Then for r[2,p]r\in[2,p] and ρ[1,r/2)\rho\in[1,r/2),

ω2αθL~tρBr,14α/rΦ1(t).\displaystyle\|\omega-\mathcal{R}_{2\alpha}\theta\|_{\widetilde{L}^{\rho}_{t}B^{4\alpha/r}_{r,1}}\leq\Phi_{1}(t). (5.2)
Proof.

For qq\in\mathbb{N} we set Γq=ΔqΓ.\Gamma_{q}=\Delta_{q}\Gamma. Then, we localise in frequencies the equation (5.1) for Γ\Gamma to get

tΓq+uΓq+Λ2αΓq\displaystyle\partial_{t}\Gamma_{q}+u\cdot\nabla\Gamma_{q}+\Lambda^{2\alpha}\Gamma_{q} =[Δq,u]Γ+Δq([2α,u]θ)\displaystyle=-\left[\Delta_{q},u\cdot\nabla\right]\Gamma+\Delta_{q}([\mathcal{R}_{2\alpha},u\cdot\nabla]\theta)
:=fq.\displaystyle:=f_{q}.

Multiplying the above equation by |Γq|r2Γq\left|\Gamma_{q}\right|^{r-2}\Gamma_{q} and integrating in the space variable we find

1rddtΓq(t)Lrr+2(Λ2αΓq)|Γq|r2ΓqdxΓq(t)Lrr1fq(t)Lr.\frac{1}{r}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\left\|\Gamma_{q}(t)\right\|_{L^{r}}^{r}+\int_{\mathbb{R}^{2}}\left(\Lambda^{2\alpha}\Gamma_{q}\right)\left|\Gamma_{q}\right|^{r-2}\Gamma_{q}\mathop{}\!\mathrm{d}x\leq\left\|\Gamma_{q}(t)\right\|_{L^{r}}^{r-1}\left\|f_{q}(t)\right\|_{L^{r}}.

From [9] or [40], we have the following generalised Bernstein inequality,

d(Λ2αΓq)|Γq|r2Γqdxc22αqΓqLrr,\int_{\mathbb{R}^{d}}\left(\Lambda^{2\alpha}\Gamma_{q}\right)\left|\Gamma_{q}\right|^{r-2}\Gamma_{q}\mathop{}\!\mathrm{d}x\geq c2^{2\alpha q}\left\|\Gamma_{q}\right\|_{L^{r}}^{r},

valid for some c>0c>0 independent of qq, and hence we find

1rddtΓq(t)Lrr+c22αqΓq(t)LrrΓq(t)Lrr1fq(t)Lr.\frac{1}{r}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\left\|\Gamma_{q}(t)\right\|_{L^{r}}^{r}+c2^{2\alpha q}\|\Gamma_{q}(t)\|_{L^{r}}^{r}\leq\left\|\Gamma_{q}(t)\right\|_{L^{r}}^{r-1}\left\|f_{q}(t)\right\|_{L^{r}}.

This yields (writing Γ0:=ω02αθ0\Gamma_{0}:=\omega_{0}-\mathcal{R}_{2\alpha}\theta_{0} and (Γ0)q:=ΔqΓ0(\Gamma_{0})_{q}:=\Delta_{q}\Gamma_{0})

Γq(t)Lrect22αq(Γ0)qLr+0tec(tτ)22αqfq(τ)Lrdτ.\left\|\Gamma_{q}(t)\right\|_{L^{r}}\leq e^{-ct2^{2\alpha q}}\left\|(\Gamma_{0})_{q}\right\|_{L^{r}}+\int_{0}^{t}e^{-c(t-\tau)2^{2\alpha q}}\left\|f_{q}(\tau)\right\|_{L^{r}}\mathop{}\!\mathrm{d}\tau.

By taking the Lρ[0,t]L^{\rho}[0,t] norm and by using convolution inequalities, we find (since fq=[Δq,u]Γ+Δq([2α,u]θ)f_{q}=-\left[\Delta_{q},u\cdot\nabla\right]\Gamma+\Delta_{q}([\mathcal{R}_{2\alpha},u\cdot\nabla]\theta), and ρ[1,p/2]\rho\in[1,p/2])

22αq2rΓqLtρLr\displaystyle 2^{2\alpha q\frac{2}{r}}\left\|\Gamma_{q}\right\|_{L_{t}^{\rho}L^{r}} (5.3)
p22αq(2r1ρ)(Γ0)qLr+22αq(2r1ρ)0t[Δq,u]Γ(τ)Lrdτ\displaystyle\lesssim_{p}2^{2\alpha q\left(\frac{2}{r}-\frac{1}{\rho}\right)}\left\|(\Gamma_{0})_{q}\right\|_{L^{r}}+2^{2\alpha q\left(\frac{2}{r}-\frac{1}{\rho}\right)}\int_{0}^{t}\left\|\left[\Delta_{q},u\cdot\nabla\right]\Gamma(\tau)\right\|_{L^{r}}\mathop{}\!\mathrm{d}\tau (5.4)
+22αq(2r1ρ)0tΔq([2α,u]θ)(τ)Lrdτ.\displaystyle\quad+2^{2\alpha q\left(\frac{2}{r}-\frac{1}{\rho}\right)}\int_{0}^{t}\left\|\Delta_{q}([\mathcal{R}_{2\alpha},u\cdot\nabla]\theta)(\tau)\right\|_{L^{r}}\mathop{}\!\mathrm{d}\tau. (5.5)

To estimate the second integral of the RHS of (5.5), we use Proposition 3.1. This gives uniformly in q0q\geq 0,

Δq[2α,u]θLr\displaystyle\left\|\Delta_{q}\left[\mathcal{R}_{2\alpha},u\cdot\nabla\right]\theta\right\|_{L^{r}} [2α,u]θBr,0\displaystyle\leq\|\left[\mathcal{R}_{2\alpha},u\cdot\nabla\right]\theta\|_{B^{0}_{r,\infty}} (5.6)
uLr(θB,12α+θL2)\displaystyle\leq\|\nabla u\|_{L^{r}}(\|\theta\|_{B_{\infty,\infty}^{1-2\alpha}}+\|\theta\|_{L^{2}}) (5.7)
Φ1(t).\displaystyle\leq\Phi_{1}(t). (5.8)

For the first integral of the RHS of (5.5) we use the following lemma:

Lemma 5.3 (Lemma 6.4 of [28]).

Let vv be a smooth divergence-free vector field and ff be a smooth scalar function. Then, for all a[1,]a\in[1,\infty] and q1q\geq-1,

[Δq,v]fLavLrfBa,12/a.\left\|\left[\Delta_{q},v\cdot\nabla\right]f\right\|_{L^{a}}\lesssim\|\nabla v\|_{L^{r}}\|f\|_{B_{a,1}^{2/a}}.

Using Lemma 5.3, the L2L^{2} bound (Proposition 4.2), and the LpL^{p} bound on ω\omega (Proposition 5.1), we can interpolate (recall r[2,p]r\in[2,p]) to bound vLrΦ1(t)\|\nabla v\|_{L^{r}}\leq\Phi_{1}(t), giving

0t[Δq,u]ΓLrdτuLrΓBr,12/rΦ1(t)ΓL1Br,12/r\displaystyle\int_{0}^{t}\|[\Delta_{q},u\cdot\nabla]\Gamma\|_{L^{r}}\mathop{}\!\mathrm{d}\tau\leq\|\nabla u\|_{L^{r}}\|\Gamma\|_{B_{r,1}^{2/r}}\leq\Phi_{1}(t)\|\Gamma\|_{L^{1}B_{r,1}^{2/r}} (5.9)

Now we show how to estimate the sum in q1q\geq-1 of (5.5). For high frequencies, since r>2ρr>2\rho i.e. 2r1ρ<0\frac{2}{r}-\frac{1}{\rho}<0, we have

qN22αq(2r1ρ)0t[Δq,u]ΓLrdτ2αN(1ρ2r)ΓL1Br,12/r,\displaystyle\sum_{q\geq N}2^{2\alpha q{(\frac{2}{r}-\frac{1}{\rho})}}\int_{0}^{t}\|[\Delta_{q},u\cdot\nabla]\Gamma\|_{L^{r}}\mathop{}\!\mathrm{d}\tau\leq 2^{-\alpha N(\frac{1}{\rho}-\frac{2}{r})}\|\Gamma\|_{L^{1}B^{2/r}_{r,1}}, (5.10)

so that by (5.5), for an NN to be chosen,

qN22αq2rΓqLtρLr\displaystyle\sum_{q\geq N}2^{2\alpha q\frac{2}{r}}\left\|\Gamma_{q}\right\|_{L_{t}^{\rho}L^{r}} (5.11)
Γ0Br,12α(2r1ρ)Γ0Lr+2αN(1ρ2r)Φ1(t)(1+ΓLt1Br,12/r).\displaystyle\leq\underbrace{\|\Gamma_{0}\|_{B^{2\alpha(\frac{2}{r}-\frac{1}{\rho})}_{r,1}}}_{\leq\|\Gamma_{0}\|_{L^{r}}}+2^{-\alpha N(\frac{1}{\rho}-\frac{2}{r})}\Phi_{1}(t)(1+\|\Gamma\|_{L^{1}_{t}B^{2/r}_{r,1}}). (5.12)

(The 11 and ΓLt1Br,12/r\|\Gamma\|_{L^{1}_{t}B^{2/r}_{r,1}} is from the first and second integral terms, respectively.) For low frequencies of Γ\Gamma, we just use

q<N22αq2rΓqLtρLxr22r2αNΓLtρLxr22r2αNΓLtLxrt1/ρ22r2αNΦ1(t).\displaystyle\sum_{q<N}2^{2\alpha q\frac{2}{r}}\|\Gamma_{q}\|_{L^{\rho}_{t}L^{r}_{x}}\leq 2^{\frac{2}{r}2\alpha N}\|\Gamma\|_{L^{\rho}_{t}L^{r}_{x}}\leq 2^{\frac{2}{r}2\alpha N}\|\Gamma\|_{L^{\infty}_{t}L^{r}_{x}}t^{1/\rho}\leq 2^{\frac{2}{r}2\alpha N}\Phi_{1}(t). (5.13)

replacing ΓBr,12/r\|\Gamma\|_{B^{2/r}_{r,1}} on the right by the worse term ΓBr,14α/r\|\Gamma\|_{B^{4\alpha/r}_{r,1}}, we see that

ΓL~ρBr,14α/r\displaystyle\|\Gamma\|_{\widetilde{L}^{\rho}B^{4\alpha/r}_{r,1}} 2αN(1ρ2r)Φ1(t)(1+ΓLt1Br,14α/r)+22r2αNΦ1(t)\displaystyle\leq 2^{-\alpha N(\frac{1}{\rho}-\frac{2}{r})}\Phi_{1}(t)(1+\|\Gamma\|_{L^{1}_{t}B^{4\alpha/r}_{r,1}})+2^{\frac{2}{r}2\alpha N}\Phi_{1}(t) (5.14)
2αN(1ρ2r)Φ1(t)(1+ΓL~tρBr,14α/r)+22r2αNΦ1(t).\displaystyle\leq 2^{-\alpha N(\frac{1}{\rho}-\frac{2}{r})}\Phi_{1}(t)(1+\|\Gamma\|_{\widetilde{L}^{\rho}_{t}B^{4\alpha/r}_{r,1}})+2^{\frac{2}{r}2\alpha N}\Phi_{1}(t). (5.15)

Taking N1N\gg 1, we obtain ΓL~ρBr,14α/rΦ1(t)\|\Gamma\|_{\widetilde{L}^{\rho}B^{4\alpha/r}_{r,1}}\leq\Phi_{1}(t) as claimed. ∎

Proposition 5.4.

Let (u,θ)(u,\theta) be a smooth solution of (1.6), and let u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p}, with p(2,)p\in(2,\infty), u0=0\nabla\cdot u_{0}=0, and θ0L2L\theta_{0}\in L^{2}\cap L^{\infty}. Then we have

uL~tρB,12r(2α1)+1Φ1(t),\displaystyle\|u\|_{\widetilde{L}_{t}^{\rho}B^{\frac{2}{r}(2\alpha-1)+1}_{\infty,1}}\leq\Phi_{1}(t), (5.16)

for every ρ[1,2p2)\rho\in[1,2\wedge\frac{p}{2}) and r>2r>2.

Proof.

The result is stronger for rr closer to 2, so without loss of generality, make rr smaller so that rpr\leq p (this is for the application of Theorem 5.2). By triangle inequality and ω=Γ+2αθ\omega=\Gamma+\mathcal{R}_{2\alpha}\theta,

ωL~tρB,12r(2α1)ΓL~tρB,12r(2α1)+2αθL~tρB,12r(2α1).\displaystyle\|\omega\|_{\widetilde{L}_{t}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}}\leq\|\Gamma\|_{\widetilde{L}_{t}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}}+\|\mathcal{R}_{2\alpha}\theta\|_{\widetilde{L}_{t}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}}. (5.17)

The first term is controlled by some Φ1\Phi_{1} by the embedding L~ρBr,14α/rL~ρB,12r(2α1)\widetilde{L}^{\rho}B^{4\alpha/r}_{r,1}\hookrightarrow\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1} and the previous result, Proposition 5.3. Recalling that 2α=Λ12α\mathcal{R}_{2\alpha}=\Lambda^{1-2\alpha}\mathcal{R}, we can use Bernstein inequalities to see that

2αθL~ρB,12r(2α1)\displaystyle\|\mathcal{R}_{2\alpha}\theta\|_{\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}} =q122r(2α1)qΔq2αθLtρL\displaystyle=\sum_{q\geq-1}2^{\frac{2}{r}(2\alpha-1)q}\|\Delta_{q}\mathcal{R}_{2\alpha}\theta\|_{L^{\rho}_{t}L^{\infty}} (5.18)
2αΔ1θLtρL+q022r(2α1)q(2α1)qΔqθLtρL\displaystyle\lesssim\|\mathcal{R}_{2\alpha}\Delta_{-1}\theta\|_{L^{\rho}_{t}L^{\infty}}+\sum_{q\geq 0}2^{\frac{2}{r}(2\alpha-1)q-(2\alpha-1)q}\|\Delta_{q}\theta\|_{L^{\rho}_{t}L^{\infty}} (5.19)
Δ1θLtρL2+q02(2r1)(2α1)qθLtρL\displaystyle\lesssim\|\Delta_{-1}\theta\|_{L^{\rho}_{t}L^{2}}+\sum_{q\geq 0}2^{(\frac{2}{r}-1)(2\alpha-1)q}\|\theta\|_{L^{\rho}_{t}L^{\infty}} (5.20)
t1/ρθ0L2+t1/ρθ0Lq02(2r1)(2α1)q,\displaystyle\leq t^{1/\rho}\|\theta_{0}\|_{L^{2}}+t^{1/\rho}\|\theta_{0}\|_{L^{\infty}}\sum_{q\geq 0}2^{(\frac{2}{r}-1)(2\alpha-1)q}, (5.21)

where we have used that θ(t)Lpθ0Lp\|\theta(t)\|_{L^{p}}\leq\|\theta_{0}\|_{L^{p}}. Since r>2r>2,

q02(2r1)(2α1)q<.\displaystyle\sum_{q\geq 0}2^{(\frac{2}{r}-1)(2\alpha-1)q}<\infty. (5.22)

Hence, 2αθL~ρB,12r(2α1)θ0t1/ρ\|\mathcal{R}_{2\alpha}\theta\|_{\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}}\lesssim_{\theta_{0}}t^{1/\rho}, and therefore ωL~tρB,12r(2α1)Φ1(t).\|\omega\|_{\widetilde{L}_{t}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}}\leq\Phi_{1}(t). Then we obtain (using Bernstein’s inequality for the low frequency term)

uL~ρB,12r(2α1)+1\displaystyle\|u\|_{\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)+1}_{\infty,1}} Δ1uL~ρL+ωL~ρB,12r(2α1)\displaystyle\lesssim\|\Delta_{-1}u\|_{\widetilde{L}^{\rho}L^{\infty}}+\|\omega\|_{\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}} (5.23)
1+t+ωL~ρB,12r(2α1)\displaystyle\lesssim 1+t+\|\omega\|_{\widetilde{L}^{\rho}B^{\frac{2}{r}(2\alpha-1)}_{\infty,1}} (5.24)
Φ1(t).\displaystyle\leq\Phi_{1}(t). (5.25)

6 Uniqueness of solutions

In order to prove the uniqueness, we will rely on the following Osgood lemma, which can be found as Theorem 5.2.1 in [8], or in [16]:

Lemma 6.1 (Osgood Lemma).

Let γLloc1(+;+),μ\gamma\in L_{\mathrm{loc}}^{1}\left(\mathbb{R}_{+};\mathbb{R}_{+}\right),\mu a continuous non-decreasing function, a+a\in\mathbb{R}_{+} and α\alpha a measurable function satisfying

0α(t)a+0tγ(τ)μ(α(τ))dτ,t+0\leq\alpha(t)\leq a+\int_{0}^{t}\gamma(\tau)\mu(\alpha(\tau))\mathop{}\!\mathrm{d}\tau,\quad\forall t\in\mathbb{R}_{+}

If we assume that a>0a>0 then

(α(t))+(a)0tγ(τ)dτ with (x):=x1drμ(r).-\mathcal{M}(\alpha(t))+\mathcal{M}(a)\leq\int_{0}^{t}\gamma(\tau)\mathop{}\!\mathrm{d}\tau\quad\text{ with }\quad\mathcal{M}(x):=\int_{x}^{1}\frac{\mathop{}\!\mathrm{d}r}{\mu(r)}.

If we assume a=0a=0 and limx0+(x)=+,\lim_{x\rightarrow 0^{+}}\mathcal{M}(x)=+\infty, then α(t)=0,t+\alpha(t)=0,\forall t\in\mathbb{R}_{+}.

Remark 6.2.

In the case μ(x)=x(1logx)\mu(x)=x(1-\log x), an explicit bound is also given in [8]. We will use the similar function μ(x):=xlog(e+1/x)\mu(x):=x\log(e+1/x), for which in the case a<1/ea<1/e, we have the estimate

α(t)exp[exp(e1+(a)0tγ)].\displaystyle\alpha(t)\leq\exp\left[-\exp\left(e-1+\mathcal{M}(a)-\int_{0}^{t}\gamma\right)\right]. (6.1)

This estimate follows elementarily from μ(r)rlogr\mu(r)\geq-r\log r for 0<r<1/e0<r<1/e. In particular, note that α(t)0\alpha(t)\to 0 as a0a\to 0.

We now recall some results (with mild modifications) from Section 4 of Hmidi–Keraani–Rousset’s paper [28].

6.1 Estimates for transport-diffusion models

First, we have a result for transported scalars, which we will apply to the temperature θ\theta of our system.

Proposition 6.3.

Let vv be a smooth divergence-free vector field. Then every scalar solution ψ\psi of the equation

tψ+vψ=f,ψ|t=0=ψ0,\partial_{t}\psi+v\cdot\nabla\psi=f,\quad\psi|_{t=0}=\psi_{0},

satisfies for every p[1,]p\in[1,\infty],

ψ(t)Bp,1CeC0tv(τ)B,11dτ(ψ0Bp,1+0tf(τ)Bp,1dτ).\displaystyle\|\psi(t)\|_{B^{-1}_{p,\infty}}\leq Ce^{C\int_{0}^{t}\|v(\tau)\|_{B^{1}_{\infty,1}}\mathop{}\!\mathrm{d}\tau}\left(\|\psi_{0}\|_{B^{-1}_{p,\infty}}+\int_{0}^{t}\|f(\tau)\|_{B^{-1}_{p,\infty}}\mathop{}\!\mathrm{d}\tau\right). (6.2)

Secondly, we have the following proposition for the linearised velocity equation. The proof is similar to the proof in Hmidi–Keraani–Rousset, so we omit the details. We will only apply this proposition with s=0s=0.

Proposition 6.4.

Let vv be a smooth divergence-free vector field, s(1,1)s\in(-1,1), and ρ[1,]\rho\in[1,\infty]. Let uu be a smooth solution of the linear system

tu+vu+Λ2αu+p=f,u=0.\partial_{t}u+v\cdot\nabla u+\Lambda^{2\alpha}u+\nabla p=f,\quad\nabla\cdot u=0.

Then we have for each t[0,)t\in[0,\infty), with s:=s2α(11ρ)s^{\prime}:=s-2\alpha(1-\frac{1}{\rho}),

uLtB2,sCeC0tv(τ)Ldτ(u0B2,s+(1+t11/ρ)fL~tρB2,s).\|u\|_{L^{\infty}_{t}B^{s}_{2,\infty}}\!\!\leq Ce^{C\int_{0}^{t}\|\nabla v(\tau)\|_{L^{\infty}}\mathop{}\!\mathrm{d}\tau}\!\left(\|u_{0}\|_{B^{s}_{2,\infty}}\!\!+(1+t^{1-1/\rho})\|f\|_{\widetilde{L}^{\rho}_{t}B^{s^{\prime}}_{2,\infty}}\right).

We also need the following estimates adapted from the Appendix of [28]; essentially the same proofs can prove these lemmas, so we will omit them.

Lemma 6.5.

For every s[1,0]s\in[-1,0], if vv is a smooth divergence-free vector field, and ff is a smooth function, then

vfB2,sfB,1+svB2,10.\|v\cdot\nabla f\|_{B^{s}_{2,\infty}}\lesssim\|f\|_{B^{1+s}_{\infty,\infty}}\|v\|_{B^{0}_{2,1}}.
Lemma 6.6.

If vH1v\in H^{1}, then vB2,10vB2,0log(e+vH1vB2,0)\|v\|_{B^{0}_{2,1}}\lesssim\|v\|_{B^{0}_{2,\infty}}\log\left(e+\frac{\|v\|_{H^{1}}}{\|v\|_{B^{0}_{2,\infty}}}\right).

Now we prove the uniqueness of solutions.

Theorem 6.7.

Let α(12,1)\alpha\in(\frac{1}{2},1). Then the equation (1.6) can have at most one solution pair (u,θ)(u,\theta) (in the sense of Definition 1.1) in the space

𝒳t:=(Lt1B,11LtH1)×(LtB2,1Lt1L).\displaystyle\mathcal{X}_{t}:=(L^{1}_{t}B^{1}_{\infty,1}\cap L^{\infty}_{t}H^{1})\times(L^{\infty}_{t}B^{-1}_{2,\infty}\cap L_{t}^{1}L^{\infty}). (6.3)
Proof.

Let (u1,θ1)(u^{1},\theta^{1}) and (u2,θ2)(u^{2},\theta^{2}) be two solutions to (1.6) in the space 𝒳T\mathcal{X}_{T}. Set u=u1u2u=u^{1}-u^{2} and θ=θ1θ2\theta=\theta^{1}-\theta^{2}. Then, they solve

tu+u2u+Λ2αu+p\displaystyle\partial_{t}u+u^{2}\cdot\nabla u+\Lambda^{2\alpha}u+\nabla p =uu1+θe2,\displaystyle=-u\cdot\nabla u^{1}+\theta e_{2}, (6.4)
tθ+u2θ\displaystyle\partial_{t}\theta+u^{2}\cdot\nabla\theta =uθ1,\displaystyle=-u\cdot\nabla\theta^{1}, (6.5)

with zero initial data. For the velocity uu, we write u=V1+V2u=V_{1}+V_{2} where ViV_{i} solve the equations

tV1+u2V1+Λ2αV1+pi\displaystyle\partial_{t}V_{1}+u^{2}\cdot\nabla V_{1}+\Lambda^{2\alpha}V_{1}+\nabla p_{i} =uu1,\displaystyle=-u\cdot\nabla u^{1}, (6.6)
tV2+u2V2+Λ2αV2+pi\displaystyle\partial_{t}V_{2}+u^{2}\cdot\nabla V_{2}+\Lambda^{2\alpha}V_{2}+\nabla p_{i} =θe2,\displaystyle=\theta e_{2}, (6.7)

Proposition 6.4, first with ρ=1\rho=1 and then with ρ=\rho=\infty, gives

V1LtB2,0\displaystyle\|V_{1}\|_{L^{\infty}_{t}B^{0}_{2,\infty}} CeC0tu2(τ)Ldτ(u0B2,0+uu1Lt1B2,0),\displaystyle\leq Ce^{C\int_{0}^{t}\|\nabla u^{2}(\tau)\|_{L^{\infty}}\mathop{}\!\mathrm{d}\tau}\left(\|u_{0}\|_{B^{0}_{2,\infty}}+\|u\cdot\nabla u^{1}\|_{L^{1}_{t}B^{0}_{2,\infty}}\right), (6.8)
V2LtB2,0\displaystyle\|V_{2}\|_{L^{\infty}_{t}B^{0}_{2,\infty}} CeC0tu2(τ)Ldτ(u0B2,0+(1+t)θLtB2,1),\displaystyle\leq Ce^{C\int_{0}^{t}\|\nabla u^{2}(\tau)\|_{L^{\infty}}\mathop{}\!\mathrm{d}\tau}\left(\|u_{0}\|_{B^{0}_{2,\infty}}+(1+t)\|\theta\|_{L^{\infty}_{t}B^{-1}_{2,\infty}}\right), (6.9)

since LtB2,1LtB2,2α=L~tB2,2αL^{\infty}_{t}B^{-1}_{2,\infty}\hookrightarrow L^{\infty}_{t}B^{-2\alpha}_{2,\infty}=\widetilde{L}^{\infty}_{t}B^{-2\alpha}_{2,\infty}. Together, we obtain

uLtB2,0\displaystyle\|u\|_{L^{\infty}_{t}B^{0}_{2,\infty}} CeC0tu2(τ)Ldτ\displaystyle\leq Ce^{C\int_{0}^{t}\|\nabla u^{2}(\tau)\|_{L^{\infty}}\mathop{}\!\mathrm{d}\tau} (6.10)
×(u0B2,0+uu1Lt1B2,0+(1+t)θLtB2,1).\displaystyle\times\left(\|u_{0}\|_{B^{0}_{2,\infty}}+\|u\cdot\nabla u^{1}\|_{L^{1}_{t}B^{0}_{2,\infty}}+(1+t)\|\theta\|_{L^{\infty}_{t}B^{-1}_{2,\infty}}\right).\quad (6.11)

By Lemma 6.5 with s=0s=0, we have

uu1B2,0u1B,1uB2,10.\displaystyle\|u\cdot\nabla u^{1}\|_{B^{0}_{2,\infty}}\lesssim\|u^{1}\|_{B^{1}_{\infty,\infty}}\|u\|_{B^{0}_{2,1}}. (6.12)

By using the logarithmic interpolation inequality of Lemma 6.6, we obtain

uB2,10\displaystyle\|u\|_{B^{0}_{2,1}} uB2,0log(e+1uB2,0)log(e+uH1).\displaystyle\lesssim\|u\|_{B^{0}_{2,\infty}}\log\left(e+\frac{1}{\|u\|_{B^{0}_{2,\infty}}}\right)\log\left(e+\|u\|_{H^{1}}\right). (6.13)

We have used log(e+ab)log(e+a)log(e+b)\log(e+ab)\leq\log(e+a)\log(e+b) which can be proven for instance by using Bernoulli’s inequality for a,b0a,b\geq 0 in the form (1+(1+a)b)(1+a)1+b(1+(1+a)b)\leq(1+a)^{1+b}. Writing μ(a):=alog(e+1/a)\mu(a):=a\log(e+1/a), the above inequality can be rewritten as

uB2,10μ(uB2,0)log(e+uH1).\displaystyle\|u\|_{B^{0}_{2,1}}\lesssim\mu(\|u\|_{B^{0}_{2,\infty}})\log\left(e+\|u\|_{H^{1}}\right). (6.14)

For the temperature, Proposition 6.3 with p=2p=2 gives

θLtB2,1CeCu2Lt1B,11(θ0B2,1+0tvθ1B2,1dτ).\displaystyle\|\theta\|_{L^{\infty}_{t}B^{-1}_{2,\infty}}\!\!\leq Ce^{C\|u^{2}\|_{L^{1}_{t}B^{1}_{\infty,1}}}\!\!\left(\|\theta_{0}\|_{B^{-1}_{2,\infty}}\!\!+\int_{0}^{t}\!\|v\cdot\nabla\theta^{1}\|_{B^{-1}_{2,\infty}}\mathop{}\!\mathrm{d}\tau\right).\quad (6.15)

By Lemma 6.5 with s=1s=-1,

uθ1B2,1θ1B,0uB2,10θ0LuB2,10.\displaystyle\|u\cdot\nabla\theta^{1}\|_{B^{-1}_{2,\infty}}\leq\|\theta^{1}\|_{B^{0}_{\infty,\infty}}\|u\|_{B^{0}_{2,1}}\leq\|\theta_{0}\|_{L^{\infty}}\|u\|_{B^{0}_{2,1}}. (6.16)

By combining (6.11) with (6.12), (6.14), (6.15), (6.16), we obtain

uLtB2,0+θLtB2,1\displaystyle\|u\|_{L^{\infty}_{t}B^{0}_{2,\infty}}+\|\theta\|_{L^{\infty}_{t}B^{-1}_{2,\infty}} (6.17)
CeCu2Lt1B,11(θ0B2,1+θ0L0tμ(uB2,0)log(e+uH1)dτ)\displaystyle\leq Ce^{C\|u^{2}\|_{L^{1}_{t}B^{1}_{\infty,1}}}\left(\|\theta_{0}\|_{B^{-1}_{2,\infty}}+\|\theta_{0}\|_{L^{\infty}}\int_{0}^{t}\mu(\|u\|_{B^{0}_{2,\infty}})\log\left(e+\|u\|_{H^{1}}\right)\mathop{}\!\mathrm{d}\tau\right) (6.18)
+CeCu2(τ)Lt1L(u0B2,0+0tu1B,1μ(uB2,0)log(e+uH1)dτ\displaystyle+Ce^{C\|\nabla u^{2}(\tau)\|_{L^{1}_{t}L^{\infty}}}\Bigg{(}\|u_{0}\|_{B^{0}_{2,\infty}}\!+\int_{0}^{t}\|u^{1}\|_{B^{1}_{\infty,\infty}}\mu(\|u\|_{B^{0}_{2,\infty}})\log\left(e+\|u\|_{H^{1}}\right)\mathop{}\!\mathrm{d}\tau (6.19)
+C(1+t)eCu2Lt1B,11[θ0B2,1+θ0L0tμ(uB2,0)log(e+uH1)dτ])\displaystyle+C(1+t)e^{C\|u^{2}\|_{L^{1}_{t}B^{1}_{\infty,1}}}\left[\|\theta_{0}\|_{B^{-1}_{2,\infty}}\!+\|\theta_{0}\|_{L^{\infty}}\int_{0}^{t}\mu(\|u\|_{B^{0}_{2,\infty}})\log\left(e+\|u\|_{H^{1}}\right)\mathop{}\!\mathrm{d}\tau\right]\Bigg{)} (6.20)
C(1+t)eCu2Lt1B,11(u0B2,0+θ0B2,1+log(e+uLtH1)(1+θ0L)\displaystyle\leq C(1+t)e^{C\|u^{2}\|_{L^{1}_{t}B^{1}_{\infty,1}}}\Bigg{(}\|u_{0}\|_{B^{0}_{2,\infty}}+\|\theta_{0}\|_{B^{-1}_{2,\infty}}+\log(e+\|u\|_{L^{\infty}_{t}H^{1}})(1+\|\theta_{0}\|_{L^{\infty}}) (6.21)
×0t(1+u1(τ)B,1)μ(uLτB2,0+θLτB2,1)dτ).\displaystyle\times\int_{0}^{t}(1+\|u^{1}(\tau)\|_{B^{1}_{\infty,\infty}})\mu(\|u\|_{L^{\infty}_{\tau}B^{0}_{2,\infty}}+\|\theta\|_{L^{\infty}_{\tau}B^{-1}_{2,\infty}})\mathop{}\!\mathrm{d}\tau\Bigg{)}. (6.22)

Setting

X(t)\displaystyle X(t) :-uLtB2,0+θLtB2,1,\displaystyle\coloneq\|u\|_{L^{\infty}_{t}B^{0}_{2,\infty}}+\|\theta\|_{L^{\infty}_{t}B^{-1}_{2,\infty}}, (6.23)
f(t)\displaystyle f(t) :-C(1+t)eCu2Lt1B,11+log(e+uLtH1)(1+θ0L),\displaystyle\coloneq C(1+t)e^{C\|u^{2}\|_{L^{1}_{t}B^{1}_{\infty,1}}}+\log(e+\|u\|_{L^{\infty}_{t}H^{1}})(1+\|\theta_{0}\|_{L^{\infty}}), (6.24)
g(t)\displaystyle g(t) :-1+u1(τ)B,1,\displaystyle\coloneq 1+\|u^{1}(\tau)\|_{B^{1}_{\infty,\infty}}, (6.25)

We obtain the following integral inequality for XX:

X(t)f(t)(X(0)+0tg(τ)μ(X(τ))dτ).\displaystyle X(t)\leq f(t)\left(X(0)+\int_{0}^{t}g(\tau)\mu(X(\tau))\mathop{}\!\mathrm{d}\tau\right). (6.26)

Since X(0)=0X(0)=0, by Lemma 6.1, X(t)=0X(t)=0 for every tt, which proves the uniqueness of solutions. ∎

7 Existence of solutions

We now use the above results to prove our main technical result.

Proof of Theorem 1.2.

The uniqueness is guaranteed by Theorem 6.7 so we focus on the existence. Following [6] (or Chapter 3 of [41]), if the initial data are smooth, we can obtain smooth solutions in HmH^{m} to (1.6). By the a priori estimate of Proposition 5.4 and the Beale–Kato–Majda type blow-up criterion [6], the solution is globally defined.

Now let u0H1W˙1,pu_{0}\in H^{1}\cap\dot{W}^{1,p} with u0=0\nabla\cdot u_{0}=0, for some p>2p>2, and θ0L1L\theta_{0}\in L^{1}\cap L^{\infty}. Consider the sequence of initial data u0,n,θ0,nu_{0,n},\theta_{0,n} defined by Littlewood–Paley projections,

u0,n\displaystyle u_{0,n} :-Snu0,\displaystyle\coloneq S_{n}u_{0}, (7.1)
θ0,n\displaystyle\theta_{0,n} :-Snθ0.\displaystyle\coloneq S_{n}\theta_{0}. (7.2)

These functions are smooth, so they define smooth solutions un,θnu_{n},\theta_{n}. Moreover, we have a priori estimates uniform in nn that imply for any ε>0\varepsilon>0 and ρ[1,2p2)\rho\in[1,2\wedge\frac{p}{2}),

θn\displaystyle\theta_{n} Lt(L1L),\displaystyle\in L^{\infty}_{t}(L^{1}\cap L^{\infty}), (7.3)
un\displaystyle u_{n} LtW1,pLtρB,12αε.\displaystyle\in L^{\infty}_{t}W^{1,p}\cap L^{\rho}_{t}B^{2\alpha-\varepsilon}_{\infty,1}. (7.4)

The spaces B,12αεB^{2\alpha-\varepsilon}_{\infty,1} and C2αε=B,2αεC^{2\alpha-\varepsilon}=B^{2\alpha-\varepsilon}_{\infty,\infty} are essentially equivalent by standard embeddings of Besov spaces due to the small loss in ε\varepsilon, so we can replace one with the other at any point.

Up to subsequences, we have that θn\theta_{n} and unu_{n} converge weakly to functions θ\theta and uu. By taking (Snu0Smu0,Snθ0Smθ0)(S_{n}u_{0}-S_{m}u_{0},S_{n}\theta_{0}-S_{m}\theta_{0}) as initial data, the proof of Theorem 6.7 and Remark 6.2 gives that as soon as

Snu0Smu0B2,0+Snθ0Smθ0B2,11/e,\|S_{n}u_{0}-S_{m}u_{0}\|_{B^{0}_{2,\infty}}+\|S_{n}\theta_{0}-S_{m}\theta_{0}\|_{B^{-1}_{2,\infty}}\leq 1/e,

we obtain

unumLtB2,0+θnθmLTB2,1\displaystyle\|u_{n}-u_{m}\|_{L^{\infty}_{t}B^{0}_{2,\infty}}+\|\theta_{n}-\theta_{m}\|_{L^{\infty}_{T}B^{-1}_{2,\infty}} (7.5)
F(t,Snu0Smu0B2,0+Snθ0Smθ0B2,1),\displaystyle\leq F(t,\|S_{n}u_{0}-S_{m}u_{0}\|_{B^{0}_{2,\infty}}+\|S_{n}\theta_{0}-S_{m}\theta_{0}\|_{B^{-1}_{2,\infty}}), (7.6)

for an explicit function FF given by (6.1) with F(t,s)0F(t,s)\to 0 as s0s\to 0. Thus umu_{m} is Cauchy and hence strongly convergent to uu in LtB2,0L^{\infty}_{t}B^{0}_{2,\infty}. Interpolation yields strong convergence of unu_{n} to uu in L2([0,t]×2)L^{2}([0,t]\times\mathbb{R}^{2}). This implies ununu_{n}\otimes u_{n} to uuu\otimes u strongly in L1([0,t]×2)L^{1}([0,t]\times\mathbb{R}^{2}). Also, since θnθ\theta_{n}\rightharpoonup\theta in L2L^{2}, the product unθnuθu_{n}\theta_{n}\rightharpoonup u\theta in L1L^{1}. Thus, all terms in Definition 1.1 make sense, and (u,θ)(u,\theta) is a solution to (1.6). ∎

Finally, for completeness, we give a standard argument that shows the boundary regularity for the temperature patches is preserved.

Proof of Theorem 1.3.

Let X=X(t,ξ)X=X(t,\xi) denote the flow of the vector field uLt1C2αu\in L^{1}_{t}C^{2\alpha^{\prime}}, i.e. the solution of

Xt(t,ξ)\displaystyle X_{t}(t,\xi) =u(X(t,ξ),t),\displaystyle=u(X(t,\xi),t), (7.7)
X(0,ξ)\displaystyle X(0,\xi) =ξ.\displaystyle=\xi. (7.8)

Taking the gradient in ξ\xi gives

(ξX)t=u(X,t)ξX,\displaystyle(\nabla_{\xi}X)_{t}=\nabla u(X,t)\nabla_{\xi}X, (7.9)

so that

ξX(t,ξ)LξαX(0,ξ)Lξe0tu(s)Lds,\|\nabla_{\xi}X(t,\xi)\|_{L^{\infty}_{\xi}}\leq\|\nabla_{\alpha}X(0,\xi)\|_{L^{\infty}_{\xi}}e^{\int_{0}^{t}\|\nabla u(s)\|_{L^{\infty}}\mathop{}\!\mathrm{d}s},\\

and similarly,

[ξX(t,ξ)]C2α1ξX(0,ξ)C2α1eC0tu(s)C2αds.[\nabla_{\xi}X(t,\xi)]_{C^{2\alpha^{\prime}-1}}\leq\|\nabla_{\xi}X(0,\xi)\|_{C^{2\alpha^{\prime}-1}}e^{C\int_{0}^{t}\|u(s)\|_{C^{2\alpha^{\prime}}}\mathop{}\!\mathrm{d}s}.

This shows that the flow remains in C2α1C^{2\alpha^{\prime}-1}. To apply this to the boundary of the patch, suppose that at time t=0t=0, D0\partial D_{0} is parameterised by the curve γ0C2α([0,1];2)\gamma_{0}\in C^{2\alpha^{\prime}}([0,1];\mathbb{R}^{2}). Then at time tt, the flow transports it to the curve

γ(t,s):=X(t,γ0(s)).\gamma(t,s):=X(t,\gamma_{0}(s)).

Then sγ=ξX(t,γ0(s))γ0\partial_{s}\gamma=\nabla_{\xi}X(t,\gamma_{0}(s))\gamma_{0}^{\prime}, and the Hölder seminorm of sγ\partial_{s}\gamma is controlled:

[sγ]C2α1[ξX(t,γ0)]C2α1γ0L2+ξXL[γ0]C2α1.[\partial_{s}\gamma]_{C^{2\alpha^{\prime}-1}}\leq[\nabla_{\xi}X(t,\gamma_{0})]_{C^{2\alpha^{\prime}-1}}\|\gamma_{0}^{\prime}\|_{L^{\infty}}^{2}+\|\nabla_{\xi}X\|_{L^{\infty}}[\gamma_{0}^{\prime}]_{C^{2\alpha^{\prime}-1}}.

This proves that the Hölder regularity is preserved for all time. ∎

Acknowledgements

Both authors were partially supported by the NSF of China (Grant No. 11771045, 11871087).

References

  • [1] H. Abidi and T. Hmidi “On the global well-posedness for Boussinesq system” In Journal of Differential Equations 233.1, 2007, pp. 199–220 DOI: 10.1016/j.jde.2006.10.008
  • [2] H. Abidi and P. Zhang “On the global well-posedness of 2-D Boussinesq system with variable viscosity” In Advances in Mathematics 305, 2017, pp. 1202–1249 DOI: 10.1016/j.aim.2016.09.036
  • [3] H. Bahouri, J.Y. Chemin and R. Danchin “Fourier Analysis and Nonlinear Partial Differential Equations”, Grundlehren der mathematischen Wissenschaften Springer Berlin Heidelberg, 2011 DOI: 10.1007/978-3-642-16830-7
  • [4] D. Chae “Global regularity for the 2D Boussinesq equations with partial viscosity terms” In Advances in Mathematics 203.2, 2006, pp. 497–513 DOI: 10.1016/j.aim.2005.05.001
  • [5] D. Chae, P. Constantin and J. Wu “Inviscid Models Generalizing the Two-dimensional Euler and the Surface Quasi-Geostrophic Equations” In Archive for Rational Mechanics and Analysis 202.1, 2011, pp. 35–62 DOI: 10.1007/s00205-011-0411-5
  • [6] D. Chae and H.-S. Nam “Local existence and blow-up criterion for the Boussinesq equations” In Proceedings of the Royal Society of Edinburgh: Section A Mathematics 127.5 Royal Society of Edinburgh Scotland Foundation, 1997, pp. 935–946 DOI: 10.1017/S0308210500026810
  • [7] D. Chae et al. “Generalized Surface Quasi-Geostrophic equations with singular velocities” In Communications on Pure and Applied Mathematics 65.8 Wiley Subscription Services, Inc., A Wiley Company, 2012, pp. 1037–1066 DOI: 10.1002/cpa.21390
  • [8] Jean-Yves Chemin “Perfect incompressible fluids”, Oxford Lecture Series in Mathematics and Its Applications 14 Oxford University Press, 1998
  • [9] Q. Chen, C. Miao and Z. Zhang “A New Bernstein’s Inequality and the 2D Dissipative Quasi-Geostrophic Equation” In Communications in Mathematical Physics 271.3, 2007, pp. 821–838 DOI: 10.1007/s00220-007-0193-7
  • [10] P. Constantin and J. Wu “Inviscid limit for vortex patches” In Nonlinearity 8.5, 1995, pp. 735–742 DOI: 10.1088/0951-7715/8/5/005
  • [11] P. Constantin and J. Wu “The inviscid limit for non-smooth vorticity” In Indiana Univ. Math. J. 45, 1996, pp. 67–82
  • [12] A. Córdoba, D. Córdoba and F. Gancedo “Interface evolution: the Hele–Shaw and Muskat problems” In Ann. of Math. (2) 173.1, 2011, pp. 477–542 DOI: 10.4007/annals.2011.173.1.10
  • [13] A. Córdoba, D. Córdoba and F. Gancedo “Uniqueness for SQG patch solutions” In ArXiv e-print, 2016 arXiv:1605.06663 [math.AP]
  • [14] R. Danchin and M. Paicu “Les théorèmes de Leray et de Fujita–Kato pour le système de Boussinesq partiellement visqueux” In Bulletin de la Société Mathématique de France 136.2 Société mathématique de France, 2008, pp. 261–309 DOI: 10.24033/bsmf.2557
  • [15] R. Danchin and X. Zhang “Global persistence of geometrical structures for the Boussinesq equation with no diffusion” In Communications in Partial Differential Equations 42.1 Taylor & Francis, 2017, pp. 68–99 DOI: 10.1080/03605302.2016.1252394
  • [16] T.. Elgindi “Osgood’s Lemma and some results for the Slightly Supercritical 2D Euler Equations for Incompressible Flow” In Archive for Rational Mechanics and Analysis 211.3, 2014, pp. 965–990 DOI: 10.1007/s00205-013-0691-z
  • [17] L.. Evans “Partial Differential Equations” In Graduate Studies in Mathematics 19 American Mathematical Society, 2010
  • [18] C. Fefferman, G. Luli and J. Rodrigo “The spine of an SQG almost-sharp front” In Nonlinearity 25.2 IOP Publishing, 2012, pp. 329 DOI: 10.1088/0951-7715/25/2/329
  • [19] C. Fefferman and J.. Rodrigo “Almost sharp fronts for SQG: the limit equations” In Communications in Mathematical Physics 313.1 Springer, 2012, pp. 131–153 DOI: 10.1007/s00220-012-1486-z
  • [20] C. Fefferman and J.. Rodrigo “Analytic sharp fronts for the Surface Quasi-Geostrophic equation” In Communications in Mathematical Physics 303.1 Springer, 2011, pp. 261–288 DOI: 10.1007/s00220-011-1190-4
  • [21] F. Gancedo “Existence for the α\alpha-patch model and the QG sharp front in Sobolev spaces” In Advances in Mathematics 217.6, 2008, pp. 2569–2598 DOI: 10.1016/j.aim.2007.10.010
  • [22] F. Gancedo and E. García-Juárez “Global Regularity for 2D Boussinesq Temperature Patches with No Diffusion” In Annals of PDE 3.2, 2017, pp. 14 DOI: 10.1007/s40818-017-0031-y
  • [23] F. Gancedo and E. García-Juárez “Global Regularity of 2D Density Patches for Inhomogeneous Navier–Stokes” In Archive for Rational Mechanics and Analysis 229.1, 2018, pp. 339–360 DOI: 10.1007/s00205-018-1218-4
  • [24] F. Gancedo and E. García-Juárez “Regularity Results for Viscous 3D Boussinesq Temperature Fronts” In Communications in Mathematical Physics 376.3, 2020, pp. 1705–1736 DOI: 10.1007/s00220-020-03767-4
  • [25] F. Gancedo and N. Patel “On the local existence and blow-up for generalized SQG patches” In arXiv e-prints, 2018, pp. arXiv:1811.00530 arXiv:1811.00530 [math.AP]
  • [26] F. Gancedo, E. Garcia-Juarez, N. Patel and R. Strain “Global Regularity for Gravity Unstable Muskat Bubbles”, 2019 arXiv:1902.02318 [math.AP]
  • [27] T. Hmidi and S. Keraani “On the global well-posedness of the two-dimensional Boussinesq system with a zero diffusivity” In Adv. Differential Equations 12.4 Khayyam Publishing, Inc., 2007, pp. 461–480 URL: https://projecteuclid.org:443/euclid.ade/1355867459
  • [28] T. Hmidi, S. Keraani and F. Rousset “Global well-posedness for a Boussinesq–Navier–Stokes system with critical dissipation” In Journal of Differential Equations 249.9, 2010, pp. 2147–2174 DOI: 10.1016/j.jde.2010.07.008
  • [29] T. Hmidi, S. Keraani and F. Rousset “Global Well-Posedness for Euler–Boussinesq System with Critical Dissipation” In Communications in Partial Differential Equations 36.3 Taylor & Francis, 2010, pp. 420–445 DOI: 10.1080/03605302.2010.518657
  • [30] T. Hmidi and M. Zerguine “On the global well-posedness of the Euler–Boussinesq system with fractional dissipation” Evolution Equations in Pure and Applied Sciences In Physica D: Nonlinear Phenomena 239.15, 2010, pp. 1387–1401 DOI: 10.1016/j.physd.2009.12.009
  • [31] T.. Hou and C. Li “Global well-posedness of the viscous Boussinesq equations” In Discrete Contin. Dyn. Syst. 12.1, 2005, pp. 1–12 DOI: 10.3934/dcds.2005.12.1
  • [32] J.. Hunter and J. Shu “Regularized and approximate equations for sharp fronts in the Surface Quasi-Geostrophic equation and its generalizations” In Nonlinearity 31.6 IOP Publishing, 2018, pp. 2480–2517 DOI: 10.1088/1361-6544/aab1cc
  • [33] J.. Hunter, J. Shu and Q. Zhang “Contour Dynamics for Surface Quasi-Geostrophic Fronts”, 2019 arXiv:1907.06593 [math.AP]
  • [34] J.. Hunter, J. Shu and Q. Zhang “Global Solutions of a Surface Quasi-Geostrophic Front Equation”, 2018 arXiv:1808.07631 [math.AP]
  • [35] J.. Hunter, J. Shu and Q. Zhang “Local Well-Posedness of an Approximate Equation for SQG Fronts” In Journal of Mathematical Fluid Mechanics 20.4 Springer ScienceBusiness Media LLC, 2018, pp. 1967–1984 DOI: 10.1007/s00021-018-0396-z
  • [36] J.. Hunter, J. Shu and Q. Zhang “Two-Front Solutions of the SQG Equation and its Generalizations”, 2019 arXiv:1904.13380 [math.AP]
  • [37] Q. Jiu and Y. Wang “Inviscid limit for vortex patches in a bounded domain” In Applied Mathematics Letters 25.10, 2012, pp. 1367–1372 DOI: 10.1016/j.aml.2011.12.003
  • [38] C. Khor and J.. Rodrigo “Local Existence of Analytic Sharp Fronts for Singular SQG”, 2020 arXiv:2001.10412 [math.AP]
  • [39] C. Khor and J.. Rodrigo “On Sharp Fronts and Almost-Sharp Fronts for singular SQG”, 2020 arXiv:2001.10332 [math.AP]
  • [40] D. Li “On a frequency localized Bernstein inequality and some generalized Poincaré-type inequalities” In Mathematical Research Letters 20.5 International Press of Boston, 2013, pp. 933–945 DOI: 10.4310/MRL.2013.v20.n5.a9
  • [41] A. Majda and A. Bertozzi “Vorticity and Incompressible Flow” Cambridge Texts in Applied Mathematics, 2002
  • [42] O. Melkemi and M. Zerguine “Local persistence of geometric structures of the inviscid nonlinear Boussinesq system”, 2020 arXiv:2005.11605 [math.AP]
  • [43] C. Miao and L. Xue “On the global well-posedness of a class of Boussinesq–Navier–Stokes systems” In Nonlinear Differential Equations and Applications NoDEA 18.6, 2011, pp. 707–735 DOI: 10.1007/s00030-011-0114-5
  • [44] J Pedlosky “Geophysical Fluid Dynamics” Springer-Verlag New York, 1987 DOI: 10.1007/978-1-4612-4650-3
  • [45] J. Peetre “New Thoughts on Besov Spaces”, Duke University mathematics series Mathematics Department, Duke University, 1976 URL: https://books.google.com/books?id=-kEeAQAAIAAJ
  • [46] J.. Rodrigo “On the evolution of Sharp Fronts for the Quasi-Geostrophic equation” In Communications on Pure and Applied Mathematics 58.6 New York: Interscience Publishers, c1949-, 2005, pp. 821–866 DOI: 10.1002/cpa.20059
  • [47] J.. Rodrigo “The vortex patch problem for the surface quasi-geostrophic equation” In Proceedings of the National Academy of Sciences 101.9 National Acad Sciences, 2004, pp. 2684–2686 DOI: 10.1073/pnas.0308158101
  • [48] W. Rudin “Functional Analysis”, International series in pure and applied mathematics McGraw-Hill, 2006 URL: https://books.google.com/books?id=l7XFfDmjp5IC
  • [49] Y. Sawano “Theory of Besov Spaces”, Developments in Mathematics Springer Singapore, 2018 URL: https://books.google.com/books?id=lw92%7BD%7DwAAQBAJ
  • [50] A. Stefanov and J. Wu “A global regularity result for the 2D Boussinesq equations with critical dissipation” In Journal d’Analyse Mathématique 137.1, 2019, pp. 269–290 DOI: 10.1007/s11854-018-0073-4
  • [51] T. Tao “Nonlinear Dispersive Equations: Local and Global Analysis”, Conference Board of the Mathematical Sciences. Regional conference series in mathematics American Mathematical Society, 2006 URL: https://books.google.com/books?id=pUDiDgAAQBAJ
  • [52] G. Wu and L. Xue “Global well-posedness for the 2D inviscid Bénard system with fractional diffusivity and Yudovich’s type data” In Journal of Differential Equations 253.1, 2012, pp. 100–125 DOI: https://doi.org/10.1016/j.jde.2012.02.025
  • [53] J. Wu and X. Xu “Well-posedness and inviscid limits of the Boussinesq equations with fractional Laplacian dissipation” In Nonlinearity 27.9 IOP Publishing, 2014, pp. 2215–2232 DOI: 10.1088/0951-7715/27/9/2215
  • [54] X. Xu “Global regularity of solutions of 2D Boussinesq equations with fractional diffusion” In Nonlinear Analysis: Theory, Methods & Applications 72.2, 2010, pp. 677–681 DOI: 10.1016/j.na.2009.07.008
  • [55] M. Zerguine “The regular vortex patch problem for stratified Euler equations with critical fractional dissipation” In Journal of Evolution Equations 15.3, 2015, pp. 667–698 DOI: 10.1007/s00028-015-0277-3

C. Khor
Laboratory of Mathematics and Complex Systems (Ministry of Education), School of Mathematical Sciences, Beijing Normal University, Beijing 100875, People’s Republic of China. [email protected]

X. Xu
Laboratory of Mathematics and Complex Systems (Ministry of Education), School of Mathematical Sciences, Beijing Normal University, Beijing 100875, People’s Republic of China. [email protected]