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

Weak and mild solutions to the MHD equations and the viscoelastic Navier–Stokes equations with damping in Wiener amalgam spaces

Chen-Chih Lai Email address: cclai.math@gmail.com Department of Mathematics, Columbia University, New York, NY 10027, USA
Abstract

We study the three-dimensional incompressible magnetohydrodynamic (MHD) equations and the incompressible viscoelastic Navier–Stokes equations with damping. Building on techniques developed by Bradshaw, Lai, and Tsai (Math. Ann. 2024), we prove the existence of mild solutions in Wiener amalgam spaces that satisfy the corresponding spacetime integral bounds. In addition, we construct global-in-time local energy weak solutions in these amalgam spaces using the framework introduced by Bradshaw and Tsai (SIAM J. Math. Anal. 2021). As part of this construction, we also establish several properties of local energy solutions with Luloc2L^{2}_{\mathrm{uloc}} initial data, including initial and eventual regularity as well as small-large uniqueness, extending analogous results obtained for the Navier–Stokes equations by Bradshaw and Tsai (Comm. Partial Differential Equations 2020).

1 Introduction

The incompressible magnetohydrodynamic (MHD) equations describe the interaction of a fluid’s velocity field and a magnetic field within a conducting medium, coupling the incompressible Navier–Stokes equations with Maxwell’s equations of electromagnetism. These fundamental equations are given by

tvΔv+vvbb+π=0tbΔb+vbbv=0v=b=0} in 3×(0,),\left.\begin{array}[]{ll}\partial_{t}v-\Delta v+v\cdot\nabla v-b\cdot\nabla b+\nabla\pi&=0\\ \partial_{t}b-\Delta b+v\cdot\nabla b-b\cdot\nabla v&=0\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\,\nabla\cdot v=\nabla\cdot b&=0\end{array}\right\}\text{ in }\mathbb{R}^{3}\times(0,\infty), (MHD)

where vv is the velocity, bb the magnetic field, and π\pi the pressure. The study of MHD equations has attracted considerable attention. The foundational result of Duvaut and Lions [11] established the global existence of weak solutions with finite energy. Building upon this, Sermange and Temam [38] investigated regularity criteria for weak solutions. Subsequent effort refined these regularity conditions under various assumptions. For instance, Wu [41] and Zhou [44] established Serrin-type criteria and scaling-invariant regularity conditions, while He and Xin [15] and Kang and Lee [19] developed partial regularity results for suitable weak solutions. Further improvements were made via harmonic analysis methods, as in [9], and directionally-restricted criteria, such as those by Cao and Wu [7]. More recently, local regularity theory for MHD has been advanced in parabolic Morrey spaces [8], and Fernández-Dalgo and Jarrín [13] provided weak-strong uniqueness results in weighted L2L^{2} spaces, alongside constructions of weak suitable solutions in local Morrey spaces.

On the existence side, Miao, Yuan, and Zhang [35] proved global mild solutions for small data in BMO1BMO^{-1}, and He and Xin [16] constructed self-similar solutions under small homogeneous initial data. Moreover, the existence of forward discretely self-similar and self-similar local Leray solutions is established in the critical space L3,L^{3,\infty} [25] and in the weighted L2L^{2} spaces [12]. The criticality of the L3,L^{3,\infty} class was also highlighted in [34], where global regularity of weak solutions was established under this condition. Additional contributions include the construction of global smooth solutions under spectral constraints [30], the use of Morrey spaces to ensure global well-posedness for small data [31], and the construction of forward self-similar solutions via topological and blow-up methods [42].

Complementing the MHD system, the incompressible viscoelastic Navier–Stokes equations with damping (vNSEd) model non-Newtonian fluids with both viscous and elastic characteristics. In the simplified setting where both relaxation and retardation times are infinite, the vNSEd system reads

tvΔv+vv(𝐅𝐅)+π=0t𝐅+v𝐅(v)𝐅=0v=0} in 3×(0,),\left.\begin{array}[]{ll}\partial_{t}v-\Delta v+v\cdot\nabla v-\nabla\cdot({\bf F}{\bf F}^{\top})+\nabla\pi&=0\\ \partial_{t}{\bf F}+v\cdot\nabla{\bf F}-(\nabla v){\bf F}&=0\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\nabla\cdot v&=0\end{array}\right\}\text{ in }\mathbb{R}^{3}\times(0,\infty), (1.1)

with initial data

v|t=0=v0 and 𝐅|t=0=𝐅0 in 3,v|_{t=0}=v_{0}\ \text{ and }\ {\bf F}|_{t=0}={\bf F}_{0}\ \text{ in }\mathbb{R}^{3},

where vv is the velocity field, 𝐅{\bf F} is the local deformation tensor of the fluid, and π\pi is the pressure. This model arises from Oldroyd-type theories for viscoelastic fluids and captures the interplay between fluid motion and elastic stresses. The addition of a damping term in the equation for 𝐅{\bf F} (following Lin–Liu–Zhang [29]) is critical for obtaining global solutions, particularly in the absence of intrinsic dissipative mechanisms. To be more precise, they introduced the following viscoelastic Navier-Stokes equations with damping as a way to approximate solutions of (1.1):

tvΔv+vv(𝐅𝐅)+π=0t𝐅μΔ𝐅+v𝐅(v)𝐅=0v=0} in 3×(0,),\left.\begin{array}[]{ll}\partial_{t}v-\Delta v+v\cdot\nabla v-\nabla\cdot({\bf F}{\bf F}^{\top})+\nabla\pi&=0\\ \partial_{t}{\bf F}-\mu\Delta{\bf F}+v\cdot\nabla{\bf F}-(\nabla v){\bf F}&=0\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\nabla\cdot v&=0\end{array}\right\}\text{ in }\mathbb{R}^{3}\times(0,\infty), (1.2)

for a damping parameter μ>0\mu>0. Existence results for smooth solutions under smallness conditions or specific symmetries have been established by various authors [10, 26, 29]. Note that if 𝐅=0\nabla\cdot{\bf F}=0 at some instance of time, then 𝐅=0\nabla\cdot{\bf F}=0 at all later times. In fact, by taking divergence of (1.2)2\eqref{vNSEd0}_{2} and using (1.2)3\eqref{vNSEd0}_{3}, one have the following equation for 𝐅\nabla\cdot{\bf F}:

t(𝐅)+v(𝐅)=μΔ(𝐅).\partial_{t}(\nabla\cdot{\bf F})+v\cdot\nabla(\nabla\cdot{\bf F})=\mu\Delta(\nabla\cdot{\bf F}).

Hence it is natural to assume

𝐅=0.\nabla\cdot{\bf F}=0.

The authors [29] noted that using standard weak convergence methods to pass the limit of solutions to (1.2) as μ0+\mu\to 0^{+} does not yield weak solutions of (1.1). Despite this, system (1.2) remains an interesting subject of study. For instance, Lai, Lin, and Wang [24] established the existence of forward self-similar classical solution to (1.2) for locally Hölder continuous, (1)(-1)-homogeneous initial data. Additionally, the existence of forward discretely self-similar and self-similar local Leray solutions in the critical space L3,L^{3,\infty} is established in [25], following the analysis in [2].

Regularity issues for weak solutions of the viscoelastic Navier–Stokes equations with damping have been investigated from several perspectives. Hynd [17] proved a version of the Caffarelli–Kohn–Nirenberg partial regularity theorem adapted to the viscoelastic system with damping, while Kim [22] established Serrin-type regularity criteria in weak-LpL^{p} spaces. These results have been further extended in [39], which proved global existence of mild solutions in scaling-invariant spaces for small data and derived various regularity criteria in Lorentz, multiplier, BMO, and Besov spaces. Additional contributions include the construction of global classical solutions with symmetry assumptions in periodic domains by Liu and Lin [32], and refined local energy bounds leading to improved ϵ\epsilon-regularity conditions in the sense of Caffarelli–Kohn–Nirenberg in [43].

Since the damping parameter μ\mu does not affect our analysis, we set μ=1\mu=1 throughout this paper. Then, columnwisely, (1.2) can be rewritten as

tvΔv+vvn=13fnfn+π=0tfmΔfm+vfmfmv=0fm=v=0} in 3×(0,),m=1,2,3,\left.\begin{array}[]{ll}\partial_{t}v-\Delta v+v\cdot\nabla v-\underset{n=1}{\overset{3}{\sum}}f_{n}\cdot\nabla f_{n}+\nabla\pi&=0\\ \partial_{t}f_{m}-\Delta f_{m}+v\cdot\nabla f_{m}-f_{m}\cdot\nabla v&=0\\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\nabla\cdot f_{m}=\nabla\cdot v&=0\end{array}\right\}\text{ in }\mathbb{R}^{3}\times(0,\infty),\ m=1,2,3, (vNSEd)

where fmf_{m} is the mm-th column vector of 𝐅{\bf F}.

A central theme in the analysis of both the (MHD) and (vNSEd) systems is the interplay between the nonlinear couplings, scaling symmetries, and the functional framework chosen for solutions. While much progress has been made in classical Lebesgue, Sobolev, and Besov spaces, recent advances have highlighted the utility of Wiener amalgam spaces in studying fluid systems. In this paper, the Wiener amalgam spaces are denoted EqpE^{p}_{q} and defined by the norm

fEqp:=fLp(B1(k))q(k3)<.\|f\|_{E^{p}_{q}}:=\bigg{\|}\left\|f\right\|_{L^{p}(B_{1}(k))}\bigg{\|}_{\ell^{q}(k\in{\mathbb{Z}^{3}})}<\infty.

These spaces, which blend local integrability and global decay properties, provide a flexible setting that accommodates non-decaying or large initial data while retaining control over both local and global behaviors. We identify EpE^{p}_{\infty} with LulocpL^{p}_{\mathrm{uloc}} with the norm fLulocp:=supx03fLp(B1(x0))\left\|f\right\|_{L^{p}_{\mathrm{uloc}}}:=\sup_{x_{0}\in\mathbb{R}^{3}}\left\|f\right\|_{L^{p}(B_{1}(x_{0}))}. The closure of Cc(3)C^{\infty}_{c}(\mathbb{R}^{3}) under the LulocpL^{p}_{\mathrm{uloc}} norm is denoted by EpE^{p}. Note that for p,p1,p2,q,q1,q2[1,]p,p_{1},p_{2},q,q_{1},q_{2}\in[1,\infty] we have the Hölder inequality:

fgEqpfEq1p1gEq2p2,1p=1p1+1p2,1q1q1+1q2.\left\|fg\right\|_{E^{p}_{q}}\leq\left\|f\right\|_{E^{p_{1}}_{q_{1}}}\left\|g\right\|_{E^{p_{2}}_{q_{2}}},\quad\frac{1}{p}=\frac{1}{p_{1}}+\frac{1}{p_{2}},\quad\frac{1}{q}\leq\frac{1}{q_{1}}+\frac{1}{q_{2}}. (1.3)

We will consider two kinds of spacetime integrals: For 0<T0<T\leq\infty, x3x\in\mathbb{R}^{3}, and 1s,p,q1\leq s,p,q\leq\infty, define the norms LTsEqpL^{s}_{T}E^{p}_{q} and ET,qs,pE^{s,p}_{T,q} as follows:

fLTsEqp:=fLs(0,T;Eqp(3)),\|f\|_{L^{s}_{T}E^{p}_{q}}:=\|f\|_{L^{s}(0,T;E^{p}_{q}(\mathbb{R}^{3}))}, (1.4)

and

fET,qs,p:=fLTsLp(B1(k))q(k3).\left\|f\right\|_{E^{s,p}_{T,q}}:=\left\|\left\|f\right\|_{L^{s}_{T}L^{p}(B_{1}(k))}\right\|_{\ell^{q}(k\in\mathbb{Z}^{3})}. (1.5)

These norms are different from each other when sqs\neq q. By Minkowski’s integral inequality,

fLTsEqpfET,qs,p, if qs,\left\|f\right\|_{L^{s}_{T}E^{p}_{q}}\leq\left\|f\right\|_{E^{s,p}_{T,q}},\qquad\text{ if }q\leq s, (1.6)

and

fET,qs,pfLTsEqp, if qs.\left\|f\right\|_{E^{s,p}_{T,q}}\leq\left\|f\right\|_{L^{s}_{T}E^{p}_{q}},\qquad\text{ if }q\geq s. (1.7)

Previous works [4, 1] developed a detailed theory for the incompressible Navier–Stokes equations in Wiener amalgam spaces, establishing mild and weak solutions, spacetime integral bounds, and eventual regularity results for different ranges of the Lebesgue exponent qq. In this paper, we extend these techniques to the (MHD) and (vNSEd) systems. Specifically, we prove the existence of mild solutions for small data in critical and subcritical Wiener amalgam spaces, as well as global weak solutions under appropriate integrability and decay conditions. Our analysis demonstrates the robustness of the Wiener amalgam framework in addressing the intricate coupling structures and nonlinearities in these models.

In the following subsections, we introduce the definitions of mild and local energy solutions for the systems (MHD) and (vNSEd), and present our main results.

1.1 Mild solutions of MHD equations

A pair of vector fields (v,b)(v,b) is called a mild solution to (MHD) if it satisfies

(v,b)(x,t)=(etΔv0,etΔb0)B((v,b),(v,b))(t),(v,b)(x,t)=(e^{t\Delta}v_{0},e^{t\Delta}b_{0})-B((v,b),(v,b))(t), (1.8)

where BB is a bilinear operator defined by B=(B1,B2)B=(B_{1},B_{2}),

B1((v,b),(u,a))(t)=0te(ts)Δ(vuba)𝑑s,B2((v,b),(u,a))(t)=0te(ts)Δ(vabu)𝑑s,\begin{split}B_{1}((v,b),(u,a))(t)=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(v\otimes u-b\otimes a)\,ds,\\ B_{2}((v,b),(u,a))(t)=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(v\otimes a-b\otimes u)\,ds,\end{split} (1.9)

in which \mathbb{P} is the Helmholtz projection operator. More precisely, the vector components of the bilinear operators B1B_{1} and B2B_{2} can be expressed by

B1((v,b),(u,a))i(x,t)=0tdlSij(xy,ts)(vlujblaj)(y,s)dyds,i=1,2,3,B2((v,b),(u,a))i(x,t)=0tdlSij(xy,ts)(vlajbluj)(y,s)dyds,i=1,2,3,\begin{split}B_{1}((v,b),(u,a))_{i}(x,t)=\int_{0}^{t}\int_{\mathbb{R}^{d}}\partial_{l}S_{ij}(x-y,t-s)(v_{l}u_{j}-b_{l}a_{j})(y,s)\,dyds,\quad i=1,2,3,\\ B_{2}((v,b),(u,a))_{i}(x,t)=\int_{0}^{t}\int_{\mathbb{R}^{d}}\partial_{l}S_{ij}(x-y,t-s)(v_{l}a_{j}-b_{l}u_{j})(y,s)\,dyds,\quad i=1,2,3,\end{split}

where SijS_{ij} is the Ossen tensor derived by Oseen in [37]. We refer the readers to [18, Section 2.2] for a brief introduction of the Oseen tensor.

We first consider the case of data (v0,b0)Eqr×Eqr(v_{0},b_{0})\in E^{r}_{q}\times E^{r}_{q} with r>3r>3, which we refer to as subcritical, and state the existence of mild solutions in the amalgam spaces in the following theorem.

Theorem 1.1 (Subcritical data (MHD)).

Let r(3,]r\in(3,\infty] and q[1,]q\in[1,\infty]. If v0,b0Eqrv_{0},b_{0}\in E^{r}_{q} are divergence free, then, for any positive time T=T((v0,b0)Eqr×Eqr)T=T(\|(v_{0},b_{0})\|_{E^{r}_{q}\times E^{r}_{q}}) chosen so that

T1/23/(2r)+T1/2(v0,b0)Eqr×Eqr1,T^{1/2-3/(2r)}+T^{1/2}{\ \lesssim\ }\|(v_{0},b_{0})\|_{E^{r}_{q}\times E^{r}_{q}}^{-1}, (1.10)

there exists a unique mild solution (v,b)L(0,T;Eqr×Eqr)C((0,T);Eqr×Eqr)(v,b)\in L^{\infty}(0,T;E^{r}_{q}\times E^{r}_{q})\cap C((0,T);E^{r}_{q}\times E^{r}_{q}) to (MHD). Moreover, (v,b)(v,b) satisfies

sup0tT(v,b)(t)Eqr×EqrC(v0,b0)Eqr×Eqr.\sup_{0\leq t\leq T}\|(v,b)(t)\|_{E^{r}_{q}\times E^{r}_{q}}\leq C\|(v_{0},b_{0})\|_{E^{r}_{q}\times E^{r}_{q}}. (1.11)

If q,r<q,r<\infty, then (v,b)C([0,T];Eqr×Eqr)(v,b)\in C([0,T];E^{r}_{q}\times E^{r}_{q}). If q=q=\infty or r=r=\infty, then we still have (etΔv0v(t),etΔb0b(t))Eqr×Eqr0\|(e^{t\Delta}v_{0}-{v(t)},e^{t\Delta}b_{0}-{b(t)})\|_{E^{r}_{q}\times E^{r}_{q}}\to 0 as t0+t\to 0^{+}.

Furthermore, if r<r<\infty, then for any s[r,]s\in[r,\infty] and p[r,3r]p\in[r,3r] with 2s+3p=3r\frac{2}{s}+\frac{3}{p}=\frac{3}{r},

(v,b)ET,ms,p×ET,ms,pC(v0,b0)Eqr×Eqr,mq,\left\|(v,b)\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}\leq C\|(v_{0},b_{0})\|_{E^{r}_{q}\times E^{r}_{q}},\qquad m\geq q,

provided (1+T1s+ϵ)(T1232r+T11s)(v0,b0)Eqr×Eqr1(1+T^{\frac{1}{s}+\epsilon})(T^{\frac{1}{2}-\frac{3}{2r}}+T^{1-\frac{1}{s}}){\ \lesssim\ }\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}^{-1} for all ϵ>0\epsilon>0.

Theorem 1.1 is proved in Section 2.1.

We now turn to the critical case, i.e., the case when data (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q}. When the data is sufficiently small, we have the following existence theorem of mild solutions.

Theorem 1.2 (Critical data I (MHD)).

Let q[1,]q\in[1,\infty]. Fix T>0T>0. There exists ε=ε(T)>0\varepsilon=\varepsilon(T)>0 such that for all divergence-free v0,b0Eq3v_{0},b_{0}\in E^{3}_{q} with (v0,b0)Eq3×Eq3ε\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}\leq\varepsilon, there exists a mild solution (v,b)(v,b) to (MHD) with

(v,b)L(0,T;Eq3×Eq3)andt12(v,b)L(0,T;Eq×Eq).(v,b)\in L^{\infty}(0,T;E^{3}_{q}\times E^{3}_{q})\quad\text{and}\quad t^{\frac{1}{2}}(v,b)\in L^{\infty}(0,T;E^{\infty}_{q}\times E^{\infty}_{q}).

The solution is unique in the class

sup0<t<Tt14(v,b)Eq6×Eq62sup0<t<Tt14(etΔv0,etΔb0)Eq6×Eq6.\sup_{0<t<T}t^{\frac{1}{4}}\left\|(v,b)\right\|_{E^{6}_{q}\times E^{6}_{q}}\leq 2\sup_{0<t<T}t^{\frac{1}{4}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{6}_{q}\times E^{6}_{q}}. (1.12)

Furthermore, (v,b)LT(Eq3×Eq3)+t1/2(v,b)LT(Eq×Eq)(v0,b0)Eq3×Eq3\left\|(v,b)\right\|_{L^{\infty}_{T}(E^{3}_{q}\times E^{3}_{q})}+\left\|t^{1/2}(v,b)\right\|_{L^{\infty}_{T}(E^{\infty}_{q}\times E^{\infty}_{q})}{\ \lesssim\ }\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}. We have (v,b)C((0,T);Eq3×Eq3)(v,b)\in C((0,T);E^{3}_{q}\times E^{3}_{q}) for q=q=\infty and (v,b)C([0,T);Eq3×Eq3)(v,b)\in C([0,T);E^{3}_{q}\times E^{3}_{q}) for q<q<\infty. If q=q=\infty, then we have for any ball BB and δ(0,2]\delta\in(0,2] that

limt0+(v,b)(t)(v0,b0)L3δ(B)×L3δ(B)=0.\lim_{t\to 0^{+}}\|(v,b)(t)-(v_{0},b_{0})\|_{L^{3-\delta}(B)\times L^{3-\delta}(B)}=0. (1.13)

For any s[3,]s\in[3,\infty] and p[3,9]p\in[3,9] given by 2s+3p=1\frac{2}{s}+\frac{3}{p}=1, by taking εε0(T,s)\varepsilon\leq\varepsilon_{0}(T,s) sufficiently small, this solution further satisfies

(v,b)ET,ms,p×ET,ms,p+𝟙qs(v,b)LTs(Emp×Emp)C(v0,b0)Eq3×Eq3,m[q,].\left\|(v,b)\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}+\mathbbm{1}_{q\leq s}\left\|(v,b)\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}\leq C\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}},\quad\forall m\in[q,\infty]. (1.14)

The following theorem concerns the critical case with enough decay, 1q31\leq q\leq 3.

Theorem 1.3 (Critical data II (MHD)).

Let 1q31\leq q\leq 3. For all divergence-free v0,b0Eq3v_{0},b_{0}\in E^{3}_{q}, there exist T=T(v0,b0)>0T=T(v_{0},b_{0})>0 and a unique mild solution (v,b)(v,b) to (MHD) satisfying

(v,b)BC([0,T);Eq3×Eq3)andt12(v,b)L(0,T;Eq2×Eq2),(v,b)\in BC([0,T);E^{3}_{q}\times E^{3}_{q})\quad\text{and}\quad t^{\frac{1}{2}}(v,b)\in L^{\infty}(0,T;E^{\infty}_{q_{2}}\times E^{\infty}_{q_{2}}),

with 1/q2=1/q1/31/q_{2}=1/q-1/3, q2[32,]q_{2}\in[\frac{3}{2},\infty]. For any s[3,)s\in[3,\infty), 2s+3p=1\frac{2}{s}+\frac{3}{p}=1, and m[q,]m\in[q,\infty], there is T1(0,T]T_{1}\in(0,T] such that

(v,b)ET1,ms,p×ET1,ms,p.(v,b)\in E^{s,p}_{T_{1},m}\times E^{s,p}_{T_{1},m}. (1.15)

Furthermore, there is ε(q)>0\varepsilon(q)>0 such that T=T=\infty if (v0,b0)Eq3×Eq3ε(q)\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}\leq\varepsilon(q). If we assume further

m>p=pp1,andmm1,2s+3m1=3q,m>p^{\prime}=\frac{p}{p-1},\quad\text{and}\quad m\geq m_{1},\quad\frac{2}{s}+\frac{3}{m_{1}}=\frac{3}{q}, (1.16)

with m>m1(s,q)m>m_{1}(s,q) when q=1q=1, then there exists ε1(s,q,m)>0\varepsilon_{1}(s,q,m)>0 such that T1=T_{1}=\infty if (v0,b0)Eq3×Eq3ε1(s,q,m)\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}\leq\varepsilon_{1}(s,q,m). Instead of (1.25), if we assume

mmax(p,m1),and{m>m1if q=1,mpif 3s<5q,m\geq\max(p^{\prime},m_{1}),\quad\text{and}\quad\left\{\begin{aligned} m>m_{1}&\quad\text{if }\quad q=1,\\ m\geq p&\quad\text{if }\quad 3s<5q,\end{aligned}\right. (1.17)

then there exists ε2(s,q,m)>0\varepsilon_{2}(s,q,m)>0 such that v,bLT=sEmpv,b\in L^{s}_{T=\infty}E^{p}_{m} if (v0,b0)Eq3×Eq3ε2(s,q,m)\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}\leq\varepsilon_{2}(s,q,m).

We prove Theorem 1.3 in Section 2.3.

1.2 Mild solutions of viscoelastic Navier–Stokes equations with damping

A pair (v,𝐅)(v,{\bf F}), 𝐅=[f1,f2,f3]3×3{\bf F}=[f_{1},f_{2},f_{3}]\in\mathbb{R}^{3\times 3}, is called a mild solution to (vNSEd) if it satisfies

v(x,t)=etΔv00((v,𝐅),(v,𝐅))(t),fm(x,t)=etΔ(fm)0m((v,𝐅),(v,𝐅))(t),m=1,2,3,\begin{split}v(x,t)&=e^{t\Delta}v_{0}-\mathcal{B}_{0}((v,{\bf F}),(v,{\bf F}))(t),\\ f_{m}(x,t)&=e^{t\Delta}(f_{m})_{0}-\mathcal{B}_{m}((v,{\bf F}),(v,{\bf F}))(t),\quad m=1,2,3,\end{split} (1.18)

where

0((v,𝐅),(u,𝐆))(t)=0te(ts)Δ(vun=13fngn)𝑑s,𝐆=[g1,g2,g3],m((v,𝐅),(u,𝐆))(t)=0te(ts)Δ(vgmfmu)𝑑s.\begin{split}\mathcal{B}_{0}((v,{\bf F}),(u,{\bf G}))(t)&=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left(v\otimes u-\sum_{n=1}^{3}f_{n}\otimes g_{n}\right)\,ds,\quad{\bf G}=[g_{1},g_{2},g_{3}],\\ \mathcal{B}_{m}((v,{\bf F}),(u,{\bf G}))(t)&=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(v\otimes g_{m}-f_{m}\otimes u)\,ds.\end{split}

The main results of the mild solutions for the viscoelastic Navier–Stokes equations with damping are stated as follows:

Theorem 1.4 (Subcritical data (vNSEd)).

Let r(3,]r\in(3,\infty] and q[1,]q\in[1,\infty]. If (v0,𝐅0)Eqr×𝐄qr(v_{0},{\bf F}_{0})\in E^{r}_{q}\times{\bf E}^{r}_{q}, 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], where v0v_{0} and (fm)0(f_{m})_{0}, m=1,2,3m=1,2,3, are divergence free, then, for any positive time T=T((v0,𝐅0)Eqr×𝐄qr)T=T(\|(v_{0},{\bf F}_{0})\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}) chosen so that

T1/23/(2r)+T1/2(v0,𝐅0)Eqr×𝐄qr1,T^{1/2-3/(2r)}+T^{1/2}{\ \lesssim\ }\|(v_{0},{\bf F}_{0})\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}^{-1}, (1.19)

there exists a unique mild solution (v,𝐅)L(0,T;Eqr×𝐄qr)C((0,T);Eqr×𝐄qr)(v,{\bf F})\in L^{\infty}(0,T;E^{r}_{q}\times{\bf E}^{r}_{q})\cap C((0,T);E^{r}_{q}\times{\bf E}^{r}_{q}) to (vNSEd). Moreover, (v,𝐅)(v,{\bf F}) satisfies

sup0tT(v,𝐅)(t)Eqr×𝐄qrC(v0,𝐅0)Eqr×𝐄qr.\sup_{0\leq t\leq T}\|(v,{\bf F})(t)\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}\leq C\|(v_{0},{\bf F}_{0})\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}. (1.20)

If q,r<q,r<\infty, then (v,𝐅)C([0,T];Eqr×𝐄qr)(v,{\bf F})\in C([0,T];E^{r}_{q}\times{\bf E}^{r}_{q}). If q=q=\infty or r=r=\infty, then we still have (etΔv0v(t),etΔ𝐅0𝐅(t))Eqr×𝐄qr0\|(e^{t\Delta}v_{0}-{v(t)},e^{t\Delta}{\bf F}_{0}-{{\bf F}(t)})\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}\to 0 as t0+t\to 0^{+}.

Furthermore, if r<r<\infty, then for any s[r,]s\in[r,\infty] and p[r,3r]p\in[r,3r] with 2s+3p=3r\frac{2}{s}+\frac{3}{p}=\frac{3}{r},

(v,𝐅)ET,ms,p×𝐄T,ms,pC(v0,𝐅0)Eqr×𝐄qr,mq,\left\|(v,{\bf F})\right\|_{E^{s,p}_{T,m}\times{\bf E}^{s,p}_{T,m}}\leq C\|(v_{0},{\bf F}_{0})\|_{E^{r}_{q}\times{\bf E}^{r}_{q}},\qquad m\geq q,

provided (1+T1s+ϵ)(T1232r+T11s)(v0,𝐅0)Eqr×𝐄qr1(1+T^{\frac{1}{s}+\epsilon})(T^{\frac{1}{2}-\frac{3}{2r}}+T^{1-\frac{1}{s}}){\ \lesssim\ }\left\|(v_{0},{\bf F}_{0})\right\|_{E^{r}_{q}\times{\bf E}^{r}_{q}}^{-1} for all ϵ>0\epsilon>0.

Theorem 1.5 (Critical data I (vNSEd)).

Let q[1,]q\in[1,\infty]. Fix T>0T>0. There exists ε=ε(T)>0\varepsilon=\varepsilon(T)>0 such that for all divergence-free v0,(fm)0Eq3v_{0},(f_{m})_{0}\in E^{3}_{q}, m=1,2,3m=1,2,3, with (v0,𝐅0)Eq3×𝐄q3ε\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}}\leq\varepsilon, 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], there exists a mild solution (v,𝐅)(v,{\bf F}) to (vNSEd) with

(v,𝐅)L(0,T;Eq3×𝐄q3)andt12(v,𝐅)L(0,T;Eq×𝐄q).(v,{\bf F})\in L^{\infty}(0,T;E^{3}_{q}\times{\bf E}^{3}_{q})\quad\text{and}\quad t^{\frac{1}{2}}(v,{\bf F})\in L^{\infty}(0,T;E^{\infty}_{q}\times{\bf E}^{\infty}_{q}).

The solution is unique in the class

sup0<t<Tt14(v,𝐅)Eq6×𝐄q62sup0<t<Tt14(etΔv0,etΔ𝐅0)Eq6×𝐄q6.\sup_{0<t<T}t^{\frac{1}{4}}\left\|(v,{\bf F})\right\|_{E^{6}_{q}\times{\bf E}^{6}_{q}}\leq 2\sup_{0<t<T}t^{\frac{1}{4}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}{\bf F}_{0})\right\|_{E^{6}_{q}\times{\bf E}^{6}_{q}}. (1.21)

Furthermore, (v,𝐅)LT(Eq3×𝐄q3)+t1/2(v,𝐅)LT(Eq×𝐄q)(v0,𝐅0)Eq3×𝐄q3\left\|(v,{\bf F})\right\|_{L^{\infty}_{T}(E^{3}_{q}\times{\bf E}^{3}_{q})}+\left\|t^{1/2}(v,{\bf F})\right\|_{L^{\infty}_{T}(E^{\infty}_{q}\times{\bf E}^{\infty}_{q})}{\ \lesssim\ }\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}}. We have (v,𝐅)C((0,T);Eq3×𝐄q3)(v,{\bf F})\in C((0,T);E^{3}_{q}\times{\bf E}^{3}_{q}) for q=q=\infty and (v,𝐅)C([0,T);Eq3×𝐄q3)(v,{\bf F})\in C([0,T);E^{3}_{q}\times{\bf E}^{3}_{q}) for q<q<\infty. If q=q=\infty, then we have for any ball BB and δ(0,2]\delta\in(0,2] that

limt0+(v,𝐅)(t)(v0,𝐅0)L3δ(B)×𝐋3δ(B)=0.\lim_{t\to 0^{+}}\|(v,{\bf F})(t)-(v_{0},{\bf F}_{0})\|_{L^{3-\delta}(B)\times{\bf L}^{3-\delta}(B)}=0. (1.22)

For any s[3,]s\in[3,\infty] and p[3,9]p\in[3,9] given by 2s+3p=1\frac{2}{s}+\frac{3}{p}=1, by taking εε0(T,s)\varepsilon\leq\varepsilon_{0}(T,s) sufficiently small, this solution further satisfies

(v,𝐅)ET,ms,p×𝐄T,ms,p+𝟙qs(v,𝐅)LTs(Emp×𝐄mp)C(v0,𝐅0)Eq3×𝐄q3,m[q,].\left\|(v,{\bf F})\right\|_{E^{s,p}_{T,m}\times{\bf E}^{s,p}_{T,m}}+\mathbbm{1}_{q\leq s}\left\|(v,{\bf F})\right\|_{L^{s}_{T}(E^{p}_{m}\times{\bf E}^{p}_{m})}\leq C\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}},\quad\forall m\in[q,\infty]. (1.23)
Theorem 1.6 (Critical data II (vNSEd)).

Let 1q31\leq q\leq 3. For all divergence-free v0,(f1)0,(f2)0,(f3)0Eq3v_{0},(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}\in E^{3}_{q}, there exist T=T(v0,𝐅0)>0T=T(v_{0},{\bf F}_{0})>0, 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], and a unique mild solution (v,𝐅)(v,{\bf F}) to (vNSEd) satisfying

(v,𝐅)BC([0,T);Eq3×𝐄q3)andt12(v,𝐅)L(0,T;Eq2×𝐄q2),(v,{\bf F})\in BC([0,T);E^{3}_{q}\times{\bf E}^{3}_{q})\quad\text{and}\quad t^{\frac{1}{2}}(v,{\bf F})\in L^{\infty}(0,T;E^{\infty}_{q_{2}}\times{\bf E}^{\infty}_{q_{2}}),

with 1/q2=1/q1/31/q_{2}=1/q-1/3, q2[32,]q_{2}\in[\frac{3}{2},\infty]. For any s[3,)s\in[3,\infty), 2s+3p=1\frac{2}{s}+\frac{3}{p}=1, and m[q,]m\in[q,\infty], there is T1(0,T]T_{1}\in(0,T] such that

(v,𝐅)ET1,ms,p×𝐄T1,ms,p.(v,{\bf F})\in E^{s,p}_{T_{1},m}\times{\bf E}^{s,p}_{T_{1},m}. (1.24)

Furthermore, there is ε(q)>0\varepsilon(q)>0 such that T=T=\infty if (v0,𝐅0)Eq3×𝐄q3ε(q)\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}}\leq\varepsilon(q). If we assume further

m>p=pp1,andmm1,2s+3m1=3q,m>p^{\prime}=\frac{p}{p-1},\quad\text{and}\quad m\geq m_{1},\quad\frac{2}{s}+\frac{3}{m_{1}}=\frac{3}{q}, (1.25)

with m>m1(s,q)m>m_{1}(s,q) when q=1q=1, then there exists ε1(s,q,m)>0\varepsilon_{1}(s,q,m)>0 such that T1=T_{1}=\infty if (v0,𝐅0)Eq3×𝐄q3ε1(s,q,m)\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}}\leq\varepsilon_{1}(s,q,m). Instead of (1.25), if we assume

mmax(p,m1),and{m>m1if q=1,mpif 3s<5q,m\geq\max(p^{\prime},m_{1}),\quad\text{and}\quad\left\{\begin{aligned} m>m_{1}&\quad\text{if }\quad q=1,\\ m\geq p&\quad\text{if }\quad 3s<5q,\end{aligned}\right. (1.26)

then there exists ε2(s,q,m)>0\varepsilon_{2}(s,q,m)>0 such that (v,𝐅)LT=s(Emp×𝐄mp)(v,{\bf F})\in L^{s}_{T=\infty}(E^{p}_{m}\times{\bf E}^{p}_{m}) if (v0,𝐅0)Eq3×𝐄q3ε2(s,q,m)\left\|(v_{0},{\bf F}_{0})\right\|_{E^{3}_{q}\times{\bf E}^{3}_{q}}\leq\varepsilon_{2}(s,q,m).

1.3 Weak solutions of MHD equations

We first introduce the notion of local energy solutions for the MHD equations, which is consistent with the concept introduced in [4] for the Navier–Stokes equations.

Definition 1.7 (local energy solution (MHD)).

Let 0<T0<T\leq\infty. A pair of vector fields (v,b)(v,b), v,bLloc2(3×[0,T))v,b\in L^{2}_{\mathrm{loc}}(\mathbb{R}^{3}\times[0,T)), is a local energy solution to (MHD) with divergence-free initial data v0,b0Luloc2(3)v_{0},b_{0}\in L^{2}_{\mathrm{uloc}}(\mathbb{R}^{3}), denoted as (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}), if the following hold:

1. There exists πLloc3/2(3×[0,T))\pi\in L^{3/2}_{\mathrm{loc}}(\mathbb{R}^{3}\times[0,T)) such that (v,b,π)(v,b,\pi) is a distributional solution to (MHD).

2. For any R>0R>0, (v,b)(v,b) satisfies

esssup0t<R2Tsupx03BR(x0)12(|v(x,t)|2+|b(x,t)|2)𝑑x+supx030R2TBR(x0)(|v(x,t)|2+|b(x,t)|2)𝑑x𝑑t<.\begin{split}\mathop{\rm ess\,sup}_{0\leq t<R^{2}\wedge T}&\sup_{x_{0}\in\mathbb{R}^{3}}\int_{B_{R}(x_{0})}\frac{1}{2}\left(|v(x,t)|^{2}+|b(x,t)|^{2}\right)dx\\ &+\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{R^{2}\wedge T}\int_{B_{R}(x_{0})}\left(|\nabla v(x,t)|^{2}+|\nabla b(x,t)|^{2}\right)dxdt<\infty.\end{split}

3. For any R>0R>0, x03x_{0}\in\mathbb{R}^{3}, and 0<T<T0<T^{\prime}<T, there exists a function of time cx0,R(t)L3/2(0,T)c_{x_{0},R}(t)\in L^{3/2}(0,T^{\prime}) so that, for every 0<t<T0<t<T^{\prime} and xB2R(x0)x\in B_{2R}(x_{0}),

π(x,t)=Δ1divdiv[(vvbb)χ4R(xx0)]3(K(xy)K(x0y))(vvbb)(y,t)(1χ4R(yx0))𝑑y+cx0,R(t)\begin{split}\pi(x,t)&=-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[(v\otimes v-b\otimes b)\chi_{4R}(x-x_{0})\right]\\ &\quad-\int_{\mathbb{R}^{3}}\left(K(x-y)-K(x_{0}-y)\right)(v\otimes v-b\otimes b)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy+c_{x_{0},R}(t)\end{split} (1.27)

in L3/2(B2R(x0)×(0,T))L^{3/2}(B_{2R}(x_{0})\times(0,T^{\prime})) where K(x)K(x) is the kernel of Δ1divdiv\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits, Kij(x)=ij14π|x|K_{ij}(x)=\partial_{i}\partial_{j}\frac{-1}{4\pi|x|}, and χ4R(x)\chi_{4R}(x) is the characteristic function of B4RB_{4R}.

4. For all compact subsets KK of 3\mathbb{R}^{3} we have v(t)v0v(t)\to v_{0} and b(t)b0b(t)\to b_{0} in L2(K)L^{2}(K) as t0+t\to 0^{+}.

5. For all cylinders QQ compactly supported in 3×(0,T)\mathbb{R}^{3}\times(0,T) and all nonnegative ϕCc(Q)\phi\in C^{\infty}_{c}(Q), we have the local energy inequality

2(|v|2+|b|2)ϕ𝑑x𝑑t(|v|2+|b|2)(tϕ+Δϕ)𝑑x𝑑t+(|v|2+|b|2+2π)(vϕ)𝑑x𝑑t2(vb)(bϕ)𝑑x𝑑t.\begin{split}2\int\int\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi\,dxdt\leq&\int\int\left(|v|^{2}+|b|^{2}\right)\left(\partial_{t}\phi+\Delta\phi\right)dxdt\\ &+\int\int\left(|v|^{2}+|b|^{2}+2\pi\right)(v\cdot\nabla\phi)\,dxdt\\ &-2\int\int(v\cdot b)(b\cdot\nabla\phi)\,dxdt.\end{split} (1.28)

6. The functions

t3v(x,t)w(x)𝑑xt3b(x,t)w(x)𝑑xt\mapsto\int_{\mathbb{R}^{3}}v(x,t)\cdot w(x)\,dx\qquad t\mapsto\int_{\mathbb{R}^{3}}b(x,t)\cdot w(x)\,dx

are continuous in t[0,T)t\in[0,T) for any compactly supported wL2(3)w\in L^{2}(\mathbb{R}^{3}).

For given divergence-free v0,b0Luloc2v_{0},b_{0}\in L^{2}_{\mathrm{uloc}}, let 𝒩MHD(v0,b0)\mathcal{N}_{\rm MHD}(v_{0},b_{0}) denote the set of all local energy solutions to (MHD) with initial data (v0,b0)(v_{0},b_{0}).

Theorem 1.8 (Eventual regularity in Eq2E^{2}_{q} (MHD)).

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q}, where 1q31\leq q\leq 3, are divergence free, and (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}). Then (v,b)(v,b) has eventual regularity, i.e., there is t1<t_{1}<\infty such that vv and bb are regular at (x,t)(x,t) whenever tt1t\geq t_{1}, and

(v,b)(,t)L×Lt1/2,\left\|(v,b)(\cdot,t)\right\|_{L^{\infty}\times L^{\infty}}{\ \lesssim\ }t^{1/2},

for sufficiently large tt.

The proof of Theorem 1.8 is given in Section 3.1

Define the q\ell^{q} local energy

(v,b)𝐋𝐄q(0,T):=(v,b)ET,q,2×ET,q,2+(v,b)ET,q2,2×ET,q2,2.\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}:=\left\|(v,b)\right\|_{E^{\infty,2}_{T,q}\times E^{\infty,2}_{T,q}}+\left\|(\nabla v,\nabla b)\right\|_{E^{2,2}_{T,q}\times E^{2,2}_{T,q}}. (1.29)
Theorem 1.9 (Explicit growth rate in Eq2E^{2}_{q} (MHD)).

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q}, where 1q<1\leq q<\infty, are divergence free, and (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}) satisfies, for some T2>0T_{2}>0,

(v,b)𝐋𝐄q(0,T1)<,T1(0,T2).\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T_{1})}<\infty,\quad\forall T_{1}\in(0,T_{2}).

Then, for any R1R\geq 1, with T=min(λ1(1+(v0,b0)Eq2×Eq2)4R2,T2)T=\min\left(\lambda_{1}(1+\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}})^{-4}R^{2},T_{2}\right), we have

esssup0tTBR(Rk)(|v|2+|b|2)𝑑x+0TBR(Rk)(|v|2+|b|2)𝑑x𝑑tq/2(k3)C(v0,b0)Eq2×Eq22,\bigg{\|}\mathop{\rm ess\,sup}_{0\leq t\leq T}\int_{B_{R}(Rk)}\left(|v|^{2}+|b|^{2}\right)dx+\int_{0}^{T}\int_{B_{R}(Rk)}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dx\,dt\bigg{\|}_{\ell^{q/2}(k\in\mathbb{Z}^{3})}\leq C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}^{2}, (1.30)

for positive constants λ1\lambda_{1} and CC independent of (v0,b0)(v_{0},b_{0}) and RR. In particular, if T2=T_{2}=\infty then TT\to\infty as RR\to\infty.

The proof of Theorem 1.9 is given in Section 3.2

Theorem 1.10 (Existence in Eq2E^{2}_{q} (MHD)).

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q}, where 1q<1\leq q<\infty, are divergence free. Then, there exists a time-global local energy solution (v,b)(v,b) and associated pressure π\pi to (MHD) in 3\mathbb{R}^{3} with initial data v0,b0v_{0},b_{0} so that, for any 0<T<0<T<\infty,

(v,b)𝐋𝐄q(0,T)<.\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}<\infty. (1.31)

In particular, (v,b)L(0,T;Eq2×Eq2)(v,b)\in L^{\infty}(0,T;E^{2}_{q}\times E^{2}_{q}).

The proof of Theorem 1.10 is divided into two cases: q2q\geq 2 and 1q<21\leq q<2. The case q2q\geq 2 is addressed in Section 3.3.1, and the case 1q<21\leq q<2 is handled separately in Section 3.3.2.

1.4 Weak solutions of viscoelastic Navier–Stokes equations with damping

We now define analogous local energy solutions to the viscoelastic Navier–Stokes equations with damping as follows.

Definition 1.11 (local energy solution (vNSEd)).

Let 0<T0<T\leq\infty. A pair of a vector field and a tensor (v,𝐅)(v,{\bf F}), 𝐅=[f1,f2,f3]{\bf F}=[f_{1},f_{2},f_{3}], v,fmLloc2(3×[0,T))v,f_{m}\in L^{2}_{\mathrm{loc}}(\mathbb{R}^{3}\times[0,T)), m=1,2,3m=1,2,3, is a local energy solution to (vNSEd) with initial data v0,𝐅0Luloc2(3)v_{0},{\bf F}_{0}\in L^{2}_{\mathrm{uloc}}(\mathbb{R}^{3}), 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], where v0,(fm)0v_{0},(f_{m})_{0} are divergence free, denoted as (v,𝐅)𝒩vNSEd(v0,𝐅0)(v,{\bf F})\in\mathcal{N}_{\rm vNSEd}(v_{0},{\bf F}_{0}), if the following hold:

1. There exists πLloc3/2(3×[0,T))\pi\in L^{3/2}_{\mathrm{loc}}(\mathbb{R}^{3}\times[0,T)) such that (v,𝐅,π)(v,{\bf F},\pi) is a distributional solution to (vNSEd).

2. For any R>0R>0, (v,𝐅)(v,{\bf F}) satisfies

esssup0t<R2Tsupx03BR(x0)12(|v(x,t)|2+|𝐅(x,t)|2)𝑑x+supx030R2TBR(x0)(|v(x,t)|2+|𝐅(x,t)|2)𝑑x𝑑t<.\begin{split}\mathop{\rm ess\,sup}_{0\leq t<R^{2}\wedge T}&\sup_{x_{0}\in\mathbb{R}^{3}}\int_{B_{R}(x_{0})}\frac{1}{2}\left(|v(x,t)|^{2}+|{\bf F}(x,t)|^{2}\right)dx\\ &+\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{R^{2}\wedge T}\int_{B_{R}(x_{0})}\left(|\nabla v(x,t)|^{2}+|\nabla{\bf F}(x,t)|^{2}\right)dxdt<\infty.\end{split}

3. For any R>0R>0, x03x_{0}\in\mathbb{R}^{3}, and 0<T<T0<T^{\prime}<T, there exists a function of time cx0,R(t)L3/2(0,T)c_{x_{0},R}(t)\in L^{3/2}(0,T^{\prime}) so that, for every 0<t<T0<t<T^{\prime} and xB2R(x0)x\in B_{2R}(x_{0}),

π(x,t)=Δ1divdiv[(vvn=13fnfn)χ4R(xx0)]3(K(xy)K(x0y))(vvn=13fnfn)(y,t)(1χ4R(yx0))𝑑y+cx0,R(t)\begin{split}\pi(x,t)&=-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[\left(v\otimes v-\sum_{n=1}^{3}f_{n}\otimes f_{n}\right)\chi_{4R}(x-x_{0})\right]\\ &\quad-\int_{\mathbb{R}^{3}}\left(K(x-y)-K(x_{0}-y)\right)\left(v\otimes v-\sum_{n=1}^{3}f_{n}\otimes f_{n}\right)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy+c_{x_{0},R}(t)\end{split} (1.32)

in L3/2(B2R(x0)×(0,T))L^{3/2}(B_{2R}(x_{0})\times(0,T^{\prime})) where K(x)K(x) is the kernel of Δ1divdiv\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits, Kij(x)=ij14π|x|K_{ij}(x)=\partial_{i}\partial_{j}\frac{-1}{4\pi|x|}, and χ4R(x)\chi_{4R}(x) is the characteristic function of B4RB_{4R}.

4. For all compact subsets KK of 3\mathbb{R}^{3} we have v(t)v0v(t)\to v_{0} and fm(t)(fm)0f_{m}(t)\to(f_{m})_{0}, m=1,2,3m=1,2,3, in L2(K)L^{2}(K) as t0+t\to 0^{+}.

5. For all cylinders QQ compactly supported in 3×(0,T)\mathbb{R}^{3}\times(0,T) and all nonnegative ϕCc(Q)\phi\in C^{\infty}_{c}(Q), we have the local energy inequality

2(|v|2+|𝐅|2)ϕ𝑑x𝑑t(|v|2+|𝐅|2)(tϕ+Δϕ)𝑑x𝑑t+(|v|2+|𝐅|2+2π)(vϕ)𝑑x𝑑t2n=13(vfn)(fnϕ)𝑑x𝑑t.\begin{split}2\int\int\left(|\nabla v|^{2}+|\nabla{\bf F}|^{2}\right)\phi\,dxdt\leq&\int\int\left(|v|^{2}+|{\bf F}|^{2}\right)\left(\partial_{t}\phi+\Delta\phi\right)dxdt\\ &+\int\int\left(|v|^{2}+|{\bf F}|^{2}+2\pi\right)(v\cdot\nabla\phi)\,dxdt\\ &-2\sum_{n=1}^{3}\int\int(v\cdot f_{n})(f_{n}\cdot\nabla\phi)\,dxdt.\end{split} (1.33)

6. The functions

t3v(x,t)w(x)𝑑xt3𝐅(x,t)w(x)𝑑xt\mapsto\int_{\mathbb{R}^{3}}v(x,t)\cdot w(x)\,dx\qquad t\mapsto\int_{\mathbb{R}^{3}}{\bf F}(x,t)\cdot w(x)\,dx

are continuous in t[0,T)t\in[0,T) for any compactly supported wL2(3)w\in L^{2}(\mathbb{R}^{3}).

For given divergence-free v0,(fm)0Luloc2v_{0},(f_{m})_{0}\in L^{2}_{\mathrm{uloc}}, m=1,2,3m=1,2,3, let 𝒩MHD(v0,𝐅0)\mathcal{N}_{\rm MHD}(v_{0},{\bf F}_{0}), 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], denote the set of all local energy solutions to (vNSEd) with initial data (v0,𝐅0)(v_{0},{\bf F}_{0}).

The main results concerning weak solutions of the viscoelastic Navier–Stokes equations with damping are stated as follows:

Theorem 1.12 (Eventual regularity in Eq2E^{2}_{q} (vNSEd)).

Assume v0,(f1)0,(f2)0,(f3)0Eq2v_{0},(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}\in E^{2}_{q}, where 1q31\leq q\leq 3, are divergence free, and (v,𝐅)𝒩vNSEd(v0,𝐅0)(v,{\bf F})\in\mathcal{N}_{\rm vNSEd}(v_{0},{\bf F}_{0}), 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}]. Then (v,𝐅)(v,{\bf F}) has eventual regularity, i.e., there is t1<t_{1}<\infty such that vv and 𝐅{\bf F} are regular at (x,t)(x,t) whenever tt1t\geq t_{1}, and

(v,𝐅)(,t)L×𝐋t1/2,\left\|(v,{\bf F})(\cdot,t)\right\|_{L^{\infty}\times{\bf L}^{\infty}}{\ \lesssim\ }t^{1/2},

for sufficiently large tt.

Define the q\ell^{q} local energy

(v,𝐅)𝐋𝐄q(0,T):=(v,𝐅)ET,q,2×𝐄T,q,2+(v,𝐅)ET,q2,2×𝐄T,q2,2.\begin{split}\left\|(v,{\bf F})\right\|_{{\bf LE}_{q}(0,T)}:=\left\|(v,{\bf F})\right\|_{E^{\infty,2}_{T,q}\times{\bf E}^{\infty,2}_{T,q}}+\left\|(\nabla v,\nabla{\bf F})\right\|_{E^{2,2}_{T,q}\times{\bf E}^{2,2}_{T,q}}.\end{split}
Theorem 1.13 (Explicit growth rate in Eq2E^{2}_{q} (vNSEd)).

Assume v0,(f1)0,(f2)0,(f3)0Eq2v_{0},(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}\in E^{2}_{q}, where 1q<1\leq q<\infty, are divergence free, and (v,𝐅)𝒩vNSEd(v0,𝐅0)(v,{\bf F})\in\mathcal{N}_{\rm vNSEd}(v_{0},{\bf F}_{0}), 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], satisfies, for some T2>0T_{2}>0,

(v,𝐅)𝐋𝐄q(0,T1)<,T1(0,T2).\left\|(v,{\bf F})\right\|_{{\bf LE}_{q}(0,T_{1})}<\infty,\quad\forall T_{1}\in(0,T_{2}).

Then, for any R1R\geq 1, with T=min(λ1(1+(v0,𝐅0)Eq2×𝐄q2)4R2,T2)T=\min\left(\lambda_{1}(1+\left\|(v_{0},{\bf F}_{0})\right\|_{E^{2}_{q}\times{\bf E}^{2}_{q}})^{-4}R^{2},T_{2}\right), we have

esssup0tTBR(Rk)(|v|2+|𝐅|2)𝑑x+0TBR(Rk)(|v|2+|𝐅|2)𝑑x𝑑tq/2(k3)C(v0,𝐅0)Eq2×𝐄q22,\begin{split}\bigg{\|}\mathop{\rm ess\,sup}_{0\leq t\leq T}\int_{B_{R}(Rk)}\left(|v|^{2}+|{\bf F}|^{2}\right)dx+\int_{0}^{T}\int_{B_{R}(Rk)}\left(|\nabla v|^{2}+|\nabla{\bf F}|^{2}\right)dx\,dt\bigg{\|}_{\ell^{q/2}(k\in\mathbb{Z}^{3})}\leq C\left\|(v_{0},{\bf F}_{0})\right\|_{E^{2}_{q}\times{\bf E}^{2}_{q}}^{2},\end{split}

for positive constants λ1\lambda_{1} and CC independent of (v0,𝐅0)(v_{0},{\bf F}_{0}) and RR. In particular, if T2=T_{2}=\infty then TT\to\infty as RR\to\infty.

Theorem 1.14 (Existence in Eq2E^{2}_{q} (vNSEd)).

Assume v0,(f1)0,(f2)0,(f3)0Eq2v_{0},(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}\in E^{2}_{q}, where 1q<1\leq q<\infty, are divergence free. Then, there exists a time-global local energy solution (v,𝐅)(v,{\bf F}) and associated pressure π\pi to (vNSEd) in 3\mathbb{R}^{3} with initial data v0,𝐅0v_{0},{\bf F}_{0}, 𝐅0=[(f1)0,(f2)0,(f3)0]{\bf F}_{0}=[(f_{1})_{0},(f_{2})_{0},(f_{3})_{0}], so that, for any 0<T<0<T<\infty,

(v,𝐅)𝐋𝐄q(0,T)<.\begin{split}\left\|(v,{\bf F})\right\|_{{\bf LE}_{q}(0,T)}<\infty.\end{split}

In particular, (v,𝐅)L(0,T;Eq2×𝐄q2)(v,{\bf F})\in L^{\infty}(0,T;E^{2}_{q}\times{\bf E}^{2}_{q}).

The remainder of the paper is organized as follows: In Section 2, we construct mild solutions in critical and subcritical spaces and prove Theorems 1.1, 1.2, 1.3, 1.4, 1.5, and 1.6. In Section 3, we examine properties of local energy solutions, including uniqueness and regularity, establish a priori bounds and explicit growth rate, and prove the global existence results: Theorems 1.8, 1.9, 1.10, 1.12, 1.13, and 1.14.

2 Construction of mild solutions in Wiener amalgam spaces

This section is dedicated to proving Theorems 1.1, 1.2, 1.3, 1.4, 1.5, and 1.6. The proof techniques closely follow those outlined in [1, Section 3]. Specifically, we apply the Picard iteration scheme as in [1, Section 3] to the problems (1.8) and (1.18) to construct mild solutions for both the MHD equations (MHD) and the viscoelastic Navier–Stokes equations with damping (vNSEd), respectively. Since the structure of (vNSEd) is analogous to that of (MHD), we prove Theorems 1.1, 1.2, and 1.3 for (MHD). The proofs of Theorems 1.4, 1.5, and 1.6 for (vNSEd) are omitted for brevity.

2.1 Mild solutions in subcritical spaces: Proof of Theorem 1.1

The proof of Theorem 1.1 is an adaption of the proof of [1, Theorem 1.1] for the Navier–Stokes equations to the MHD equations.

Define

(v,b)T=sup0tT(v,b)(t)Eqr×Eqr.\left\|(v,b)\right\|_{\mathcal{E}_{T}}=\sup_{0\leq t\leq T}\left\|(v,b)(t)\right\|_{E^{r}_{q}\times E^{r}_{q}}.

By [1, Lemma 2.1], we have the linear estimate

(etΔv0,etΔb0)Eqr×Eqr(v0,b0)Eqr×Eqr.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}{\ \lesssim\ }\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}.

So,

(etΔv0,etΔb0)TC1(v0,b0)Eqr×Eqr.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\mathcal{E}_{T}}\leq C_{1}\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}. (2.1)

For bilinear estimate, again by [1, Lemma 2.1], we estimate:

B1((v,b),(u,a))(t)Eqr0t(1(ts)12+32r+1(ts)12)((vu)(s)Eqr/2+(ba)(s)Eqr/2)𝑑s(t1232r+t12)(sup0<τ<tv(τ)Eqrsup0<τ<tu(τ)Er+sup0<τ<tb(τ)Eqrsup0<τ<ta(τ)Er)(t1232r+t12)(sup0<τ<tv(τ)Eqrsup0<τ<tu(τ)Eqr+sup0<τ<tb(τ)Eqrsup0<τ<ta(τ)Eqr)(t1232r+t12)(v,b)T(u,a)T,\begin{split}\|B_{1}&((v,b),(u,a))(t)\|_{E^{r}_{q}}\\ &{\ \lesssim\ }\int_{0}^{t}\bigg{(}\frac{1}{(t-s)^{\frac{1}{2}+\frac{3}{2r}}}+\frac{1}{(t-s)^{\frac{1}{2}}}\bigg{)}\left(\|(v\otimes u)(s)\|_{E^{r/2}_{q}}+\|(b\otimes a)(s)\|_{E^{r/2}_{q}}\right)ds\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left(\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|u(\tau)\|_{E^{r}_{\infty}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|a(\tau)\|_{E^{r}_{\infty}}\right)\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left(\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|u(\tau)\|_{E^{r}_{q}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|a(\tau)\|_{E^{r}_{q}}\right)\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\|(v,b)\|_{\mathcal{E}_{T}}\|(u,a)\|_{\mathcal{E}_{T}},\end{split}

where we used the embedding EqrErE^{r}_{q}\subset E^{r}_{\infty}. Similarly,

B2((v,b),(u,a))(t)Eqr0t(1(ts)12+32r+1(ts)12)((va)(s)Eqr/2+(bu)(s)Eqr/2)𝑑s(t1232r+t12)(sup0<τ<tv(τ)Eqrsup0<τ<ta(τ)Er+sup0<τ<tb(τ)Eqrsup0<τ<tu(τ)Er)(t1232r+t12)(sup0<τ<tv(τ)Eqrsup0<τ<ta(τ)Eqr+sup0<τ<tb(τ)Eqrsup0<τ<tu(τ)Eqr)(t1232r+t12)(v,b)T(u,a)T.\begin{split}\|B_{2}&((v,b),(u,a))(t)\|_{E^{r}_{q}}\\ &{\ \lesssim\ }\int_{0}^{t}\bigg{(}\frac{1}{(t-s)^{\frac{1}{2}+\frac{3}{2r}}}+\frac{1}{(t-s)^{\frac{1}{2}}}\bigg{)}\left(\|(v\otimes a)(s)\|_{E^{r/2}_{q}}+\|(b\otimes u)(s)\|_{E^{r/2}_{q}}\right)ds\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left(\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|a(\tau)\|_{E^{r}_{\infty}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|u(\tau)\|_{E^{r}_{\infty}}\right)\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left(\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|a(\tau)\|_{E^{r}_{q}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|u(\tau)\|_{E^{r}_{q}}\right)\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\|(v,b)\|_{\mathcal{E}_{T}}\|(u,a)\|_{\mathcal{E}_{T}}.\end{split}

Thus, the full bilinear estimate becomes

B((v,b),(u,a))TC2(T1232r+T12)(v,b)T(u,a)T.\left\|B((v,b),(u,a))\right\|_{\mathcal{E}_{T}}\leq C_{2}\left(T^{\frac{1}{2}-\frac{3}{2r}}+T^{\frac{1}{2}}\right)\|(v,b)\|_{\mathcal{E}_{T}}\|(u,a)\|_{\mathcal{E}_{T}}. (2.2)

We look for a solution of the form

(v,b)=(etΔv0,etΔb0)B((v,b),(v,b)).(v,b)=(e^{t\Delta}v_{0},e^{t\Delta}b_{0})-B((v,b),(v,b)).

Suppose TT is small enough so that (v0,b0)Eqr×Eqr<(8C1C2(T1/23/(2r)+T1/2))1\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}<(8C_{1}C_{2}(T^{1/2-3/(2r)}+T^{1/2}))^{-1}. Then by the Picard contraction principle, there exists a unique strong mild solution satisfying

(v,b)T2C1(v0,b0)Eqr×Eqr.\left\|(v,b)\right\|_{\mathcal{E}_{T}}\leq 2C_{1}\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}. (2.3)

To prove continuity at time zero, assume r,q<r,q<\infty. Then

v(t)v0EqrB1(u,u)(t)Eqr+etΔv0v0Eqr(t1232r+t12)[sup0<τ<tv(τ)Eqrsup0<τ<tv(τ)Er+sup0<τ<tb(τ)Eqrsup0<τ<tb(τ)Er]+etΔv0v0Eqr(t1232r+t12)[sup0<τ<tv(τ)Eqrsup0<τ<tv(τ)Eqr+sup0<τ<tb(τ)Eqrsup0<τ<tb(τ)Eqr]+etΔv0v0Eqr.\begin{split}\|v(t)-v_{0}\|_{E^{r}_{q}}&\leq\|B_{1}(u,u)(t)\|_{E^{r}_{q}}+\|e^{t\Delta}v_{0}-v_{0}\|_{E^{r}_{q}}\\ &{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left[\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{\infty}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{\infty}}\right]\\ &\qquad+\|e^{t\Delta}v_{0}-v_{0}\|_{E^{r}_{q}}\\ &\leq\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left[\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}+\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\right]\\ &\qquad+\|e^{t\Delta}v_{0}-v_{0}\|_{E^{r}_{q}}.\end{split}

Both terms tend to zero as t0+t\to 0^{+}, the latter by [1, Lemma 2.3]. Hence, v(t)v0Eqr0\|v(t)-v_{0}\|_{E^{r}_{q}}\to 0 as t0+t\to 0^{+}. Similarly,

b(t)b0Eqr(t1232r+t12)[sup0<τ<tv(τ)Eqrsup0<τ<tb(τ)Eqr+sup0<τ<tv(τ)Eqrsup0<τ<tb(τ)Eqr]+etΔb0b0Eqr0 as t0+.\begin{split}\|b(t)-b_{0}\|_{E^{r}_{q}}&{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left[\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}+\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\right]\\ &\qquad+\|e^{t\Delta}b_{0}-b_{0}\|_{E^{r}_{q}}\to 0\text{ as }t\to 0^{+}.\end{split}

The continuity at t(0,T)t\in(0,T) can be shown as usual, see e.g., [40, lines 3-8, page 86], including r=r=\infty or q=q=\infty.

If either r=r=\infty or q=q=\infty, the semigroup terms no longer vanish, but the bilinear terms still tend to zero:

(t1232r+t12)sup0<τ<tv(τ)Eqrsup0<τ<tv(τ)Eqr0,(t1232r+t12)sup0<τ<tb(τ)Eqrsup0<τ<tb(τ)Eqr0,\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\to 0,\ \left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\to 0,

and

(t1232r+t12)sup0<τ<tv(τ)Eqrsup0<τ<tb(τ)Eqr0.\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\sup_{0<\tau<t}\|v(\tau)\|_{E^{r}_{q}}\sup_{0<\tau<t}\|b(\tau)\|_{E^{r}_{q}}\to 0.

Hence,

v(t)etΔv0Eqr+b(t)etΔb0Eqr0 as t0+,\|v(t)-e^{t\Delta}v_{0}\|_{E^{r}_{q}}+\|b(t)-e^{t\Delta}b_{0}\|_{E^{r}_{q}}\to 0\text{ as }t\to 0^{+},

as asserted in the theorem.

For uniqueness, let (v,b),(v,b)L(0,T;Eqr×Eqr)C((0,T);Eqr×Eqr)(v,b),(v^{\prime},b^{\prime})\in L^{\infty}(0,T;E^{r}_{q}\times E^{r}_{q})\cap C((0,T);E^{r}_{q}\times E^{r}_{q}) be two mild solutions with initial data (v0,b0)Eqr×Eqr(v_{0},b_{0})\in E^{r}_{q}\times E^{r}_{q}. Then, for 0<t<TT0<t<T^{\prime}\leq T,

(vv)(t)EqrB1((vv,bb),(v,b))(t)Eqr+B1((v,b),(vv,bb))(t)Eqr(t1232r+t12)((v,b)T+(v,b)T)(sup0<t<T(vv)(t)Eqr+sup0<t<T(bb)(t)Eqr),\begin{split}&\left\|(v-v^{\prime})(t)\right\|_{E^{r}_{q}}\leq\left\|B_{1}((v-v^{\prime},b-b^{\prime}),(v,b))(t)\right\|_{E^{r}_{q}}+\left\|B_{1}((v^{\prime},b^{\prime}),(v-v^{\prime},b-b^{\prime}))(t)\right\|_{E^{r}_{q}}\\ &\quad{\ \lesssim\ }\left(t^{\frac{1}{2}-\frac{3}{2r}}+t^{\frac{1}{2}}\right)\left(\left\|(v,b)\right\|_{\mathcal{E}_{T}}+\left\|(v^{\prime},b^{\prime})\right\|_{\mathcal{E}_{T}}\right)\left(\sup_{0<t<T^{\prime}}\|(v-v^{\prime})(t)\|_{E^{r}_{q}}+\sup_{0<t<T^{\prime}}\|(b-b^{\prime})(t)\|_{E^{r}_{q}}\right),\end{split}

so that we have

sup0<t<T(vv)(t)Eqr(T1232r+T12)((v,b)T+(v,b)T)sup0<t<T(sup0<t<T(vv)(t)Eqr+sup0<t<T(bb)(t)Eqr).\begin{split}\sup_{0<t<T^{\prime}}\|(v-v^{\prime})(t)\|_{E^{r}_{q}}&{\ \lesssim\ }\left(T^{\prime\frac{1}{2}-\frac{3}{2r}}+T^{\prime\frac{1}{2}}\right)\left(\left\|(v,b)\right\|_{\mathcal{E}_{T}}+\left\|(v^{\prime},b^{\prime})\right\|_{\mathcal{E}_{T}}\right)\\ &\qquad\cdot\sup_{0<t<T^{\prime}}\left(\sup_{0<t<T^{\prime}}\|(v-v^{\prime})(t)\|_{E^{r}_{q}}+\sup_{0<t<T^{\prime}}\|(b-b^{\prime})(t)\|_{E^{r}_{q}}\right).\end{split}

Similarly,

sup0<t<T(bb)(t)Eqr(T1232r+T12)((v,b)T+(v,b)T)sup0<t<T(sup0<t<T(vv)(t)Eqr+sup0<t<T(bb)(t)Eqr).\begin{split}\sup_{0<t<T^{\prime}}\|(b-b^{\prime})(t)\|_{E^{r}_{q}}&{\ \lesssim\ }\left(T^{\prime\frac{1}{2}-\frac{3}{2r}}+T^{\prime\frac{1}{2}}\right)\left(\left\|(v,b)\right\|_{\mathcal{E}_{T}}+\left\|(v^{\prime},b^{\prime})\right\|_{\mathcal{E}_{T}}\right)\\ &\qquad\cdot\sup_{0<t<T^{\prime}}\left(\sup_{0<t<T^{\prime}}\|(v-v^{\prime})(t)\|_{E^{r}_{q}}+\sup_{0<t<T^{\prime}}\|(b-b^{\prime})(t)\|_{E^{r}_{q}}\right).\end{split}

Thus, for small enough T>0T^{\prime}>0, this implies (v,b)=(v,b)(v,b)=(v^{\prime},b^{\prime}) on (0,T)(0,T^{\prime}), and repeating the argument yields uniqueness on (0,T)(0,T).

To obtain the spacetime integral bound, assume 3<rs3<r\leq s\leq\infty, rp<r\leq p<\infty, 2s+3p=3r\frac{2}{s}+\frac{3}{p}=\frac{3}{r}, and 1q=m1\leq q=m\leq\infty. Consider the Banach space

XT=T(ET,qs,p×ET,qs,p).X_{T}=\mathcal{E}_{T}\cap(E^{s,p}_{T,q}\times E^{s,p}_{T,q}).

We can assume m=qm=q since fET,ms,pfET,qs,p\left\|f\right\|_{E^{s,p}_{T,m}}\leq\left\|f\right\|_{E^{s,p}_{T,q}} for mqm\geq q. From 3p=3r2s3r2r\frac{3}{p}=\frac{3}{r}-\frac{2}{s}\geq\frac{3}{r}-\frac{2}{r} we get p3r<p\leq 3r<\infty. For the linear term, by (2.1) and [1, Lemma 2.4] (which needs r<r<\infty and rsr\leq s), we have for a fixed ϵ>0\epsilon>0 that

(etΔv0,etΔb0)XT=(etΔv0,etΔb0)T+(etΔv0,etΔb0)ET,qs,p×ET,qs,pC3(1+T1/s+ϵ)(v0,b0)Eqr×Eqr.\begin{split}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{X_{T}}&=\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\mathcal{E}_{T}}+\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{s,p}_{T,q}\times E^{s,p}_{T,q}}\\ &\leq C_{3}(1+T^{1/s+\epsilon})\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}.\end{split}

For the bilinear term, by (2.2) and [1, Lemma 2.7] with p~=p/2\tilde{p}=p/2 and s~=s/2\tilde{s}=s/2 so that σ=1232r>0\sigma=\frac{1}{2}-\frac{3}{2r}>0 due to r>3r>3, (allowing s=s=\infty),

B((v,b),(u,a))XT=B((v,b),(u,a))T+B1((v,b),(u,a))ET,qs,p+B2((v,b),(u,a))ET,qs,pC4[(T1232r+T12)(v,b)T(u,a)T+(T1232r+T11s)(vuET,qs2,p2+baET,qs2,p2+vaET,qs2,p2+buET,qs2,p2)].\begin{split}&\left\|B((v,b),(u,a))\right\|_{X_{T}}=\left\|B((v,b),(u,a))\right\|_{\mathcal{E}_{T}}+\left\|B_{1}((v,b),(u,a))\right\|_{E^{s,p}_{T,q}}+\left\|B_{2}((v,b),(u,a))\right\|_{E^{s,p}_{T,q}}\\ &\quad\leq C_{4}\left[\left(T^{\frac{1}{2}-\frac{3}{2r}}+T^{\frac{1}{2}}\right)\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{E}_{T}}\right.\\ &\qquad\left.+\left(T^{\frac{1}{2}-\frac{3}{2r}}+T^{1-\frac{1}{s}}\right)\left(\left\|v\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,q}}+\left\|b\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,q}}+\left\|v\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,q}}+\left\|b\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,q}}\right)\right].\end{split}

Since

fgET,qs2,p2fET,qs,pgET,s,pfET,qs,pgET,qs,p,\left\|f\otimes g\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,q}}\leq\left\|f\right\|_{E^{s,p}_{T,q}}\left\|g\right\|_{E^{s,p}_{T,\infty}}\leq\left\|f\right\|_{E^{s,p}_{T,q}}\left\|g\right\|_{E^{s,p}_{T,q}},

we derive

B((v,b),(u,a))XT2C4(T1232r+T11s)(v,b)XT(u,a)XT.\left\|B((v,b),(u,a))\right\|_{X_{T}}\leq 2C_{4}\left(T^{\frac{1}{2}-\frac{3}{2r}}+T^{1-\frac{1}{s}}\right)\left\|(v,b)\right\|_{X_{T}}\left\|(u,a)\right\|_{X_{T}}.

Choose T>0T>0 small enough such that (v0,b0)Eqr×Eqr<[8C3C4(1+T1s+ϵ)(T1232r+T11s)]1\left\|(v_{0},b_{0})\right\|_{E^{r}_{q}\times E^{r}_{q}}<[8C_{3}C_{4}(1+T^{\frac{1}{s}+\epsilon})(T^{\frac{1}{2}-\frac{3}{2r}}+T^{1-\frac{1}{s}})]^{-1}, the Picard iteration yields a unique strong mild solution (v~,b~)XT(\tilde{v},\tilde{b})\in X_{T}. Since XTLT(Eqr×Eqr)X_{T}\subset L^{\infty}_{T}(E^{r}_{q}\times E^{r}_{q}), uniqueness implies (v~,b~)(\tilde{v},\tilde{b}), and the solution satisfies the desired spacetime integral bound. This completes the proof of Theorem 1.1. ∎

2.2 Mild solutions in critical spaces with small data: Proof of Theorem 1.2

The proof of Theorem 1.2 is an adaption of the proof of [1, Theorem 1.2] for the Navier–Stokes equations to the MHD equations.

Let

(v,b)~T:=sup0<t<T(v,b)(,t)Eq3×Eq3+sup0<t<Tt12(v,b)(,t)Eq×Eq\left\|(v,b)\right\|_{\tilde{\mathcal{E}}_{T}}:=\sup_{0<t<T}\left\|(v,b)(\cdot,t)\right\|_{E^{3}_{q}\times E^{3}_{q}}+\sup_{0<t<T}t^{\frac{1}{2}}\left\|(v,b)(\cdot,t)\right\|_{E^{\infty}_{q}\times E^{\infty}_{q}}

and

(v,b)~T:=sup0<t<Tt14(v,b)(,t)Eq6×Eq6.\left\|(v,b)\right\|_{\tilde{\mathcal{F}}_{T}}:=\sup_{0<t<T}t^{\frac{1}{4}}\left\|(v,b)(\cdot,t)\right\|_{E^{6}_{q}\times E^{6}_{q}}.

The inclusion ~T~T\tilde{\mathcal{E}}_{T}\subset\tilde{\mathcal{F}}_{T} is obvious. We define the spaces

T:=~TET,ms,p and T:=~TET,ms,p,\mathcal{E}_{T}:=\tilde{\mathcal{E}}_{T}\cap E^{s,p}_{T,m}\qquad\text{ and }\qquad\mathcal{F}_{T}:=\tilde{\mathcal{F}}_{T}\cap E^{s,p}_{T,m}, (2.4)

with corresponding norms

(v,b)T:=(v,b)~T+(v,b)ET,ms,p and (v,b)T:=(v,b)~T+(v,b)ET,ms,p.\left\|(v,b)\right\|_{\mathcal{E}_{T}}:=\left\|(v,b)\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|(v,b)\right\|_{E^{s,p}_{T,m}}\qquad\text{ and }\qquad\left\|(v,b)\right\|_{\mathcal{F}_{T}}:=\left\|(v,b)\right\|_{\tilde{\mathcal{F}}_{T}}+\left\|(v,b)\right\|_{E^{s,p}_{T,m}}.

The inclusion TT\mathcal{E}_{T}\subset\mathcal{F}_{T} follows immediately.

From [1, (2.16)], and choosing ϵ>32m32q\epsilon>\frac{3}{2m}-\frac{3}{2q} from the definition of β\beta in [1, Lemma 2.4], we have

(etΔv0,etΔb0)ET,ms,p×ET,ms,p(1+T1s+ϵ)(v0,b0)Eq3×Eq3.\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}{\ \lesssim\ }(1+T^{\frac{1}{s}+\epsilon})\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}. (2.5)

Also, by [1, Lemma 2.1],

t14(etΔv0,etΔb0)Eq6×Eq6(1+T14)(v0,b0)Eq3×Eq3.t^{\frac{1}{4}}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{E^{6}_{q}\times E^{6}_{q}}{\ \lesssim\ }(1+T^{\frac{1}{4}})\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}. (2.6)

Combining, we obtain

(etΔv0,etΔb0)TC1(1+T1s+ϵ+T14)(v0,b0)Eq3×Eq3.\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}\leq C_{1}(1+T^{\frac{1}{s}+\epsilon}+T^{\frac{1}{4}})\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}. (2.7)

Using [1, Lemma 2.7] with s~=s/2\tilde{s}=s/2, p~=p/2\tilde{p}=p/2, and m~=m\tilde{m}=m (so that σ=0\sigma=0), and applying Hölder’s inequality, we estimate the bilinear terms:

B1((v,b),(u,a))ET,ms,pC(1+T11s)(vuET,ms2,p2+baET,ms2,p2)C(1+T11s)(vET,ms,puET,s,p+bET,ms,paET,s,p)C(1+T11s)(vET,ms,puET,ms,p+bET,ms,paET,ms,p),\begin{split}\left\|B_{1}((v,b),(u,a))\right\|_{E^{s,p}_{T,m}}&\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,m}}+\left\|b\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,m}}\right)\\ &\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\right\|_{E^{s,p}_{T,m}}\left\|u\right\|_{E^{s,p}_{T,\infty}}+\left\|b\right\|_{E^{s,p}_{T,m}}\left\|a\right\|_{E^{s,p}_{T,\infty}}\right)\\ &\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\right\|_{E^{s,p}_{T,m}}\left\|u\right\|_{E^{s,p}_{T,m}}+\left\|b\right\|_{E^{s,p}_{T,m}}\left\|a\right\|_{E^{s,p}_{T,m}}\right),\end{split} (2.8)

where we used the inclusion ET,ms,pET,s,pE^{s,p}_{T,m}\subset E^{s,p}_{T,\infty} in the last inequality. Similarly, we have

B2((v,b),(u,a))ET,ms,pC(1+T11s)(vaET,ms2,p2+buET,ms2,p2)C(1+T11s)(vET,ms,paET,s,p+bET,ms,puET,s,p)C(1+T11s)(vET,ms,paET,ms,p+bET,ms,puET,ms,p).\begin{split}\left\|B_{2}((v,b),(u,a))\right\|_{E^{s,p}_{T,m}}&\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,m}}+\left\|b\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,m}}\right)\\ &\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\right\|_{E^{s,p}_{T,m}}\left\|a\right\|_{E^{s,p}_{T,\infty}}+\left\|b\right\|_{E^{s,p}_{T,m}}\left\|u\right\|_{E^{s,p}_{T,\infty}}\right)\\ &\leq C(1+T^{1-\frac{1}{s}})\left(\left\|v\right\|_{E^{s,p}_{T,m}}\left\|a\right\|_{E^{s,p}_{T,m}}+\left\|b\right\|_{E^{s,p}_{T,m}}\left\|u\right\|_{E^{s,p}_{T,m}}\right).\end{split} (2.9)

Hence,

B((v,b),(u,a))ET,ms,p×ET,ms,pC(1+T11s)(v,b)T(u,a)T.\begin{split}\left\|B((v,b),(u,a))\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}&\leq C(1+T^{1-\frac{1}{s}})\left\|(v,b)\right\|_{\mathcal{F}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}.\end{split} (2.10)

Additionally, applying [1, Lemma 2.1] and Hölder inequality (1.3), we obtain

B1((v,b),(u,a)Eq6(t)0t(1(tτ)12+1(tτ)34)((vu)(τ)Eq3+(ba)(τ)Eq3)𝑑τ0t(1(tτ)12+1(tτ)34)(v(τ)Eq6u(τ)Eq6+b(τ)Eq6a(τ)Eq6)𝑑τ(1+t1/4)(v,b)~T(u,a)~T.\begin{split}\left\|B_{1}((v,b),(u,a)\right\|_{E^{6}_{q}}(t)&{\ \lesssim\ }\int_{0}^{t}\left(\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\right)\left(\left\|(v\otimes u)(\tau)\right\|_{E^{3}_{q}}+\left\|(b\otimes a)(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\left(\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\right)\left(\left\|v(\tau)\right\|_{E^{6}_{q}}\left\|u(\tau)\right\|_{E^{6}_{q}}+\left\|b(\tau)\right\|_{E^{6}_{q}}\left\|a(\tau)\right\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+t^{-1/4})\left\|(v,b)\right\|_{\tilde{\mathcal{F}}_{T}}\left\|(u,a)\right\|_{\tilde{\mathcal{F}}_{T}}.\end{split}

Similarly,

B2((v,b),(u,a)Eq6(t)0t(1(tτ)12+1(tτ)34)((va)(τ)Eq3+(bu)(τ)Eq3)𝑑τ0t(1(tτ)12+1(tτ)34)(v(τ)Eq6a(τ)Eq6+b(τ)Eq6u(τ)Eq6)𝑑τ(1+t1/4)(v,b)~T(u,a)~T.\begin{split}\left\|B_{2}((v,b),(u,a)\right\|_{E^{6}_{q}}(t)&{\ \lesssim\ }\int_{0}^{t}\left(\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\right)\left(\left\|(v\otimes a)(\tau)\right\|_{E^{3}_{q}}+\left\|(b\otimes u)(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\left(\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\right)\left(\left\|v(\tau)\right\|_{E^{6}_{q}}\left\|a(\tau)\right\|_{E^{6}_{q}}+\left\|b(\tau)\right\|_{E^{6}_{q}}\left\|u(\tau)\right\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+t^{-1/4})\left\|(v,b)\right\|_{\tilde{\mathcal{F}}_{T}}\left\|(u,a)\right\|_{\tilde{\mathcal{F}}_{T}}.\end{split}

Hence

B((v,b),(u,a))TC2(1+T14+T11s)(v,b)T(u,a)T.\|B((v,b),(u,a))\|_{\mathcal{F}_{T}}\leq C_{2}(1+T^{\frac{1}{4}}+T^{1-\frac{1}{s}})\|(v,b)\|_{\mathcal{F}_{T}}\|(u,a)\|_{\mathcal{F}_{T}}.

Taking (v0,b0)Eq3×Eq3\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}} small enough, by (2.7), it is possible to make

(etΔv0,etΔb0)T[4C2(1+T1/4+T11s)]1.\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}\leq\big{[}4C_{2}(1+T^{1/4}+T^{1-\frac{1}{s}})\big{]}^{-1}. (2.11)

The Picard iteration yields a mild solution (v,b)T(v,b)\in\mathcal{F}_{T} to (MHD) so that

(v,b)T2(etΔv0,etΔb0)T.\|(v,b)\|_{\mathcal{F}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}.

This solution is unique among all mild solutions (u,a)(u,a) with data (v0,b0)(v_{0},b_{0}) satisfying (u,a)T2(etΔv0,etΔb0)T\|(u,a)\|_{\mathcal{F}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}. In fact, since we can also apply the Picard contraction to ~T\tilde{\mathcal{F}}_{T}, the solution is also unique in the class (u,a)~T2(etΔv0,etΔb0)~T\|(u,a)\|_{\tilde{\mathcal{F}}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\tilde{\mathcal{F}}_{T}}.

Next, we show that a solution (v,b)T(v,b)\in\mathcal{F}_{T} with small enough initial data (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q} also belongs to T\mathcal{E}_{T}. Let {(v(n),b(n))}n1\{(v^{(n)},b^{(n)})\}_{n\geq 1} be the Picard iteration sequence in T\mathcal{F}_{T}. By construction,

(v(n),b(n))T2C1(1+T1s+ϵ+T14)(v0,b0)Eq3×Eq3<1.\left\|(v^{(n)},b^{(n)})\right\|_{{\mathcal{F}}_{T}}\leq 2C_{1}(1+T^{\frac{1}{s}+\epsilon}+T^{\frac{1}{4}})\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}<1.

Note that

(v(n),b(n))~T(etΔv0,etΔb0)~T+B((v(n1),b(n1)),(v(n1),b(n1)))~T.\left\|(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}}\leq\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|B((v^{(n-1)},b^{(n-1)}),(v^{(n-1)},b^{(n-1)}))\right\|_{\tilde{\mathcal{E}}_{T}}.

By [1, Lemma 2.1],

(etΔv0,etΔb0)~TC(1+T1/2)(v0,b0)Eq3×Eq3.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\tilde{\mathcal{E}}_{T}}\leq C(1+T^{1/2})\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}.

As is usual in arguments like this, we now seek estimates for B((v,b),(u,a))B((v,b),(u,a)) in ~T\tilde{\mathcal{E}}_{T} in terms of measurements of (v,b)(v,b) and (u,a)(u,a) in ~T\tilde{\mathcal{F}}_{T} and ~T\tilde{\mathcal{E}}_{T}. We have by [1, Lemma 2.1] and Hölder inequality (1.3),

B1((v,b),(u,a))Eq30t(1(tτ)12+1(tτ)34)((vu)(τ)Eq2+(ba)(τ)Eq2)𝑑τ0t(1(tτ)12+1(tτ)34)(v(τ)Eq3u(τ)Eq6+b(τ)Eq3a(τ)Eq6)𝑑τ(1+T1/4)(v,b)~T(u,a)~T,\begin{split}\|B_{1}((v,b),(u,a))\|_{E^{3}_{q}}&{\ \lesssim\ }\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|(v\otimes u)(\tau)\|_{E^{2}_{q}}+\|(b\otimes a)(\tau)\|_{E^{2}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|v(\tau)\|_{E^{3}_{q}}\|u(\tau)\|_{E^{6}_{q}}+\|b(\tau)\|_{E^{3}_{q}}\|a(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}},\end{split} (2.12)

and

B2((v,b),(u,a))Eq30t(1(tτ)12+1(tτ)34)((va)(τ)Eq2+(bu)(τ)Eq2)𝑑τ0t(1(tτ)12+1(tτ)34)(v(τ)Eq3a(τ)Eq6+b(τ)Eq3u(τ)Eq6)𝑑τ(1+T1/4)(v,b)~T(u,a)~T,\begin{split}\|B_{2}((v,b),(u,a))\|_{E^{3}_{q}}&{\ \lesssim\ }\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|(v\otimes a)(\tau)\|_{E^{2}_{q}}+\|(b\otimes u)(\tau)\|_{E^{2}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|v(\tau)\|_{E^{3}_{q}}\|a(\tau)\|_{E^{6}_{q}}+\|b(\tau)\|_{E^{3}_{q}}\|u(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}},\end{split} (2.13)

which imply

B((v,b),(u,a))Eq3×Eq3(1+T1/4)(v,b)~T(u,a)~T.\begin{split}\|B((v,b),(u,a))\|_{E^{3}_{q}\times E^{3}_{q}}{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}}.\end{split} (2.14)

Moreover, we have

t12B1((v,b),(u,a))Eqt120t(1(tτ)12+1(tτ)34)((vu)(τ)Eq6+(ba)(τ)Eq6)𝑑τt120t(1(tτ)12+1(tτ)34)(v(τ)Equ(τ)Eq6+b(τ)Eqa(τ)Eq6)𝑑τ(1+T1/4)(v,b)~T(u,a)~T,\begin{split}t^{\frac{1}{2}}\|B_{1}((v,b),(u,a))\|_{E^{\infty}_{q}}&{\ \lesssim\ }t^{\frac{1}{2}}\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|(v\otimes u)(\tau)\|_{E^{6}_{q}}+\|(b\otimes a)(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &\leq t^{\frac{1}{2}}\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|v(\tau)\|_{E^{\infty}_{q}}\|u(\tau)\|_{E^{6}_{q}}+\|b(\tau)\|_{E^{\infty}_{q}}\|a(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}},\end{split}

and

t12B2((v,b),(u,a))Eqt120t(1(tτ)12+1(tτ)34)((va)(τ)Eq6+(bu)(τ)Eq6)𝑑τt120t(1(tτ)12+1(tτ)34)(v(τ)Eqa(τ)Eq6+b(τ)Equ(τ)Eq6)𝑑τ(1+T1/4)(v,b)~T(u,a)~T,\begin{split}t^{\frac{1}{2}}\|B_{2}((v,b),(u,a))\|_{E^{\infty}_{q}}&{\ \lesssim\ }t^{\frac{1}{2}}\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|(v\otimes a)(\tau)\|_{E^{6}_{q}}+\|(b\otimes u)(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &\leq t^{\frac{1}{2}}\int_{0}^{t}\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}}+\frac{1}{(t-\tau)^{\frac{3}{4}}}\bigg{)}\left(\|v(\tau)\|_{E^{\infty}_{q}}\|a(\tau)\|_{E^{6}_{q}}+\|b(\tau)\|_{E^{\infty}_{q}}\|u(\tau)\|_{E^{6}_{q}}\right)d\tau\\ &{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}},\end{split}

which imply

t12B2((v,b),(u,a))Eq×Eq(1+T1/4)(v,b)~T(u,a)~T.\begin{split}t^{\frac{1}{2}}\|B_{2}((v,b),(u,a))\|_{E^{\infty}_{q}\times E^{\infty}_{q}}{\ \lesssim\ }(1+T^{1/4})\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}}.\end{split}

By switching (v,b),(u,a)(v,b),(u,a) in the estimates,

B((v,b),(u,a))~T(1+T1/4)min((v,b)~T(u,a)~T,(u,a)~T(v,b)~T).\|B((v,b),(u,a))\|_{\tilde{\mathcal{E}}_{T}}{\ \lesssim\ }(1+T^{1/4})\min\left(\|(v,b)\|_{\tilde{\mathcal{E}}_{T}}\|(u,a)\|_{\tilde{\mathcal{F}}_{T}},\,\|(u,a)\|_{\tilde{\mathcal{E}}_{T}}\|(v,b)\|_{\tilde{\mathcal{F}}_{T}}\right).

We now return to our main objective: proving that the Picard iterates {v(n),b(n)}\{v^{(n)},b^{(n)}\} are uniformly bounded in ~T\tilde{\mathcal{E}}_{T}. From the recursive relation, we have

(v(n),b(n))~T(etΔv0,etΔb0)~T+B((v(n1),b(n1)),(v(n1),b(n1)))~TCT(v0,b0)Eq3×Eq3+CT(v(n1),b(n1))~T(v(n1),b(n1))~TCT(v0,b0)Eq3×Eq3+CTCT′′(v0,b0)Eq3×Eq3(v(n1),b(n1))~T.\begin{split}\left\|(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}}&\leq\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|B((v^{(n-1)},b^{(n-1)}),(v^{(n-1)},b^{(n-1)}))\right\|_{\tilde{\mathcal{E}}_{T}}\\ &\leq C_{T}\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}+C_{T}^{\prime}\left\|(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{F}}_{T}}\left\|(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{E}}_{T}}\\ &\leq C_{T}\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}+C_{T}^{\prime}C_{T}^{\prime\prime}\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}\left\|(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{E}}_{T}}.\end{split}

Thus, if (v0×b0)Eq3×Eq3(2CTCT′′)1\|(v_{0}\times b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}\leq(2C_{T}^{\prime}C_{T}^{\prime\prime})^{-1}, then (v(n),b(n))~T\left\|(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}} is uniformly bounded by 2CT(v0,b0)Eq3×Eq32C_{T}\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}. To show that the limit (v,b)~T(v,b)\in\tilde{\mathcal{E}}_{T}, consider the difference of successive iterates:

(v(n+1),b(n+1))(v(n),b(n))~T=B((v(n),b(n)),(v(n),b(n)))B((v(n1),b(n1)),(v(n1),b(n1)))~TB((v(n),b(n))(v(n1),b(n1)),(v(n),b(n)))~T+B((v(n1),b(n1)),(v(n),b(n))(v(n1),b(n1)))~T(v(n),b(n))(v(n1),b(n1))~T((v(n),b(n))~T+(v(n1),b(n1))~T)(v(n),b(n))(v(n1),b(n1))~T((v(n),b(n))~T+(v(n1),b(n1))~T).\begin{split}&\left\|(v^{(n+1)},b^{(n+1)})-(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}}\\ &=\left\|B((v^{(n)},b^{(n)}),(v^{(n)},b^{(n)}))-B((v^{(n-1)},b^{(n-1)}),(v^{(n-1)},b^{(n-1)}))\right\|_{\tilde{\mathcal{E}}_{T}}\\ &\leq\left\|B((v^{(n)},b^{(n)})-(v^{(n-1)},b^{(n-1)}),(v^{(n)},b^{(n)}))\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|B((v^{(n-1)},b^{(n-1)}),(v^{(n)},b^{(n)})-(v^{(n-1)},b^{(n-1)}))\right\|_{\tilde{\mathcal{E}}_{T}}\\ &{\ \lesssim\ }\left\|(v^{(n)},b^{(n)})-(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{F}}_{T}}\left(\left\|(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{E}}_{T}}\right)\\ &{\ \lesssim\ }\left\|(v^{(n)},b^{(n)})-(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{E}}_{T}}\left(\left\|(v^{(n)},b^{(n)})\right\|_{\tilde{\mathcal{E}}_{T}}+\left\|(v^{(n-1)},b^{(n-1)})\right\|_{\tilde{\mathcal{E}}_{T}}\right).\end{split} (2.15)

This shows the sequence {(v(n+1),b(n+1))(v(n),b(n))}\{(v^{(n+1)},b^{(n+1)})-(v^{(n)},b^{(n)})\} is Cauchy in ~T\tilde{\mathcal{E}}_{T}, and the limit of {(v(n),b(n))}\{(v^{(n)},b^{(n)})\} lies in T\mathcal{E}_{T}, since convergence in T\mathcal{F}_{T} implies convergence in the full T\mathcal{E}_{T}-norm.

If qsq\leq s, we use the embeddings

fLTsEmpfLTsEqpfET,qs,p,\left\|f\right\|_{L^{s}_{T}E^{p}_{m}}\leq\left\|f\right\|_{L^{s}_{T}E^{p}_{q}}\leq\left\|f\right\|_{E^{s,p}_{T,q}},

from (1.6) to justify the LTsEmpL^{s}_{T}E^{p}_{m}-estimate in (1.23).

Next, we prove convergence to the initial data when q<q<\infty. By [1, Lemma 2.3] we have

limT0+sup0<t<Tt14(etΔv0,etΔb0)Eq6×Eq6=limT0(etΔv0,etΔb0)~T=0,\lim_{T^{\prime}\to 0^{+}}\sup_{0<t<T^{\prime}}t^{\frac{1}{4}}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{E^{6}_{q}\times E^{6}_{q}}=\lim_{T^{\prime}\to 0}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}=0, (2.16)

whenever (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q}. We extend this to the Picard sequence by induction. The base case (v(0),b(0))=(etΔv0,etΔb0)(v^{(0)},b^{(0)})=(e^{t\Delta}v_{0},e^{t\Delta}b_{0}) satisfies (2.16). Suppose the inductive hypothesis holds:

limT0(v(n1),b(n1))~T=0.\lim_{T^{\prime}\to 0}\|(v^{(n-1)},b^{(n-1)})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}=0. (2.17)

Then, from the iteration estimate in the class ~T\tilde{\mathcal{F}}_{T^{\prime}} where we are taking TTT^{\prime}\leq T, we have

(v(n),b(n))~T(etΔv0,etΔb0)~T+B((v(n1),b(n1)),(v(n1),b(n1)))~T(etΔv0,etΔb0)~T+(v(n1),b(n1))~T2,\begin{split}\|(v^{(n)},b^{(n)})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}&\leq\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}+\|B((v^{(n-1)},b^{(n-1)}),(v^{(n-1)},b^{(n-1)}))\|_{\tilde{\mathcal{F}}_{T^{\prime}}}\\ &{\ \lesssim\ }\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}+\|(v^{(n-1)},b^{(n-1)})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}^{2},\end{split}

which implies

limT0(v(n),b(n))~T=0\lim_{T^{\prime}\to 0}\|(v^{(n)},b^{(n)})\|_{\tilde{\mathcal{F}}_{T^{\prime}}}=0 (2.18)

by (2.16) and (2.17). This completes the induction.

Since the Picard sequence converges in ~T\tilde{\mathcal{F}}_{T}, the limit (v,b)(v,b) also satisfies

limT0+(v,b)~T=0.\lim_{T^{\prime}\to 0^{+}}\|(v,b)\|_{\tilde{\mathcal{F}}_{T^{\prime}}}=0. (2.19)

for T>0T^{\prime}>0 sufficiently small. Using (2.14) and (2.19), we find

limT0+sup0<t<TB((v,b),(v,b))Eq3×Eq3(t)=0.\begin{split}\lim_{T^{\prime}\to 0^{+}}\sup_{0<t<T^{\prime}}\|B((v,b),(v,b))\|_{E^{3}_{q}\times E^{3}_{q}}(t)=0.\end{split}

Combining this with [1, Lemma 2.3], we conclude

limt0(v,b)(v0,b0)Eq3×Eq3=0.\lim_{t\to 0}\|(v,b)-(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}=0.

If q=q=\infty then we have a weaker mode of convergence. Fix a ball BB. Take R>0R>0 large so that BBR(0)B\subset B_{R}(0). We re-write the bilinear form as

B((v,b),(v,b))=B((v,b),(vχBR(0),bχBR(0)))+B((v,b),(v(1χBR(0)),b(1χBR(0)))).B((v,b),(v,b))=B((v,b),(v\chi_{B_{R}(0)},b\chi_{B_{R}(0)}))+B((v,b),(v(1-\chi_{B_{R}(0)}),b(1-\chi_{B_{R}(0)}))).

If 1<ω<31<\omega<3 then we have

B1((v,b),(vχBR(0),bχBR(0)))Lω0t1(tτ)12+32(231ω)(|v|2χBR(0)L3/2+|b|2χBR(0)L3/2)(τ)𝑑τRt1232(231ω)(v,b)L(Luloc3×Luloc3)2,\begin{split}\|B_{1}&((v,b),(v\chi_{B_{R}(0)},b\chi_{B_{R}(0)}))\|_{L^{\omega}}\\ &{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}+\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}}\left(\||v|^{2}\chi_{B_{R}(0)}\|_{L^{3/2}}+\||b|^{2}\chi_{B_{R}(0)}\|_{L^{3/2}}\right)(\tau)\,d\tau\\ &\lesssim_{R}t^{\frac{1}{2}-\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}\|(v,b)\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}})}^{2},\end{split}
B2((v,b),(vχBR(0),bχBR(0)))Lω0t1(tτ)12+32(231ω)(vbχBR(0)L3/2+bvχBR(0)L3/2)(τ)𝑑τ0t1(tτ)12+32(231ω)(|v|2χBR(0)L3/2+|b|2χBR(0)L3/2)(τ)𝑑τRt1232(231ω)(v,b)L(Luloc3×Luloc3)2,\begin{split}\|B_{2}&((v,b),(v\chi_{B_{R}(0)},b\chi_{B_{R}(0)}))\|_{L^{\omega}}\\ &{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}+\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}}\left(\|v\otimes b\chi_{B_{R}(0)}\|_{L^{3/2}}+\|b\otimes v\chi_{B_{R}(0)}\|_{L^{3/2}}\right)(\tau)\,d\tau\\ &{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}+\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}}\left(\||v|^{2}\chi_{B_{R}(0)}\|_{L^{3/2}}+\||b|^{2}\chi_{B_{R}(0)}\|_{L^{3/2}}\right)(\tau)\,d\tau\\ &\lesssim_{R}t^{\frac{1}{2}-\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}\|(v,b)\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}})}^{2},\end{split}

by Young’s inequality, so that

B((v,b),(vχBR(0),bχBR(0)))Lω×LωRt1232(231ω)(v,b)L(Luloc3×Luloc3)2.\begin{split}\|B((v,b),(v\chi_{B_{R}(0)},b\chi_{B_{R}(0)}))\|_{L^{\omega}\times L^{\omega}}\lesssim_{R}t^{\frac{1}{2}-\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}\|(v,b)\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}})}^{2}.\end{split}

For any R>0R>0, the above vanishes as t0t\to 0 provided ω<3\omega<3.

By taking R=2maxxB|x|R=2\max_{x\in B}|x|, we can ensure that for all xBx\in B and |y|R|y|\geq R we have 12|y||xy|32|y|\frac{1}{2}|y|\leq|x-y|\leq\frac{3}{2}|y|. Hence, for xBx\in B,

|B((v,b),(vχBR(0)c,bχBR(0)c))(x,t)|0tyBR(0)c1(|xy|+tτ)4(|v|2+|b|2)(y,τ)𝑑y𝑑τtsup0<τ<tk=1124k4R4R2k1|y|<R2k(|v|2+|b|2)(y,τ)𝑑ytsup0<τ<tk=1R2kR424k4(|y|<R2k(|v|3+|b|3)(y,τ)𝑑y)23tk=1R323kR424k4sup0<τ<t(v,b)(τ)Luloc3×Luloc32tRsup0<τ<t(v,b)(τ)Luloc3×Luloc32.\begin{split}|B&((v,b),(v\chi_{B_{R}(0)^{c}},b\chi_{B_{R}(0)^{c}}))(x,t)|\\ &\lesssim\int_{0}^{t}\int_{y\in B_{R}(0)^{c}}\frac{1}{(|x-y|+\sqrt{t-\tau})^{4}}(|v|^{2}+|b|^{2})(y,\tau)\,dy\,d\tau\\ &\lesssim t\sup_{0<\tau<t}\sum_{k=1}^{\infty}\frac{1}{2^{4k-4}R^{4}}\int_{R2^{k-1}\leq|y|<R2^{k}}(|v|^{2}+|b|^{2})(y,\tau)\,dy\\ &\lesssim t\sup_{0<\tau<t}\sum_{k=1}^{\infty}\frac{R2^{k}}{R^{4}2^{4k-4}}\bigg{(}\int_{|y|<R2^{k}}(|v|^{3}+|b|^{3})(y,\tau)\,dy\bigg{)}^{\frac{2}{3}}\\ &\lesssim t\sum_{k=1}^{\infty}\frac{R^{3}2^{3k}}{R^{4}2^{4k-4}}\sup_{0<\tau<t}\|(v,b)(\tau)\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}^{2}\\ &\lesssim\frac{t}{R}\sup_{0<\tau<t}\|(v,b)(\tau)\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}^{2}.\end{split}

Now,

B((v,b),(v,b))(t)Lω(B)×Lω(B)RB((v,b),(vχBR(0)c,bχBR(0)c))(t)Lω+B((v,b),(vχBR(0)c,bχBR(0)c))(t)L(B)Rt1232(231ω)(v,b)L(Luloc3×Luloc3)2+tsup0<τ<t(v,b)(τ)Luloc3×Luloc32.\begin{split}\|B&((v,b),(v,b))(t)\|_{L^{\omega}(B)\times L^{\omega}(B)}\\ &\lesssim_{R}\|B((v,b),(v\chi_{B_{R}(0)^{c}},b\chi_{B_{R}(0)^{c}}))(t)\|_{L^{\omega}}+\|B((v,b),(v\chi_{B_{R}(0)^{c}},b\chi_{B_{R}(0)^{c}}))(t)\|_{L^{\infty}(B)}\\ &\lesssim_{R}t^{\frac{1}{2}-\frac{3}{2}(\frac{2}{3}-\frac{1}{\omega})}\|(v,b)\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}})}^{2}+t\sup_{0<\tau<t}\|(v,b)(\tau)\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}^{2}.\end{split}

Hence

limt0+B((v,b),(v,b))(t)Lω(B)×Lω(B)=0.\lim_{t\to 0^{+}}\|B((v,b),(v,b))(t)\|_{L^{\omega}(B)\times L^{\omega}(B)}=0.

Referring to [33, p. 394], we have

limt0(etΔv0,etΔb0)(v0,b0)Lω(B)×Lω(B)=0.\lim_{t\to 0}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})-(v_{0},b_{0})\|_{L^{\omega}(B)\times L^{\omega}(B)}=0.

It follows that

limt0(v,b)(,t)(v0,b0)()Lω(B)×Lω(B)=0.\lim_{t\to 0}\|(v,b)(\cdot,t)-(v_{0},b_{0})(\cdot)\|_{L^{\omega}(B)\times L^{\omega}(B)}=0.

We now prove continuity at positive times. Let t1>0t_{1}>0 be fixed. We will send tt1t\to t_{1}. Note that by [1, Lemma 2.3] we have (etΔv0,etΔb0)(et1Δv0,et1Δb0)0(e^{t\Delta}v_{0},e^{t\Delta}b_{0})-(e^{t_{1}\Delta}v_{0},e^{t_{1}\Delta}b_{0})\to 0 in Eq3×Eq3E^{3}_{q}\times E^{3}_{q} as tt1t\to t_{1}. We therefore only need to show B((v,b),(v,b))(t)B((v,b),(v,b))(t1)B((v,b),(v,b))(t)\to B((v,b),(v,b))(t_{1}). Following [40, p. 86], we take ρ\rho slightly less than 11 so that ρt1<t\rho t_{1}<t and write

B((v,b),(v,b))(t)B((v,b),(v,b))(t1)=ρt1te(tτ)ΔF𝑑τρt1t1e(t1τ)ΔF𝑑τ+0ρt1(e(tρt1)Δe(t1ρt1)Δ)e(ρt1τ)ΔF𝑑τ\begin{split}B((v,b),(v,b))(t)-B((v,b),(v,b))(t_{1})&=\int_{\rho t_{1}}^{t}e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot F\,d\tau{-}\int_{\rho t_{1}}^{t_{1}}e^{(t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F\,d\tau\\ &+\int_{0}^{\rho t_{1}}\big{(}e^{(t-\rho t_{1})\Delta}-e^{(t_{1}-\rho t_{1})\Delta}\big{)}e^{(\rho t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F\,d\tau\end{split}

where F=(vv)(τ)F=(v\otimes v)(\tau), (bb)(τ)(b\otimes b)(\tau), (vb)(τ)(v\otimes b)(\tau), or (bv)(τ)(b\otimes v)(\tau). For the first and second terms, by [1, (2.3)] with p=p~=3p=\tilde{p}=3, q~=q\tilde{q}=q and using the embedding EqEE^{\infty}_{q}\subset E^{\infty}_{\infty}, we have

ρt1te(tτ)ΔFEq3𝑑τρt1t1(tτ)12τ12v,bEq3(τ)τ12v,bEqdτ(tρt1)12(ρt1)12(v,b)T2,\begin{split}\int_{\rho t_{1}}^{t}\|e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot F\|_{E^{3}_{q}}\,d\tau&\lesssim\int_{\rho t_{1}}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}\tau^{\frac{1}{2}}}\|v,b\|_{E^{3}_{q}}(\tau)\tau^{\frac{1}{2}}\|v,b\|_{E^{\infty}_{q}}\,d\tau\lesssim\frac{(t-\rho t_{1})^{\frac{1}{2}}}{(\rho t_{1})^{\frac{1}{2}}}\|(v,b)\|_{\mathcal{E}_{T}}^{2},\end{split}

and

ρt1t1e(t1τ)ΔFEq3𝑑τρt1t11(t1τ)12τ12v,bEq3(τ)τ12v,bEqdτ(t1ρt1)12(ρt1)12(v,b)T2,\begin{split}\int_{\rho t_{1}}^{t_{1}}\|e^{(t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F\|_{E^{3}_{q}}\,d\tau&\lesssim\int_{\rho t_{1}}^{t_{1}}\frac{1}{(t_{1}-\tau)^{\frac{1}{2}}\tau^{\frac{1}{2}}}\|v,b\|_{E^{3}_{q}}(\tau)\tau^{\frac{1}{2}}\|v,b\|_{E^{\infty}_{q}}\,d\tau\lesssim\frac{(t_{1}-\rho t_{1})^{\frac{1}{2}}}{(\rho t_{1})^{\frac{1}{2}}}\|(v,b)\|_{\mathcal{E}_{T}}^{2},\end{split}

both of which can be made arbitrarily small by taking ρt1\rho t_{1} close to t1t_{1} and tt close to t1t_{1}.

For the third term we note that by [1, Lemma 2.3], for each 0<τ<ρt10<\tau<\rho t_{1}, we have

(e(tρt1)Δe(t1ρt1)Δ)e(ρt1τ)ΔF(τ)Eq30 as tt1,\|\big{(}e^{(t-\rho t_{1})\Delta}-e^{(t_{1}-\rho t_{1})\Delta}\big{)}e^{(\rho t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F(\tau)\|_{E^{3}_{q}}\to 0\text{ as }t\to t_{1},

which follows (even if q=q=\infty) the fact that e(ρt1τ)ΔF(τ)Eq3e^{(\rho t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F(\tau)\in E^{3}_{q}, which is a consequence of [1, Lemma 2.1]. Additionally,

(e(tρt1)Δe(t1ρt1)Δ)e(ρt1τ)ΔF(τ)Eq3(1(tτ)12τ12+1(t1τ)12τ12)(v,b)T2L1(0,ρt1),\begin{split}&\|\big{(}e^{(t-\rho t_{1})\Delta}-e^{(t_{1}-\rho t_{1})\Delta}\big{)}e^{(\rho t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F(\tau)\|_{E^{3}_{q}}\\ &\lesssim\bigg{(}\frac{1}{(t-\tau)^{\frac{1}{2}}\tau^{\frac{1}{2}}}+\frac{1}{(t_{1}-\tau)^{\frac{1}{2}}\tau^{\frac{1}{2}}}\bigg{)}\|(v,b)\|_{\mathcal{E}_{T}}^{2}\in L^{1}(0,\rho t_{1}),\end{split}

where integration in L1(0,ρt1)L^{1}(0,\rho t_{1}) is with respect to τ\tau. So, by Lebesgue’s dominated convergence theorem,

0ρt1(e(tρt1)Δe(t1ρt1)Δ)e(ρt1τ)ΔF(τ)Eq3𝑑τ0 as tt1.\int_{0}^{\rho t_{1}}\|\big{(}e^{(t-\rho t_{1})\Delta}-e^{(t_{1}-\rho t_{1})\Delta}\big{)}e^{(\rho t_{1}-\tau)\Delta}\mathbb{P}\nabla\cdot F(\tau)\|_{E^{3}_{q}}\,d\tau\to 0\text{ as }t\to t_{1}.

The above show the continuity of (v,b)(t)(v,b)(t) at positive times.

To prove the spacetime integral bound (1.23) for p=3p=3, s=s=\infty, we work in the spaces with the norms

(v,b)T=(v,b)ET,q,3×ET,q,3+(t12v,t12b)ET,q,×ET,q,,(v,b)T=(t14v,t14b)ET,q,6×ET,q,6.\left\|(v,b)\right\|_{\mathcal{E}_{T}^{*}}=\left\|(v,b)\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}}+\left\|(t^{\frac{1}{2}}v,t^{\frac{1}{2}}b)\right\|_{E^{\infty,\infty}_{T,q}\times E^{\infty,\infty}_{T,q}},\quad\left\|(v,b)\right\|_{\mathcal{F}_{T}^{*}}=\left\|(t^{\frac{1}{4}}v,t^{\frac{1}{4}}b)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}.

Note that TT\mathcal{E}_{T}^{*}\subset\mathcal{F}_{T}^{*}.

For the linear estimate in T\mathcal{F}^{*}_{T}, taking p=6p=6 so that a=1/4a=1/4 in [1, Lemma 3.1], we have

(etΔv0,etΔb0)T=(t14etΔv0,t14etΔb0)ET,q,6×ET,q,6C1(1+T14)(v0,b0)Eq3×Eq3.\begin{split}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\mathcal{F}^{*}_{T}}=\left\|(t^{\frac{1}{4}}e^{t\Delta}v_{0},t^{\frac{1}{4}}e^{t\Delta}b_{0})\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\leq C_{1}^{*}(1+T^{\frac{1}{4}})\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}.\end{split} (2.20)

For bilinear estimate, taking a=b=1/4a=b=1/4, p=p~=6p=\tilde{p}=6 in [1, Lemma 3.2],

B((v,b),(u,a))T=t14B((v,b),(u,a))ET,q,6×ET,q,6C2(1+T34)(t14v,t14b)ET,q,6×ET,q,6(t14u,t14a)ET,q,6×ET,q,6=C2(1+T34)(v,b)T(u,a)T.\begin{split}\left\|B((v,b),(u,a))\right\|_{\mathcal{F}_{T}^{*}}&=\left\|t^{\frac{1}{4}}B((v,b),(u,a))\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\\ &\leq C_{2}^{*}(1+T^{\frac{3}{4}})\left\|(t^{\frac{1}{4}}v,t^{\frac{1}{4}}b)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\left\|(t^{\frac{1}{4}}u,t^{\frac{1}{4}}a)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\\ &=C_{2}^{*}(1+T^{\frac{3}{4}})\left\|(v,b)\right\|_{\mathcal{F}_{T}^{*}}\left\|(u,a)\right\|_{\mathcal{F}_{T}^{*}}.\end{split}

By choosing (v0,b0)Eq3×Eq3\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}} small enough so that (v0,b0)Eq3×Eq3<14C1C2(1+T14)(1+T34)\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}<\frac{1}{4C_{1}^{*}C_{2}^{*}\left(1+T^{\frac{1}{4}}\right)\left(1+T^{\frac{3}{4}}\right)}, Picard iteration yields a unique mild solution satisfying

(v,b)T2C1(1+T14)(v0,b0)Eq3×Eq3.\left\|(v,b)\right\|_{\mathcal{F}_{T}^{*}}\leq 2C_{1}^{*}(1+T^{\frac{1}{4}})\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}.

Now, we claim that any solution (v,b)T(v,b)\in\mathcal{F}^{*}_{T} with sufficiently small (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q} also belongs to T\mathcal{E}^{*}_{T}. By [1, (3.21)] of [1, Lemma 3.1], we have

(etΔv0,etΔb0)ET,q,3×ET,q,3ln(2+T)(v0,b0)Eq3×Eq3.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}}{\ \lesssim\ }\ln(2+T)\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}.

Taking p=p=\infty so that a=1/2a=1/2 in [1, Lemma 3.1], we also obtain

(t12v0,t12b0)ET,q,(1+T12)(v0,b0)Eq3×Eq3.\left\|(t^{\frac{1}{2}}v_{0},t^{\frac{1}{2}}b_{0})\right\|_{E^{\infty,\infty}_{T,q}}{\ \lesssim\ }(1+T^{\frac{1}{2}})\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}.

Combining both estimates, we conclude:

(etΔv0,etΔb0)T(1+T12)(v0,b0)Eq3×Eq3.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\mathcal{E}^{*}_{T}}{\ \lesssim\ }(1+T^{\frac{1}{2}})\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}. (2.21)

Next, we establish the bilinear estimate:

B((v,b),(u,a))TTmin((v,b)T(u,a)T,(u,a)T(v,b)T).\left\|B((v,b),(u,a))\right\|_{\mathcal{E}^{*}_{T}}{\ \lesssim\ }_{T}\min\left(\left\|(v,b)\right\|_{\mathcal{E}^{*}_{T}}\left\|(u,a)\right\|_{\mathcal{F}^{*}_{T}},\left\|(u,a)\right\|_{\mathcal{E}^{*}_{T}}\left\|(v,b)\right\|_{\mathcal{F}^{*}_{T}}\right). (2.22)

Indeed, by applying [1, Lemma 3.2] with (a,b,p,p~)=(0,1/4,3,6)(a,b,p,\tilde{p})=(0,1/4,3,6) and (a,b,p,p~)=(1/2,1/4,,6)(a,b,p,\tilde{p})=(1/2,1/4,\infty,6), we obtain

B((v,b),(u,a))ET,q,3×ET,q,3(1+T34)min((v,b)ET,q,3×ET,q,3(t14u,t14a)ET,q,6×ET,q,6,(u,a)ET,q,3×ET,q,3(t14v,t14b)ET,q,6×ET,q,6),\begin{split}&\left\|B((v,b),(u,a))\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}}\\ &\qquad{\ \lesssim\ }(1+T^{\frac{3}{4}})\min\left(\left\|(v,b)\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}}\left\|(t^{\frac{1}{4}}u,t^{\frac{1}{4}}a)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}},\right.\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\left.\left\|(u,a)\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}}\left\|(t^{\frac{1}{4}}v,t^{\frac{1}{4}}b)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\right),\end{split}

and

t12B((v,b),(u,a))ET,q,×ET,q,(1+T34)min((t12v,t12b)ET,q,×ET,q,(t14u,t14a)ET,q,6×ET,q,6,(t12u,t12a)ET,q,×ET,q,(t14v,t14b)ET,q,6×ET,q,6).\begin{split}&\left\|t^{\frac{1}{2}}B((v,b),(u,a))\right\|_{E^{\infty,\infty}_{T,q}\times E^{\infty,\infty}_{T,q}}\\ &\qquad{\ \lesssim\ }(1+T^{\frac{3}{4}})\min\left(\left\|(t^{\frac{1}{2}}v,t^{\frac{1}{2}}b)\right\|_{E^{\infty,\infty}_{T,q}\times E^{\infty,\infty}_{T,q}}\left\|(t^{\frac{1}{4}}u,t^{\frac{1}{4}}a)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}},\right.\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\left.\left\|(t^{\frac{1}{2}}u,t^{\frac{1}{2}}a)\right\|_{E^{\infty,\infty}_{T,q}\times E^{\infty,\infty}_{T,q}}\left\|(t^{\frac{1}{4}}v,t^{\frac{1}{4}}b)\right\|_{E^{\infty,6}_{T,q}\times E^{\infty,6}_{T,q}}\right).\end{split}

Using the same argument as before, we conclude that (v,b)T(v,b)\in\mathcal{E}^{*}_{T}, possibly after taking a smaller T>0T>0. In particular, (v,b)ET,q,3×ET,q,3(v,b)\in E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}, with the norm controlled by (v0,b0)Eq3×Eq3\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}. The case s=s=\infty in Theorem 1.2 then follows from the embeddings (v,b)LEq3×LEq3(v,b)ET,q,3×ET,q,3\left\|(v,b)\right\|_{L^{\infty}E^{3}_{q}\times L^{\infty}E^{3}_{q}}\leq\left\|(v,b)\right\|_{E^{\infty,3}_{T,q}\times E^{\infty,3}_{T,q}} and (t12v,t12b)LT(Eq×Eq)(t12v,t12b)ET,q,×ET,q,\left\|(t^{\frac{1}{2}}v,t^{\frac{1}{2}}b)\right\|_{L^{\infty}_{T}(E^{\infty}_{q}\times E^{\infty}_{q})}\leq\left\|(t^{\frac{1}{2}}v,t^{\frac{1}{2}}b)\right\|_{E^{\infty,\infty}_{T,q}\times E^{\infty,\infty}_{T,q}}. ∎

2.3 Mild solutions in critical spaces with enough decay: Proof of Theorem 1.3

The proof of Theorem 1.3 is an adaption of the proof of [1, Theorem 1.3] for the Navier–Stokes equations to the MHD equations.

By [1, Lemma 2.1], we have for 1q31\leq q\leq 3,

sup0<t<((etΔv0,etΔb0)Eq3×Eq3+t12(etΔv0,etΔb0)Eq2×+t14(etΔv0,etΔb0)Eq16×Eq16)C(v0,b0)Eq3×Eq3,\begin{split}\sup_{0<t<\infty}&\left(\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}+t^{\frac{1}{2}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{\infty}_{q_{2}}\times}+t^{\frac{1}{4}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{6}_{q_{1}}\times E^{6}_{q_{1}}}\right)\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\leq C\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}},\end{split} (2.23)

where

1q1=1q16,1q2=1q13,q<q1<q2.\frac{1}{q_{1}}=\frac{1}{q}-\frac{1}{6},\quad\frac{1}{q_{2}}=\frac{1}{q}-\frac{1}{3},\quad q<q_{1}<q_{2}\leq\infty. (2.24)

For 0<T0<T\leq\infty and 1q31\leq q\leq 3, let T\mathcal{E}_{T}, T\mathcal{F}_{T} be Banach spaces defined as

T:={(v,b)L(0,T;Eq3×Eq3):t12(v,b)(,t)L(0,T;Eq2×Eq2)},\mathcal{E}_{T}:=\left\{(v,b)\in L^{\infty}(0,T;E^{3}_{q}\times E^{3}_{q}):\ t^{\frac{1}{2}}(v,b)(\cdot,t)\in L^{\infty}(0,T;E^{\infty}_{q_{2}}\times E^{\infty}_{q_{2}})\right\}, (2.25)

and

T:={(v,b):t14(v,b)(,t)L(0,T;Eq16×Eq16)},\mathcal{F}_{T}:=\left\{(v,b):\ t^{\frac{1}{4}}(v,b)(\cdot,t)\in L^{\infty}(0,T;E^{6}_{q_{1}}\times E^{6}_{q_{1}})\right\}, (2.26)

with norms

(v,b)T:=sup0<t<T(v,b)(,t)Eq3×Eq3+sup0<t<Tt12(v,b)(,t)Eq2×Eq2,\left\|(v,b)\right\|_{\mathcal{E}_{T}}:=\sup_{0<t<T}\left\|(v,b)(\cdot,t)\right\|_{E^{3}_{q}\times E^{3}_{q}}+\sup_{0<t<T}t^{\frac{1}{2}}\left\|(v,b)(\cdot,t)\right\|_{E^{\infty}_{q_{2}}\times E^{\infty}_{q_{2}}},

and

(v,b)T:=sup0<t<Tt14(v,b)(,t)Eq16×Eq16,\left\|(v,b)\right\|_{\mathcal{F}_{T}}:=\sup_{0<t<T}t^{\frac{1}{4}}\left\|(v,b)(\cdot,t)\right\|_{E^{6}_{q_{1}}\times E^{6}_{q_{1}}},

respectively. Note that TT\mathcal{E}_{T}\subset\mathcal{F}_{T}.

By [1, Lemma 2.1] again and Hölder inequality (1.3) using 2qq12q\geq q_{1} due to q3q\leq 3,

B1((v,b),(u,a))Eq16(t)0t1(tτ)34(vu(τ)Eq3+ba(τ)Eq3)𝑑τ0t1(tτ)34(v(τ)Eq16u(τ)Eq16+b(τ)Eq16a(τ)Eq16)𝑑τ0t1(tτ)34τ1/4(v,b)Tτ1/4(u,a)T𝑑τt1/4(v,b)T(u,a)T,\begin{split}\left\|B_{1}((v,b),(u,a))\right\|_{E^{6}_{q_{1}}}(t)&{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\left\|v\otimes u(\tau)\right\|_{E^{3}_{q}}+\left\|b\otimes a(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\left\|v(\tau)\right\|_{E^{6}_{q_{1}}}\left\|u(\tau)\right\|_{E^{6}_{q_{1}}}+\left\|b(\tau)\right\|_{E^{6}_{q_{1}}}\left\|a(\tau)\right\|_{E^{6}_{q_{1}}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\tau^{-1/4}\left\|(v,b)\right\|_{{\mathcal{F}}_{T}}\tau^{-1/4}\left\|(u,a)\right\|_{{\mathcal{F}}_{T}}\,d\tau\\ &{\ \lesssim\ }t^{-1/4}\left\|(v,b)\right\|_{{\mathcal{F}}_{T}}\left\|(u,a)\right\|_{{\mathcal{F}}_{T}},\end{split} (2.27)
B2((v,b),(u,a))Eq16(t)0t1(tτ)34(va(τ)Eq3+bu(τ)Eq3)𝑑τ0t1(tτ)34(v(τ)Eq16a(τ)Eq16+b(τ)Eq16u(τ)Eq16)𝑑τ0t1(tτ)34τ1/4(v,b)Tτ1/4(u,a)T𝑑τt1/4(v,b)T(u,a)T,\begin{split}\left\|B_{2}((v,b),(u,a))\right\|_{E^{6}_{q_{1}}}(t)&{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\left\|v\otimes a(\tau)\right\|_{E^{3}_{q}}+\left\|b\otimes u(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\left\|v(\tau)\right\|_{E^{6}_{q_{1}}}\left\|a(\tau)\right\|_{E^{6}_{q_{1}}}+\left\|b(\tau)\right\|_{E^{6}_{q_{1}}}\left\|u(\tau)\right\|_{E^{6}_{q_{1}}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\tau^{-1/4}\left\|(v,b)\right\|_{{\mathcal{F}}_{T}}\tau^{-1/4}\left\|(u,a)\right\|_{{\mathcal{F}}_{T}}\,d\tau\\ &{\ \lesssim\ }t^{-1/4}\left\|(v,b)\right\|_{{\mathcal{F}}_{T}}\left\|(u,a)\right\|_{{\mathcal{F}}_{T}},\end{split} (2.28)

so that

B((v,b),(u,a))Eq16×Eq16(t)t1/4(v,b)T(u,a)T.\left\|B((v,b),(u,a))\right\|_{E^{6}_{q_{1}}\times E^{6}_{q_{1}}}(t){\ \lesssim\ }t^{-1/4}\left\|(v,b)\right\|_{{\mathcal{F}}_{T}}\left\|(u,a)\right\|_{{\mathcal{F}}_{T}}. (2.29)

Hence,

B((v,b),(u,a))Tc(v,b)T(u,a)T,\|B((v,b),(u,a))\|_{{\mathcal{F}}_{T}}\leq c_{*}\|(v,b)\|_{{\mathcal{F}}_{T}}\|(u,a)\|_{{\mathcal{F}}_{T}},

where cc_{*} is a universal constant.

Concerning the caloric extension of (v0,b0)(v_{0},b_{0}), we have for (v0,b0)Eq3×Eq3\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}} of any size that

limT0(etΔv0,etΔb0)T=0,\lim_{T\to 0}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{{\mathcal{F}}_{T}}=0,

by [1, (2.13)] of [1, Lemma 2.3]. Hence, there exists T=T(u0)T=T(u_{0}) so that

(etΔv0,etΔb0)Tc1.\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{{\mathcal{F}}_{T}}\lesssim c_{*}^{-1}. (2.30)

If, on the other hand, (v0,b0)Eq3×Eq3c1\|(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}\lesssim c_{*}^{-1}, then by (2.23), we have (2.30) for T=T=\infty. The Picard contraction theorem then guarantees the existence of a mild solution (v,b)(v,b) to (MHD) so that

(v,b)T2(etΔv0,etΔb0)T.\|(v,b)\|_{\mathcal{F}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}.

This solution is unique among all mild solutions (u,a)(u,a) with data (v0,b0)(v_{0},b_{0}) satisfying (u,a)T2(etΔv0,etΔb0)T\|(u,a)\|_{\mathcal{F}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}.

Next, we show that a solution (v,b)T(v,b)\in\mathcal{F}_{T} with initial data (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q} also belongs to T\mathcal{E}_{T}. Let {(v(n),b(n))}n1\{(v^{(n)},b^{(n)})\}_{n\geq 1} be the Picard iteration sequence in T\mathcal{F}_{T}. By construction,

(v(n),b(n))T2(etΔv0,etΔb0)T.\left\|(v^{(n)},b^{(n)})\right\|_{{\mathcal{F}}_{T}}\leq 2\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{\mathcal{F}_{T}}. (2.31)

Note that

(v(n),b(n))T(etΔv0,etΔb0)T+B((v(n1),b(n1)),v(n1),b(n1)))T.\left\|(v^{(n)},b^{(n)})\right\|_{\mathcal{E}_{T}}\leq\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{\mathcal{E}_{T}}+\left\|B((v^{(n-1)},b^{(n-1)}),v^{(n-1)},b^{(n-1)}))\right\|_{\mathcal{E}_{T}}.

We now bound B((v,b),(u,a))B((v,b),(u,a)) in T\mathcal{E}_{T} in terms of (v,b)(v,b) and (u,a)(u,a) in T\mathcal{F}_{T} and T\mathcal{E}_{T}. We have by [1, Lemma 2.1] and Hölder inequality (1.3) using q16q_{1}\geq 6,

B1((v,b),(u,a))Eq3(t)0t1(tτ)12(vu(τ)Eq3+ba(τ)Eq3)𝑑τ0t1(tτ)12(v(τ)Eq16u(τ)Eq16+b(τ)Eq16a(τ)Eq16)𝑑τ0t1(tτ)12τ1/4(v,b)Tτ1/4(u,a)T𝑑τ(v,b)T(u,a)T,\begin{split}\left\|B_{1}((v,b),(u,a))\right\|_{E^{3}_{q}}(t)&{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\left(\left\|v\otimes u(\tau)\right\|_{E^{3}_{q}}+\left\|b\otimes a(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\left(\left\|v(\tau)\right\|_{E^{6}_{q_{1}}}\left\|u(\tau)\right\|_{E^{6}_{q_{1}}}+\left\|b(\tau)\right\|_{E^{6}_{q_{1}}}\left\|a(\tau)\right\|_{E^{6}_{q_{1}}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\tau^{-1/4}\left\|(v,b)\right\|_{\mathcal{F}_{T}}\tau^{-1/4}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\,d\tau\\ &{\ \lesssim\ }\left\|(v,b)\right\|_{\mathcal{F}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}},\end{split} (2.32)
B2((v,b),(u,a))Eq3(t)0t1(tτ)12(va(τ)Eq3+bu(τ)Eq3)𝑑τ0t1(tτ)12(v(τ)Eq16a(τ)Eq16+b(τ)Eq16u(τ)Eq16)𝑑τ0t1(tτ)12τ1/4(v,b)Tτ1/4(u,a)T𝑑τ(v,b)T(u,a)T,\begin{split}\left\|B_{2}((v,b),(u,a))\right\|_{E^{3}_{q}}(t)&{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\left(\left\|v\otimes a(\tau)\right\|_{E^{3}_{q}}+\left\|b\otimes u(\tau)\right\|_{E^{3}_{q}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\left(\left\|v(\tau)\right\|_{E^{6}_{q_{1}}}\left\|a(\tau)\right\|_{E^{6}_{q_{1}}}+\left\|b(\tau)\right\|_{E^{6}_{q_{1}}}\left\|u(\tau)\right\|_{E^{6}_{q_{1}}}\right)d\tau\\ &\leq\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{1}{2}}}\tau^{-1/4}\left\|(v,b)\right\|_{\mathcal{F}_{T}}\tau^{-1/4}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\,d\tau\\ &{\ \lesssim\ }\left\|(v,b)\right\|_{\mathcal{F}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}},\end{split} (2.33)

so that

B((v,b),(u,a))Eq3×Eq3(t)(v,b)T(u,a)T.\left\|B((v,b),(u,a))\right\|_{E^{3}_{q}\times E^{3}_{q}}(t){\ \lesssim\ }\left\|(v,b)\right\|_{\mathcal{F}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}. (2.34)

By TT\mathcal{E}_{T}\subset\mathcal{F}_{T}, we have

B((v,b),(u,a))Eq3×Eq3(t)(v,b)T(u,a)T(u,a)T(v,b)T.\left\|B((v,b),(u,a))\right\|_{E^{3}_{q}\times E^{3}_{q}}(t){\ \lesssim\ }\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\wedge\left\|(u,a)\right\|_{\mathcal{E}_{T}}\left\|(v,b)\right\|_{\mathcal{F}_{T}}.

Also by [1, Lemma 2.1] and Hölder inequality (1.3),

B1((v,b),(u,a))Eq2(t)\displaystyle\|B_{1}((v,b),(u,a))\|_{E^{\infty}_{q_{2}}}(t) 0t1(tτ)34(vuEq16(τ)+baEq16(τ))𝑑τ\displaystyle{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\|v\otimes u\|_{E^{6}_{q_{1}}}(\tau)+\|b\otimes a\|_{E^{6}_{q_{1}}}(\tau)\right)d\tau
0t1(tτ)34τ3/4(τ1/2vEq2τ1/4uEq16τ1/2vEq2τ1/4uEq16\displaystyle{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}\tau^{3/4}}(\tau^{1/2}\|v\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|u\|_{E^{6}_{q_{1}}}\wedge\tau^{1/2}\|v\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|u\|_{E^{6}_{q_{1}}}
+τ1/2bEq2τ1/4aEq16τ1/2bEq2τ1/4aEq16)dτ\displaystyle\qquad\qquad\qquad\qquad\quad+\tau^{1/2}\|b\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|a\|_{E^{6}_{q_{1}}}\wedge\tau^{1/2}\|b\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|a\|_{E^{6}_{q_{1}}})\,d\tau
t12((v,b)T(u,a)T(u,a)T(v,b)T),\displaystyle{\ \lesssim\ }t^{-\frac{1}{2}}(\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\wedge\left\|(u,a)\right\|_{\mathcal{E}_{T}}\left\|(v,b)\right\|_{\mathcal{F}_{T}}), (2.35)
B2((v,b),(u,a))Eq2(t)\displaystyle\|B_{2}((v,b),(u,a))\|_{E^{\infty}_{q_{2}}}(t) 0t1(tτ)34(vaEq16(τ)+buEq16(τ))𝑑τ\displaystyle{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}}\left(\|v\otimes a\|_{E^{6}_{q_{1}}}(\tau)+\|b\otimes u\|_{E^{6}_{q_{1}}}(\tau)\right)d\tau
0t1(tτ)34τ3/4(τ1/2vEq2τ1/4aEq16τ1/2vEq2τ1/4aEq16\displaystyle{\ \lesssim\ }\int_{0}^{t}\frac{1}{(t-\tau)^{\frac{3}{4}}\tau^{3/4}}(\tau^{1/2}\|v\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|a\|_{E^{6}_{q_{1}}}\wedge\tau^{1/2}\|v\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|a\|_{E^{6}_{q_{1}}}
+τ1/2bEq2τ1/4uEq16τ1/2bEq2τ1/4uEq16)dτ\displaystyle\qquad\qquad\qquad\qquad\quad+\tau^{1/2}\|b\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|u\|_{E^{6}_{q_{1}}}\wedge\tau^{1/2}\|b\|_{E^{\infty}_{q_{2}}}\tau^{1/4}\|u\|_{E^{6}_{q_{1}}})\,d\tau
t12((v,b)T(u,a)T(u,a)T(v,b)T),\displaystyle{\ \lesssim\ }t^{-\frac{1}{2}}(\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\wedge\left\|(u,a)\right\|_{\mathcal{E}_{T}}\left\|(v,b)\right\|_{\mathcal{F}_{T}}), (2.36)

so that

B((v,b),(u,a))Eq2×Eq2(t)t12((v,b)T(u,a)T(u,a)T(v,b)T).\displaystyle\|B((v,b),(u,a))\|_{E^{\infty}_{q_{2}}\times E^{\infty}_{q_{2}}}(t){\ \lesssim\ }t^{-\frac{1}{2}}(\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\wedge\left\|(u,a)\right\|_{\mathcal{E}_{T}}\left\|(v,b)\right\|_{\mathcal{F}_{T}}).

Based on the above estimates we conclude

B((v,b),(u,a))T(v,b)T(u,a)T(u,a)T(v,b)T.\|B((v,b),(u,a))\|_{\mathcal{E}_{T}}{\ \lesssim\ }\left\|(v,b)\right\|_{\mathcal{E}_{T}}\left\|(u,a)\right\|_{\mathcal{F}_{T}}\wedge\left\|(u,a)\right\|_{\mathcal{E}_{T}}\left\|(v,b)\right\|_{\mathcal{F}_{T}}. (2.37)

We can now conclude that {(v(n),b(n))}\{(v^{(n)},b^{(n)})\} is Cauchy in T\mathcal{E}_{T} by the calculation preceding and including (LABEL:ineq:difference.picard.iterates). However, the smallness of the constant is now provided by (2.30)-(2.31), not by (v0,b0)Eq3×Eq3\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}.

We now show continuity. For small data, we can try to inherit continuity from Theorem 1.2. But we will provide a proof valid for general data. We first address convergence to the initial data. By [1, Lemma 2.3] we have

limT0+sup0<t<Tt14(etΔv0,etΔb0)Eq16×Eq16=limT0(etΔv0,etΔb0)T=0,\lim_{T^{\prime}\to 0^{+}}\sup_{0<t<T^{\prime}}t^{\frac{1}{4}}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{E^{6}_{q_{1}}\times E^{6}_{q_{1}}}=\lim_{T^{\prime}\to 0}\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{{\mathcal{F}}_{T^{\prime}}}=0, (2.38)

whenever (v0,b0)Eq3×Eq3(v_{0},b_{0})\in E^{3}_{q}\times E^{3}_{q}. By our estimates in the class T{\mathcal{F}}_{T^{\prime}} where we are taking TTT^{\prime}\leq T, we have

(v(n),b(n))T(etΔv0,etΔb0)T+B((v(n1),b(n1)),(v(n1),b(n1)))T(etΔv0,etΔb0)T+(v(n1),b(n1))T2.\begin{split}\|(v^{(n)},b^{(n)})\|_{{\mathcal{F}}_{T^{\prime}}}&\leq\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{{\mathcal{F}}_{T^{\prime}}}+\|B((v^{(n-1)},b^{(n-1)}),(v^{(n-1)},b^{(n-1)}))\|_{{\mathcal{F}}_{T^{\prime}}}\\ &{\ \lesssim\ }\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\|_{{\mathcal{F}}_{T^{\prime}}}+\|(v^{(n-1)},b^{(n-1)})\|_{{\mathcal{F}}_{T^{\prime}}}^{2}.\end{split}

From this and by induction, for any nn we have

limT0+(v(n),b(n))T=0.\begin{split}\lim_{T^{\prime}\to 0^{+}}\|(v^{(n)},b^{(n)})\|_{{\mathcal{F}}_{T^{\prime}}}=0.\end{split}

The limit (2.38), convergence of the Picard iterates in T{\mathcal{F}}_{T} and the above inequality imply that, by taking TT^{\prime} small, we can make sup0<t<Tt14(v,b)(t)Eq16×Eq16\sup_{0<t<T^{\prime}}t^{\frac{1}{4}}\|(v,b)(t)\|_{E^{6}_{q_{1}}\times E^{6}_{q_{1}}} small. To elaborate, we have

(v,b)T(v,b)(v(n),b(n))T+(v(n),b(n))T.\|(v,b)\|_{{\mathcal{F}}_{T^{\prime}}}\leq\|(v,b)-(v^{(n)},b^{(n)})\|_{{\mathcal{F}}_{T^{\prime}}}+\|(v^{(n)},b^{(n)})\|_{{\mathcal{F}}_{T^{\prime}}}. (2.39)

We may choose nn large so that the first term is small and then make the second term small by taking TT^{\prime} small. Hence,

limT0+(v,b)T=0.\lim_{T^{\prime}\to 0^{+}}{\|(v,b)\|_{{\mathcal{F}}_{T^{\prime}}}}=0. (2.40)

Using (2.34), this implies

limT0+sup0<t<TB((v,b),(v,b))Eq3×v(t)=0.\begin{split}\lim_{T^{\prime}\to 0^{+}}\sup_{0<t<T^{\prime}}\|B((v,b),(v,b))\|_{E^{3}_{q}\times v}(t)=0.\end{split}

This and [1, Lemma 2.3] imply

limt0(v,b)(v0,b0)Eq3×Eq3=0.\lim_{t\to 0}\|(v,b)-(v_{0},b_{0})\|_{E^{3}_{q}\times E^{3}_{q}}=0.

The proof of continuity for positive times follows from the same argument used in the proof of Theorem 1.2 and is therefore omitted for brevity.

We now prove the spacetime integral bound (1.24) for p(3,9]p\in(3,9] and 2s+3p=1\frac{2}{s}+\frac{3}{p}=1. Note that we exclude p=3p=3, i.e., s=s=\infty. By imbedding EqpEmpE^{p}_{q}\subset E^{p}_{m} for m>qm>q, we may assume m<m<\infty. (We do not take m=qm=q since we need q<mq<m for global existence). Denote the Banach space

XT=T(ET,ms,p×ET,ms,p).X_{T}=\mathcal{E}_{T}\cap(E^{s,p}_{T,m}\times E^{s,p}_{T,m}).

For the linear term, by [1, Lemma 2.1] and [1, Lemma 2.4],

(etΔv0,etΔb0)XT=sup0<t<T(etΔv0,etΔb0)Eq3×Eq3+sup0<t<Tt12(etΔv0,etΔb0)Eq2×Eq2+(etΔv0,etΔb0)ET,ms,p×ET,ms,pC3(1+Tβ0)(v0,b0)Eq3×Eq3,\begin{split}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{X_{T}}&=\sup_{0<t<T}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}}\\ &\quad+\sup_{0<t<T}t^{\frac{1}{2}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{\infty}_{q_{2}}\times E^{\infty}_{q_{2}}}+\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}\\ &\leq C_{3}(1+T^{\beta_{0}})\left\|(v_{0},b_{0})\right\|_{E^{3}_{q}\times E^{3}_{q}},\end{split} (2.41)

for any β0[0,)\beta_{0}\in[0,\infty) and β0>α0=32m32q+1s\beta_{0}>\alpha_{0}=\frac{3}{2m}-\frac{3}{2q}+\frac{1}{s}. Note that

limT0+(etΔv0,etΔb0)ET,ms,p×ET,ms,p=0.\lim_{T\to 0_{+}}\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}=0. (2.42)

For the bilinear term, by [1, Lemma 2.7] with s~=s/2\tilde{s}=s/2, p~=p/2\tilde{p}=p/2, and m~=max(1,m/2)\tilde{m}=\max(1,m/2), so that

σ=0,α={1232m1s,if2m,1+32m1s,if1<m<2,α<11s,\sigma=0,\quad\alpha=\left\{\begin{aligned} \tfrac{1}{2}-\tfrac{3}{2m}-\tfrac{1}{s},\quad&\text{if}\quad 2\leq m\leq\infty,\\[2.84526pt] -1+\tfrac{3}{2m}-\tfrac{1}{s},\quad&\text{if}\quad 1<m<2,\end{aligned}\right.\quad\alpha<1-\tfrac{1}{s}, (2.43)

we have

B((v,b),(u,a))ET,ms,p×ET,ms,pC4(1+Tβ)(vuET,m~s2,p2+baET,m~s2,p2+vaET,m~s2,p2+buET,m~s2,p2)\begin{split}\left\|B((v,b),(u,a))\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}&\leq C_{4}(1+T^{\beta})\left(\left\|v\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,\tilde{m}}}+\left\|b\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,\tilde{m}}}+\left\|v\otimes a\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,\tilde{m}}}+\left\|b\otimes u\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,\tilde{m}}}\right)\end{split}

for any β[0,11s]\beta\in[0,1-\frac{1}{s}] and β>α\beta>\alpha. Note

fgET,m~s2,p2fET,2m~s,pgET,2m~s,pfET,ms,pgET,ms,p\left\|f\otimes g\right\|_{E^{\frac{s}{2},\frac{p}{2}}_{T,\tilde{m}}}\leq\left\|f\right\|_{E^{s,p}_{T,2\tilde{m}}}\left\|g\right\|_{E^{s,p}_{T,2\tilde{m}}}\leq\left\|f\right\|_{E^{s,p}_{T,m}}\left\|g\right\|_{E^{s,p}_{T,m}}

no matter m2m\geq 2 or 1<m<21<m<2. We conclude, also using (2.37),

B((v,b),(u,a))ET,ms,p×ET,ms,pC4(1+Tβ)(v,b)ET,ms,p×ET,ms,p(u,a)ET,ms,p×ET,ms,p,B((v,b),(u,a))XT2C4(1+Tβ)(v,b)XT(u,a)XT.\begin{split}\left\|B((v,b),(u,a))\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}&\leq C_{4}(1+T^{\beta})\left\|(v,b)\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}}\left\|(u,a)\right\|_{E^{s,p}_{T,m}\times E^{s,p}_{T,m}},\\ \left\|B((v,b),(u,a))\right\|_{X_{T}}&\leq 2C_{4}(1+T^{\beta})\left\|(v,b)\right\|_{X_{T}}\left\|(u,a)\right\|_{X_{T}}.\end{split} (2.44)

By (2.42), we can find T1(0,T]T_{1}\in(0,T] so that

(etΔv0,etΔb0)ET1,ms,p×ET1,ms,pδ=[4C4(1+Tβ)]1.\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{E^{s,p}_{T_{1},m}\times E^{s,p}_{T_{1},m}}\leq\delta=[4C_{4}(1+T^{\beta})]^{-1}.

Then the Picard sequence (v(k),b(k))(v^{(k)},b^{(k)}) satisfies (v(k),b(k))ET1,ms,p×ET1,ms,p2δ\left\|(v^{(k)},b^{(k)})\right\|_{E^{s,p}_{T_{1},m}\times E^{s,p}_{T_{1},m}}\leq 2\delta for all kk\in\mathbb{N}, and we get (v,b)ET1,ms,p×ET1,ms,p2δ\left\|(v,b)\right\|_{E^{s,p}_{T_{1},m}\times E^{s,p}_{T_{1},m}}\leq 2\delta. Thus, (v,b)(v,b) satisfies the spacetime integral bound (1.24).

We now establish the global ET,ms,pE^{s,p}_{T,m}-estimates when (v0,b0)(v_{0},b_{0}) is sufficiently small in Eq3×Eq3E^{3}_{q}\times E^{3}_{q}. To this end, we aim to eliminate the dependence of constants on TT, i.e., we choose β0=0\beta_{0}=0 in (2.41) and β=0\beta=0 in (2.44).

Let us analyze the conditions under which β=0\beta=0 in (2.44) is permissible. When m2m\geq 2, we additionally assume that 2s+3m1\frac{2}{s}+\frac{3}{m}\geq 1 so that the exponent α0\alpha\leq 0. In particular, we can take β=0\beta=0 when α=0\alpha=0 since the condition 1<m~<m<1<\tilde{m}<m<\infty required in [1, Lemma 2.7] is satisfied. Note that when m=2m=2, we have m~=1\tilde{m}=1, but then α<0\alpha<0.

For 1<m<21<m<2, we impose the additional condition 3m<2s+2=33p\frac{3}{m}<\frac{2}{s}+2=3-\frac{3}{p}, which ensures α<0\alpha<0 and hence again allows us to take β=0\beta=0.

With β0=β=0\beta_{0}=\beta=0 in (2.41) and (2.44), we obtain a global-in-time estimate for (v,b)(v,b) in ET=,ms,p×ET=,ms,pE^{s,p}_{T=\infty,m}\times E^{s,p}_{T=\infty,m}, provided that the initial data satisfies (v0,b0)Eqp×Eqp<[32C3C4]1\left\|(v_{0},b_{0})\right\|_{E^{p}_{q}\times E^{p}_{q}}<[32C_{3}C_{4}]^{-1}.

Observe that when m2m\geq 2, all required conditions–including the upper bound 2s+3m1\frac{2}{s}+\frac{3}{m}\geq 1– are satisfied if we take m=pm=p. Once we have established (v,b)ET=,ms,p×ET=,ms,p(v,b)\in E^{s,p}_{T=\infty,m}\times E^{s,p}_{T=\infty,m} for some mm, then the inclusion property implies (v,b)ET=,m~s,p×ET=,m~s,p(v,b)\in E^{s,p}_{T=\infty,\tilde{m}}\times E^{s,p}_{T=\infty,\tilde{m}} for all m~[m,]\tilde{m}\in[m,\infty]. Thus, the condition 2s+3m1\frac{2}{s}+\frac{3}{m}\geq 1 can be removed entirely.

We now consider the LTsEmpL^{s}_{T}E^{p}_{m}-estimates of (v,b)(v,b), restricting to T=T=\infty for simplicity. Fix s[3,)s\in[3,\infty) and q[1,3]q\in[1,3], and define pp by the relation 3p+2s=1\frac{3}{p}+\frac{2}{s}=1. Using [1, Lemma 2.8] with p~=p/2\tilde{p}=p/2, s~=s/2\tilde{s}=s/2 and m~1\tilde{m}\geq 1 such that 1m~1m=1p~1p=1p\frac{1}{\tilde{m}}-\frac{1}{m}=\frac{1}{\tilde{p}}-\frac{1}{p}=\frac{1}{p}, and by applying Hölder inequality, we obtain

B((v,b),(u,a))LTs(Emp×Emp)vuLTs2Em~p2+baLTs2Em~p2+vaLTs2Em~p2+buLTs2Em~p2(v,b)LTs(Emp×Emp)(u,a)LTs(Epp×Epp)(v,b)LTs(Emp×Emp)(u,a)LTs(Epp×Epp), if pm.\begin{split}\left\|B((v,b),(u,a))\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}&{\ \lesssim\ }\left\|v\otimes u\right\|_{L^{\frac{s}{2}}_{T}E^{\frac{p}{2}}_{\tilde{m}}}+\left\|b\otimes a\right\|_{L^{\frac{s}{2}}_{T}E^{\frac{p}{2}}_{\tilde{m}}}+\left\|v\otimes a\right\|_{L^{\frac{s}{2}}_{T}E^{\frac{p}{2}}_{\tilde{m}}}+\left\|b\otimes u\right\|_{L^{\frac{s}{2}}_{T}E^{\frac{p}{2}}_{\tilde{m}}}\\ &{\ \lesssim\ }\left\|(v,b)\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}\left\|(u,a)\right\|_{L^{s}_{T}(E^{p}_{p}\times E^{p}_{p})}\\ &{\ \lesssim\ }\left\|(v,b)\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}\left\|(u,a)\right\|_{L^{s}_{T}(E^{p}_{p}\times E^{p}_{p})},\ \text{ if }p\geq m.\end{split}

The condition m~1\tilde{m}\geq 1 is equivalent to mp=pp1m\geq p^{\prime}{=\frac{p}{p-1}}. Hence,

B((v,b),(u,a))LTs(Emp×Emp)(v,b)LTs(Emp×Emp)(u,a)LTs(Emp×Emp), if pmp.\left\|B((v,b),(u,a))\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}{\ \lesssim\ }\left\|(v,b)\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})}\left\|(u,a)\right\|_{L^{s}_{T}(E^{p}_{m}\times E^{p}_{m})},\ \text{ if }p^{\prime}\leq m\leq p. (2.45)

Let (s,q)\mathcal{M}(s,q) denote the set of all mm for which we can establish (v,b)Ls(Emp×Emp)(v,b)\in L^{s}(E^{p}_{m}\times E^{p}_{m}). Since LsEmpLsEm2pL^{s}E^{p}_{m}\subset L^{s}E^{p}_{m_{2}} for m<m2m<m_{2}, the set (s,q)\mathcal{M}(s,q), if nonempty, must be an interval of the form

m¯<m,orm¯m,\underline{m}<m\leq\infty,\quad\text{or}\quad\underline{m}\leq m\leq\infty, (2.46)

for some m¯=m¯(s,q)[1,]\underline{m}=\underline{m}(s,q)\in[1,\infty].

Define m1m_{1} by

1m1=1q23s.\frac{1}{m_{1}}=\frac{1}{q}-\frac{2}{3s}.

Given s<s<\infty and q3q\leq 3, and recalling that 1p=1323s\frac{1}{p}=\frac{1}{3}-\frac{2}{3s}, we deduce that

q<m1p.q<m_{1}\leq p. (2.47)

From [1, Lemma 2.4] with d=r=3d=r=3, we have

etΔfLT=sEmpfEq3,\left\|e^{t\Delta}f\right\|_{L^{s}_{T=\infty}E^{p}_{m}}\lesssim\|f\|_{E^{3}_{q}}, (2.48)

with constants independent of RR, provided one of the following holds:

  1. A.

    m1msm_{1}\leq m\leq s, (and m1<msm_{1}<m\leq s if q=1q=1),

  2. E1.

    s<p=ms<p=m,

  3. E2.

    s<ms<m and 1q53s\frac{1}{q}\geq\frac{5}{3s}.

If the linear estimate (2.48) holds, and if pmpp^{\prime}\leq m\leq p so that the bilinear estimate (2.45) holds, then the Picard iteration (v(n),b(n))(v^{(n)},b^{(n)}) converges in Ls(Emp×Emp)L^{s}(E^{p}_{m}\times E^{p}_{m}) for sufficiently small initial data.

Case A: This case applies as soon as m1sm_{1}\leq s, i.e., 1q53s\frac{1}{q}\geq\frac{5}{3s}. Strict inequality m1<sm_{1}<s holds when q=1q=1. By (2.47) and since p<2<sp^{\prime}<2<s, the value

m(s,q)=max(p,m1)m^{*}(s,q)=\max(p^{\prime},m_{1})

belongs to both [p,p][p^{\prime},p] and [m1,s][m_{1},s]. Thus, (v(n),b(n))(v^{(n)},b^{(n)}) converges in Ls(Emp×Emp)L^{s}(E^{p}_{m}\times E^{p}_{m}) for m=mm=m^{*}, or for mm slightly larger than mm^{*} when q=1q=1 and 3s43\leq s\leq 4.

Case E1: This case applies when 3s<53\leq s<5, allowing us to take m=pm=p. It thus covers the parameter range 1q3s<51\leq q\leq 3\leq s<5, and 1q<53s\frac{1}{q}<\frac{5}{3s}.

Case E2: While this case also requires 1q53s\frac{1}{q}\geq\frac{5}{3s}, it does not yield smaller admissible mm than Case A.

This completes the proof of LsEmpL^{s}E^{p}_{m}-estimates, and concludes the proof of Theorem 1.3. ∎

3 Local energy solutions in Wiener amalgam spaces

In this section, we address weak solutions and establish Theorems 1.8, 1.9, 1.10 for the MHD equations (MHD), along with Theorems 1.12, 1.13, 1.14 for the viscoelastic Navier–Stokes equations with damping (vNSEd). Given the similarity between the structures of (vNSEd) and (MHD), we focus on presenting the proofs of Theorems 1.8, 1.9, 1.10 for (MHD). The details of verification of Theorems 1.12, 1.13, 1.14 for (vNSEd) are left to the readers.

Define

NR0(v0,b0)=supx031RBR(x0)(|v0|2+|b0|2)𝑑x,N_{R}^{0}(v_{0},b_{0})=\sup_{x_{0}\in\mathbb{R}^{3}}\frac{1}{R}\int_{B_{R}(x_{0})}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx,

and

Nq,R0(v0,b0)=1R[k3(BR(kR)(|v0|2+|b0|2)𝑑x)q/2]2/q,N,R0(v0,b0)=NR0(v0,b0):=supx031RBR(x0)(|v0|2+|b0|2)𝑑x.\begin{split}N_{q,R}^{0}(v_{0},b_{0})&=\frac{1}{R}\left[\sum_{k\in\mathbb{Z}^{3}}\left(\int_{B_{R}(kR)}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx\right)^{q/2}\right]^{2/q},\\ N_{\infty,R}^{0}(v_{0},b_{0})&=N_{R}^{0}(v_{0},b_{0}):=\sup_{x_{0}\in\mathbb{R}^{3}}\frac{1}{R}\int_{B_{R}(x_{0})}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx.\end{split}
Lemma 3.1.

Let v0,b0Luloc2v_{0},b_{0}\in L^{2}_{\mathrm{uloc}} be divergence free, and assume (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}). For all r>0r>0 we have

esssup0tσr2supx03Br(x0)|v|2+|b|22𝑑x𝑑t+supx030σr2Br(x0)(|v|2+|b|2)𝑑x𝑑t<CA0(r),\mathop{\rm ess\,sup}_{0\leq t\leq\sigma r^{2}}\sup_{x_{0}\in\mathbb{R}^{3}}\int_{B_{r}(x_{0})}\frac{|v|^{2}+|b|^{2}}{2}\,dxdt+\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{\sigma r^{2}}\int_{B_{r}(x_{0})}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt<CA_{0}(r), (3.1)
supx030σr2Br(x0)(|v|3+|b|3+|pcx0,r(t)|3/2)𝑑x𝑑t<Cr12(A0(r))32,\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{\sigma r^{2}}\int_{B_{r}(x_{0})}\left(|v|^{3}+|b|^{3}+|p-c_{x_{0},r}(t)|^{3/2}\right)dxdt<Cr^{\frac{1}{2}}(A_{0}(r))^{\frac{3}{2}}, (3.2)

where

A0(r)=rNr0(v0,b0)=supx03Br(x0)(|v0|2+|b0|2)𝑑x,A_{0}(r)=rN_{r}^{0}(v_{0},b_{0})=\sup_{x_{0}\in\mathbb{R}^{3}}\int_{B_{r}(x_{0})}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx,

and

σ=σ(r)=c0min{(Nr0(v0,b0))2,1},\sigma=\sigma(r)=c_{0}\min\left\{(N_{r}^{0}(v_{0},b_{0}))^{-2},1\right\}, (3.3)

for a small universal constant c0>0c_{0}>0.

Proof.

The proof of the lemma is an adaption of the proof of [3, Lemma 2.1] for the Navier–Stokes equations to the MHD equations.

By Hölder and Young inequalities, for any δ>0\delta>0, we have

vL3(0,T;L3)3vL6(0,T;L2)3/2vL2(0,T;L6)3/2(δR)3vL6(0,T;L2)6+δRvL2(0,T;L6)2,\left\|v\right\|_{L^{3}(0,T;L^{3})}^{3}{\ \lesssim\ }\left\|v\right\|_{L^{6}(0,T;L^{2})}^{3/2}\left\|v\right\|^{3/2}_{L^{2}(0,T;L^{6})}{\ \lesssim\ }(\delta R)^{-3}\left\|v\right\|_{L^{6}(0,T;L^{2})}^{6}+\delta R\left\|v\right\|_{L^{2}(0,T;L^{6})}^{2},

and

bL3(0,T;L3)3bL6(0,T;L2)3/2bL2(0,T;L6)3/2(δR)3bL6(0,T;L2)6+δRbL2(0,T;L6)2.\left\|b\right\|_{L^{3}(0,T;L^{3})}^{3}{\ \lesssim\ }\left\|b\right\|_{L^{6}(0,T;L^{2})}^{3/2}\left\|b\right\|^{3/2}_{L^{2}(0,T;L^{6})}{\ \lesssim\ }(\delta R)^{-3}\left\|b\right\|_{L^{6}(0,T;L^{2})}^{6}+\delta R\left\|b\right\|_{L^{2}(0,T;L^{6})}^{2}.

In addition, applying Sobolev inequality yields

1R0σR2B2R(x0)|v|3𝑑x𝑑tCδ3R40σR2(B2R(x0)|v|2𝑑x)3𝑑t+CδR20σR2B2R(x0)|v|2𝑑x𝑑t+Cδsupx030σR2B2R(x0)|v|2𝑑x𝑑t,\begin{split}\frac{1}{R}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|v|^{3}\,dxdt&\leq\frac{C}{\delta^{3}R^{4}}\int_{0}^{\sigma R^{2}}\left(\int_{B_{2R}(x_{0})}|v|^{2}\,dx\right)^{3}dt+\frac{C\delta}{R^{2}}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|v|^{2}\,dxdt\\ &\quad+C\delta\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|\nabla v|^{2}\,dxdt,\end{split}

and

1R0σR2B2R(x0)|b|3𝑑x𝑑tCδ3R40σR2(B2R(x0)|b|2𝑑x)3𝑑t+CδR20σR2B2R(x0)|b|2𝑑x𝑑t+Cδsupx030σR2B2R(x0)|b|2𝑑x𝑑t,\begin{split}\frac{1}{R}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|b|^{3}\,dxdt&\leq\frac{C}{\delta^{3}R^{4}}\int_{0}^{\sigma R^{2}}\left(\int_{B_{2R}(x_{0})}|b|^{2}\,dx\right)^{3}dt+\frac{C\delta}{R^{2}}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|b|^{2}\,dxdt\\ &\quad+C\delta\sup_{x_{0}\in\mathbb{R}^{3}}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|\nabla b|^{2}\,dxdt,\end{split}

where CC is a positive constant independent of σ\sigma. For the pressure term, using the local pressure expansion (1.27), we obtain

1R0σR2B2R(x0)|πcx0,R(t)|3/2CR0σR2B4R(x0)(|v|3+|b|3)𝑑x𝑑t+CR40σR2[A¯(tR2)]3/2𝑑t,\frac{1}{R}\int_{0}^{\sigma R^{2}}\int_{B_{2R}(x_{0})}|\pi-c_{x_{0},R}(t)|^{3/2}\leq\frac{C}{R}\int_{0}^{\sigma R^{2}}\int_{B_{4R}(x_{0})}\left(|v|^{3}+|b|^{3}\right)dxdt+\frac{C}{R^{4}}\int_{0}^{\sigma R^{2}}\left[\overline{A}\left(\frac{t}{R^{2}}\right)\right]^{3/2}dt, (3.4)

where

A¯(σ)=esssup0tσR2supx033|v|2+|b|22ϕ(xx0)𝑑x.\overline{A}(\sigma)=\mathop{\rm ess\,sup}_{0\leq t\leq\sigma R^{2}}\sup_{x_{0}\in\mathbb{R}^{3}}\int_{\mathbb{R}^{3}}\frac{|v|^{2}+|b|^{2}}{2}\,\phi(x-x_{0})\,dx.

Applying the local energy inequality, we deduce

3|v|2+|b|22ϕ(xx0)𝑑x+0t3(|v|2+|b|2)ϕ(xx0)𝑑x𝑑sA0(R)2+CR20σR2A¯(sR2)𝑑s+CR40σR2(A¯(sR2))3𝑑s,\begin{split}\int_{\mathbb{R}^{3}}\frac{|v|^{2}+|b|^{2}}{2}\phi(x-x_{0})\,dx&+\int_{0}^{t}\int_{\mathbb{R}^{3}}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi(x-x_{0})\,dxds\\ &\leq\frac{A_{0}(R)}{2}+\frac{C}{R^{2}}\int_{0}^{\sigma R^{2}}\overline{A}\left(\frac{s}{R^{2}}\right)ds+\frac{C}{R^{4}}\int_{0}^{\sigma R^{2}}\left(\overline{A}\left(\frac{s}{R^{2}}\right)\right)^{3}ds,\end{split} (3.5)

for sufficiently small δ\delta. Therefore, we conclude that

A¯(σ)RA0(R)2R+CR20σR2A¯(sR2)R𝑑s+CR20σR2(A¯(sR2)R)2𝑑s.\frac{\overline{A}(\sigma)}{R}\leq\frac{A_{0}(R)}{2R}+\frac{C}{R^{2}}\int_{0}^{\sigma R^{2}}\frac{\overline{A}\left(\frac{s}{R^{2}}\right)}{R}\,ds+\frac{C}{R^{2}}\int_{0}^{\sigma R^{2}}\left(\frac{\overline{A}\left(\frac{s}{R^{2}}\right)}{R}\right)^{2}ds. (3.6)

Making the change of variables τ=s/R2\tau=s/R^{2} and applying Grönwall’s inequality [3, Lemma 2.2], we derive

A¯(σ)A0(R),\overline{A}(\sigma)\leq A_{0}(R),

for t[0,TR]t\in[0,T_{R}] where TR=σR2T_{R}=\sigma R^{2}, and

σ=c0min{(NR0(v0,b0))2,1}\sigma=c_{0}\min\left\{(N_{R}^{0}(v_{0},b_{0}))^{-2},1\right\}

for some small constant c0c_{0} independent of RR and (v0,b0)(v_{0},b_{0}). Note that A¯(σ)\overline{A}(\sigma) is nondecreasing in σ\sigma, and its continuity in σ\sigma follows directly from the local energy inequality. With the above estimate for A¯(σ)\overline{A}(\sigma), the lemma follows by the standard continuation in σ\sigma argument, provided that c0c_{0} is chosen sufficiently small. ∎

We recall a local regularity criterion for suitable weak solutions that is a replacement for the case of the Navier–Stokes equations proposed in [6] (see also [28, 36]). See [15] for the definition of suitable weak solutions.

Lemma 3.2 (ϵ\epsilon-regularity criterion [34, Theorem 3.1]).

There exists a universal constant ϵ>0\epsilon_{*}>0 such that, if (v,b,π)(v,b,\pi) is a suitable weak solution of (MHD) in Qr=Qr(x0,t0)=Br(x0)×(t0r2,t0)Q_{r}=Q_{r}(x_{0},t_{0})=B_{r}(x_{0})\times(t_{0}-r^{2},t_{0}), Br(x0)3B_{r}(x_{0})\subset\mathbb{R}^{3}, and

ϵ3=1r2Qr(|v|3+|v|3+|π|3/2)𝑑x𝑑t<ϵ,\epsilon^{3}=\frac{1}{r^{2}}\int_{Q_{r}}\left(|v|^{3}+|v|^{3}+|\pi|^{3/2}\right)dxdt<\epsilon_{*},

then vv and bb are Hölder continuous on Qr/2¯\overline{Q_{r/2}}.

The corresponding ϵ\epsilon-regularity criterion for weak solutions of the viscoelastic Navier–Stokes equations with damping, (vNSEd), is established in [17, Proposition 3.2].

Theorem 3.3 (Initial and eventual regularity).

There is a small positive constant ϵ1\epsilon_{1} such that the following holds. Assume that v0,b0Luloc2(3)v_{0},b_{0}\in L^{2}_{\mathrm{uloc}}(\mathbb{R}^{3}) are divergence free and that (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}). Let

NR0:=supx031RBR(x0)(|v0|2+|b0|2)𝑑x.N_{R}^{0}:=\sup_{x_{0}\in\mathbb{R}^{3}}\frac{1}{R}\int_{B_{R}(x_{0})}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx.

1. If there exists R0>0R_{0}>0 so that

supRR0NR0<ϵ1,\sup_{R\geq R_{0}}N_{R}^{0}<\epsilon_{1},

then (v,b)(v,b) has eventual regularity. Moreover, if R02tR_{0}^{2}{\ \lesssim\ }t, then

t1/2(v,b)L×L(supRR0NR0)1/2<.t^{1/2}\left\|(v,b)\right\|_{L^{\infty}\times L^{\infty}}{\ \lesssim\ }\left(\sup_{R\geq R_{0}}N_{R}^{0}\right)^{1/2}<\infty.

2. If there exists R0>0R_{0}>0 so that

supRR0NR0<ϵ1,\sup_{R\leq R_{0}}N_{R}^{0}<\epsilon_{1},

then (v,b)(v,b) has initial regularity. Moreover, if tc0R02t\leq c_{0}R_{0}^{2}, then

t1/2(v,b)(,t)L×L(supRR0NR0)1/2<.t^{1/2}\left\|(v,b)(\cdot,t)\right\|_{L^{\infty}\times L^{\infty}}{\ \lesssim\ }\left(\sup_{R\leq R_{0}}N_{R}^{0}\right)^{1/2}<\infty.

3. If (v0,b0)(v_{0},b_{0}) satisfies

supR>0NR0<ϵ1,\sup_{R>0}N_{R}^{0}<\epsilon_{1},

then the set of singular times of (v,b)(v,b) in 3×(0,)\mathbb{R}^{3}\times(0,\infty) is empty. Moreover, for all t>0t>0,

t1/2(v,b)(,t)L×L(supR>0NR0)1/2<.t^{1/2}\left\|(v,b)(\cdot,t)\right\|_{L^{\infty}\times L^{\infty}}{\ \lesssim\ }\left(\sup_{R>0}N_{R}^{0}\right)^{1/2}<\infty.
Proof.

The proof of the theorem is an adaption of the proof of [3, Theorem 1.2] for the Navier–Stokes equations to the MHD equations.

Assume there exists R0>0R_{0}>0 such that for all RR0R\geq R_{0}, we have NR0(v0,b0)<ϵ1N_{R}^{0}(v_{0},b_{0})<\epsilon_{1}, where ϵ1(0,1)\epsilon_{1}\in(0,1) is a small constant to be determined.

Fix x03x_{0}\in\mathbb{R}^{3} and R>R0R>R_{0}. Define p~(x,t)=p(x,t)cx0,R(t)\tilde{p}(x,t)=p(x,t)-c_{x_{0},R}(t) where cx0,R(t)c_{x_{0},R}(t) is the function of tt from the local pressure expansion (1.27). Then (v,b)(v,b) is a suitable weak solution to (MHD) with associated pressure p~\tilde{p}. By the estimate (3.2), we have

0σ(R)R2BR(x0)(|v|3+|b|3+|p~|3/2)𝑑x𝑑tC(NR0)32R2.\int_{0}^{\sigma(R)R^{2}}\int_{B_{R}(x_{0})}\left(|v|^{3}+|b|^{3}+|\tilde{p}|^{3/2}\right)dxdt\leq C(N_{R}^{0})^{\frac{3}{2}}R^{2}.

Thus, if RR0R\geq R_{0} and ϵ1(c0C1ϵ)2/3\epsilon_{1}\leq\left(c_{0}C^{-1}\epsilon_{*}\right)^{2/3}, then the right side is bounded by ϵ\epsilon_{*}, and we may apply Lemma 3.2. It follows that

v,bL(Q),Q=Bc01/2R2(x0)×[3c0R24,c0R2],v,b\in L^{\infty}(Q),\quad Q=B_{\frac{c_{0}^{1/2}R}{2}}(x_{0})\times\left[\frac{3c_{0}R^{2}}{4},c_{0}R^{2}\right],

and for (x,t)Q(x,t)\in Q,

|v(x,t)|+|b(x,t)|C0(Cc0(NR0)3/2)13(c01/2R2)1C(NR0)12t12.|v(x,t)|+|b(x,t)|\leq C_{0}\left(\frac{C}{c_{0}}(N_{R}^{0})^{3/2}\right)^{\frac{1}{3}}\left(\frac{c_{0}^{1/2}R}{2}\right)^{-1}\leq C(N_{R}^{0})^{\frac{1}{2}}t^{-\frac{1}{2}}. (3.7)

Hence, (v,b)(v,b) is regular in 3×(3c0R24,c0R2]\mathbb{R}^{3}\times\left(\frac{3c_{0}R^{2}}{4},c_{0}R^{2}\right]. Since RR0R\geq R_{0} is arbitrary, we conclude that (v,b)(v,b) is regular at every point (x,t)3×(3c0R24,)(x,t)\in\mathbb{R}^{3}\times\left(\frac{3c_{0}R^{2}}{4},\infty\right), with the bound given by (3.7). Note that this threshold 3c0R024\frac{3c_{0}R_{0}^{2}}{4} depends only on (v0,b0)(v_{0},b_{0}) and is uniform for all (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}).

The argument proceeds analogously in the case where supRR0NR0<ϵ1\sup_{R\leq R_{0}}N_{R}^{0}<\epsilon_{1}; we omit the details for brevity.

Finally, if NR0<ϵ1N_{R}^{0}<\epsilon_{1} for all R>0R>0, then σ(R)=c0\sigma(R)=c_{0} for all R>0R>0. In this case, (v,b)(v,b) is regular with the same bound (3.7) throughout the entire space-time domain 0<R<3×(3c0R24,c0R2]=3×(0,)\cup_{0<R<\infty}\mathbb{R}^{3}\times\left(\frac{3c_{0}R^{2}}{4},c_{0}R^{2}\right]=\mathbb{R}^{3}\times(0,\infty). ∎

As a consequence of Theorem 3.3, the uniqueness result below can be established by adapting the argument used in the proof of [3, Theorem 1.7].

Theorem 3.4 (Uniqueness for data that is small at high frequencies).

Assume that v0,b0E2v_{0},b_{0}\in E^{2} are divergence free. Let (v,b),(u,a)𝒩MHD(v0,b0)(v,b),(u,a)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}). Then there exist universal constants 0<ϵ3,τ010<\epsilon_{3},\tau_{0}\leq 1 so that if

sup0<rRNr0ϵ3\sup_{0<r\leq R}N_{r}^{0}\leq\epsilon_{3}

for some R>0R>0, then (u,a)=(v,b)(u,a)=(v,b) as distributions on 3×(0,T)\mathbb{R}^{3}\times(0,T), T=τ0R2T=\tau_{0}R^{2}.

Proof.

Assume limR0NR0<ϵ\lim_{R\to 0}N^{0}_{R}<\epsilon for some ϵ>0\epsilon>0, and either (v0,b0)(v_{0},b_{0}) satisfies

limRsupx031R2BR(x0)(|v0(x)|2+|b0(x)|2)𝑑x=0,\lim_{R\to\infty}\sup_{x_{0}\in\mathbb{R}^{3}}\frac{1}{R^{2}}\int_{B_{R}(x_{0})}(|v_{0}(x)|^{2}+|b_{0}(x)|^{2})\,dx=0, (3.8)

or (v0,b0)E2×E2(v_{0},b_{0})\in E^{2}\times E^{2}. Let R0R_{0} satisfy supR<R0NR0<ϵ\sup_{R<R_{0}}N_{R}^{0}<\epsilon. By Theorem 3.3 we have for T=c0R02T=c_{0}R_{0}^{2},

t1/2(v,b)(,t)L×LC(ϵ),0<tT.t^{1/2}\left\|(v,b)(\cdot,t)\right\|_{L^{\infty}\times L^{\infty}}\leq C(\epsilon),\quad 0<t\leq T.

Using (3.1) we have the estimate

supr>t/c0,x31rBr(x)(|v(x,t)|2+|b(x,t)|2)𝑑x<Cϵ2,\sup_{r>\sqrt{t/c_{0}},\,x\in\mathbb{R}^{3}}\frac{1}{r}\int_{B_{r}(x)}(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx<C\epsilon^{2}, (3.9)

where c0c_{0} is defined in (3.3). Moreover, by item 3 of Theorem 3.3, we have

(v,b)(t)L×L<Cϵ2t1/2,0<t<.\left\|(v,b)(t)\right\|_{L^{\infty}\times L^{\infty}}<C\epsilon^{2}t^{-1/2},\quad 0<t<\infty. (3.10)

Combining (3.9) and (3.10), we deduce that, for r<t/c0r<\sqrt{t/c_{0}},

1rBr(x)(|v(x,t)|2+|b(x,t)|2)dx=1r(Br(x)(|v(x,t)|2+|b(x,t)|2)𝑑x)1/3(Br(x)(|v(x,t)|2+|b(x,t)|2)𝑑x)2/3(v,b)(t)L×L2/3(Bt/c0(x)(|v(x,t)|2+|b(x,t)|2)𝑑x)2/3C(ϵ)t1/3(t/c0)1/3=C(ϵ).\begin{split}\frac{1}{r}\int_{B_{r}(x)}&(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx\\ &=\frac{1}{r}\left(\int_{B_{r}(x)}(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx\right)^{1/3}\left(\int_{B_{r}(x)}(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx\right)^{2/3}\\ &{\ \lesssim\ }\left\|(v,b)(t)\right\|_{L^{\infty}\times L^{\infty}}^{2/3}\left(\int_{B_{\sqrt{t/c_{0}}}(x)}(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx\right)^{2/3}\\ &{\ \lesssim\ }C(\epsilon)t^{-1/3}(t/c_{0})^{1/3}=C(\epsilon).\end{split} (3.11)

Using the estimates (3.9) for r=Tr=\sqrt{T} and (3.11) for r<Tr<\sqrt{T}, we have for all t(0,T)t\in(0,T) and r(0,T1/2)r\in(0,T^{1/2}) that

1rBr(x)(|v(x,t)|2+|b(x,t)|2)𝑑xC(ϵ),\frac{1}{r}\int_{B_{r}(x)}(|v(x,t)|^{2}+|b(x,t)|^{2})\,dx\leq C(\epsilon),

which implies that

sup0<t<(v,b)(t)Luloc,r2×Luloc,r2<C(ϵ),\sup_{0<t<\infty}\left\|(v,b)(t)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}<C(\epsilon), (3.12)

where fLuloc,r2:=supx3fL2(x)\left\|f\right\|_{L^{2}_{\mathrm{uloc},r}}:=\sup_{x\in\mathbb{R}^{3}}\left\|f\right\|_{L^{2}(x)}.

We now check that (v,b)(v,b) satisfies the integral formula (1.8), i.e., it is a mild solution, on 3×(0,T)\mathbb{R}^{3}\times(0,T). If (v0,b0)E2×E2(v_{0},b_{0})\in E^{2}\times E^{2}, then this follows from a direct adaption of [20, §8] for the Navier–Stokes equations to the MHD equations. On the other hand, assume (v0,b0)(v_{0},b_{0}) satisfies (3.8). By (3.1), we have

esssup0tσ(r)r2(v,b)Luloc,r2×Luloc,r22<CA0(r),A0(r)=(v0,b0)Luloc,r2×Luloc,r22.\mathop{\rm ess\,sup}_{0\leq t\leq\sigma(r)r^{2}}\left\|(v,b)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}<CA_{0}(r),\qquad A_{0}(r)=\left\|(v_{0},b_{0})\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}.

Since (v0,b0)(v_{0},b_{0}) satisfies (3.8), we have σ(r)r2\sigma(r)r^{2}\to\infty as rr\to\infty. So, there exists R¯\bar{R} so that, for all R>R¯R>\bar{R}, σ(R)R2>T\sigma(R)R^{2}>T. We conclude that for any r>0r>0

esssup0tT(v,b)Luloc,r2×Luloc,r22f(r),\mathop{\rm ess\,sup}_{0\leq t\leq T}\left\|(v,b)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}\leq f(r), (3.13)

where f(r)=CA0(r)+CA0(R¯)f(r)=CA_{0}(r)+CA_{0}(\bar{R}).

Let (v~,b~)(\tilde{v},\tilde{b}) be defined as

(v~,b~)(x,t)=(etΔv0,etΔb0)B((v,b),(v,b))(t),(\tilde{v},\tilde{b})(x,t)=(e^{t\Delta}v_{0},e^{t\Delta}b_{0})-B((v,b),(v,b))(t), (3.14)

where BB is a bilinear operator defined by B=(B1,B2)B=(B_{1},B_{2}), where B1B_{1} and B2B_{2} are given in (1.9). Using [33, (1.8), (1.10)], we have

(etΔv0,etΔb0)Luloc,r2×Luloc,r2(v0,b0)Luloc,r2×Luloc,r2=(A0(r))1/2,\left\|(e^{t\Delta}v_{0},e^{t\Delta}b_{0})\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}\leq\left\|(v_{0},b_{0})\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}=(A_{0}(r))^{1/2},

and

0te(ts)Δ(vvbb)(s)𝑑sLuloc,r2+0te(ts)Δ(vbbv)(s)𝑑sLuloc,r20tC(ts)1/2(v,b)(s)L×L(v,b)(s)Luloc,r2×Luloc,r2𝑑s0tC(ϵ)(ts)1/2s1/2𝑑s=C(ϵ).\begin{split}&\left\|\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(v\otimes v-b\otimes b)(s)\,ds\right\|_{L^{2}_{\mathrm{uloc},r}}+\left\|\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(v\otimes b-b\otimes v)(s)\,ds\right\|_{L^{2}_{\mathrm{uloc},r}}\\ &\leq\int_{0}^{t}\frac{C}{(t-s)^{1/2}}\left\|(v,b)(s)\right\|_{L^{\infty}\times L^{\infty}}\left\|(v,b)(s)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}\,ds\\ &\leq\int_{0}^{t}\frac{C(\epsilon)}{(t-s)^{1/2}s^{1/2}}\,ds=C(\epsilon).\end{split}

Thus

sup0tT(v~,b~)Luloc,r2×Luloc,r22CA0(r)+C(ϵ)sup0<s<T(v,b)Luloc,r2×Luloc,r22Cf(r).\sup_{0\leq t\leq T}\left\|(\tilde{v},\tilde{b})\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}\leq CA_{0}(r)+C(\epsilon)\sup_{0<s<T}\left\|(v,b)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}\leq Cf(r). (3.15)

Then (V,B)=(v,b)(v~,b~)(V,B)=(v,b)-(\tilde{v},\tilde{b}) satisfies

esssup0tT(V,B)Luloc,r2×Luloc,r22Cf(r),r>0.\mathop{\rm ess\,sup}_{0\leq t\leq T}\left\|(V,B)\right\|_{L^{2}_{\mathrm{uloc},r}\times L^{2}_{\mathrm{uloc},r}}^{2}\leq Cf(r),\quad\forall r>0.

Following the same logic in [20, §8] and adapting it to MHD equations, we have that the mollified VϵV_{\epsilon} and BϵB_{\epsilon} are harmonic in xx and for fixed t(0,T)t\in(0,T)

|Vϵ(x,t)|+|Bϵ(x,t)|C(1r3Br(x)(|Vϵ(y,t)|2+|Bϵ(y,t)|2)𝑑y)1/2C(f(r)r3)1/20 as r,|V_{\epsilon}(x,t)|+|B_{\epsilon}(x,t)|\leq C\left(\frac{1}{r^{3}}\int_{B_{r}(x)}(|V_{\epsilon}(y,t)|^{2}+|B_{\epsilon}(y,t)|^{2})\,dy\right)^{1/2}\leq C\left(\frac{f(r)}{r^{3}}\right)^{1/2}\to 0\ \text{ as }\ r\to\infty,

This shows (Vϵ,Bϵ)=(0,0)(V_{\epsilon},B_{\epsilon})=(0,0) for t<Tt<T, for all ϵ>0\epsilon>0. Hence (V,B)=(0,0)(V,B)=(0,0) and (v,b)=(v~,b~)(v,b)=(\tilde{v},\tilde{b}). This shows that any local energy solution with data satisfying the assumptions of Theorem 3.4 is a mild solution.

Assume (u,a)𝒩MHD(v0,b0)(u,a)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}) also and satisfy the assumptions of Theorem 3.4. Then, (u,a)(u,a) is also a mild solution. Let (w,d)=(v,b)(u,a)(w,d)=(v,b)-(u,a). Then

w(,t)=0te(ts)Δ(ww+uw+wuddadda)(,s)𝑑s,w(\cdot,t)=-\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(w\otimes w+u\otimes w+w\otimes u-d\otimes d-a\otimes d-d\otimes a)(\cdot,s)\,ds,

and

d(,t)=0te(ts)Δ(wd+ud+wadwawdu)(,s)𝑑s.d(\cdot,t)=-\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot(w\otimes d+u\otimes d+w\otimes a-d\otimes w-a\otimes w-d\otimes u)(\cdot,s)\,ds.

By [33, Corollary 3.1] we have for t>0t>0 that

(w,d)(t)Luloc2×Luloc2Csup0<s<ts((v,b)(s)L×L+(u,a)(s)L×L)(w,d)L(Luloc2×Luloc2).\begin{split}&\left\|(w,d)(t)\right\|_{L^{2}_{\mathrm{uloc}}\times L^{2}_{\mathrm{uloc}}}\\ &\qquad\leq C\sup_{0<s<t}\sqrt{s}(\left\|(v,b)(s)\right\|_{L^{\infty}\times L^{\infty}}+\left\|(u,a)(s)\right\|_{L^{\infty}\times L^{\infty}})\left\|(w,d)\right\|_{L^{\infty}(L^{2}_{\mathrm{uloc}}\times L^{2}_{\mathrm{uloc}})}.\end{split}

Since s((v,b)(s)L×L+(u,a)(s)L×L)\sqrt{s}(\left\|(v,b)(s)\right\|_{L^{\infty}\times L^{\infty}}+\left\|(u,a)(s)\right\|_{L^{\infty}\times L^{\infty}}) is small we have

(w,d)(t)Luloc2×Luloc212(w,d)L(Luloc2×Luloc2).\left\|(w,d)(t)\right\|_{L^{2}_{\mathrm{uloc}}\times L^{2}_{\mathrm{uloc}}}\leq\frac{1}{2}\left\|(w,d)\right\|_{L^{\infty}(L^{2}_{\mathrm{uloc}}\times L^{2}_{\mathrm{uloc}})}.

Taking the essential supremum on the left hand side leads to uniqueness. ∎

As a corollary of Theorem 3.4, the following local uniqueness result in E3E^{3} can be derived by adapting the proof of [3, Corollary 1.8].

Corollary 3.5 (Local uniqueness in E3E^{3}).

Assume v0,b0E3v_{0},b_{0}\in E^{3} are divergence free. Let (v,b)(v,b) and (u,a)(u,a) be elements of 𝒩MHD(v0,b0)\mathcal{N}_{\rm MHD}(v_{0},b_{0}). Then, there exists T=T(v0,b0)>0T=T(v_{0},b_{0})>0 so that (v,b)=(u,a)(v,b)=(u,a) as distributions on 3×(0,T)\mathbb{R}^{3}\times(0,T).

Proof.

Assume (v0,b0)E3×E3(v_{0},b_{0})\in E^{3}\times E^{3}. Then, in particular, (v0,b0)E2×E2(v_{0},b_{0})\in E^{2}\times E^{2}. Let ϵ>0\epsilon>0 be given. Since (v0,b0)E3×E3(v_{0},b_{0})\in E^{3}\times E^{3}, there exists R0R_{0} such that

sup|x0|R0B1(x0)(|v0(x)|3+|b0(x)|3)𝑑x<ϵ.\sup_{|x_{0}|\geq R_{0}}\int_{B_{1}(x_{0})}(|v_{0}(x)|^{3}+|b_{0}(x)|^{3})\,dx<\epsilon.

On the other hand, since v0,b0E3v_{0},b_{0}\in E^{3}, their local L3L^{3}-norms are uniformly small on sufficiently small balls. That is, there exists γ(0,1]\gamma\in(0,1] such that

sup|x0|R0, 0<rγBr(x0)(|v0(x)|3+|b0(x)|3)𝑑x<ϵ.\sup_{|x_{0}|\leq R_{0},\,0<r\leq\gamma}\int_{B_{r}(x_{0})}(|v_{0}(x)|^{3}+|b_{0}(x)|^{3})\,dx<\epsilon.

Applying Hölder’s inequality, we obtain:

sup|x0|R0, 0<rγ1rBr(x0)(|v0(x)|2+|b0(x)|2)𝑑x<|B1|1/3ϵ2/3.\sup_{|x_{0}|\leq R_{0},\,0<r\leq\gamma}\frac{1}{r}\int_{B_{r}(x_{0})}(|v_{0}(x)|^{2}+|b_{0}(x)|^{2})\,dx<|B_{1}|^{1/3}\epsilon^{2/3}.

Therefore, by Theorem 3.4, any local energy solution with initial data (v0,b0)(v_{0},b_{0}) is unique in the local energy class, at least for a short time. ∎

3.1 Eventual regularity for local energy solutions

Lemma 3.6.

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q}, 2q<2\leq q<\infty. Then

limRR6q2Nq,R0(v0,b0)=0.\lim_{R\to\infty}R^{\frac{6}{q}-2}N_{q,R}^{0}(v_{0},b_{0})=0.

Consequently, if v0,b0Eq2v_{0},b_{0}\in E^{2}_{q}, then

limRNR0(v0,b0)=0 if  2q3 and limRR1Nq,R0(v0,b0)=0 if  2q6.\lim_{R\to\infty}N^{0}_{R}(v_{0},b_{0})=0\ \text{ if }\ 2\leq q\leq 3\quad\text{ and }\quad\lim_{R\to\infty}R^{-1}N^{0}_{q,R}(v_{0},b_{0})=0\ \text{ if }\ 2\leq q\leq 6.
Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 2.2] for the Navier–Stokes equations to the MHD equations.

Let ϵ>0\epsilon>0 be given. Suppose that v0,b0Eq2v_{0},b_{0}\in E^{2}_{q} for some 2q<2\leq q<\infty. Then the norms can be expressed as v0Eq2=uq(3)\left\|v_{0}\right\|_{E^{2}_{q}}=\left\|u\right\|_{\ell^{q}(\mathbb{Z}^{3})} and v0Eq2=aq(3)\left\|v_{0}\right\|_{E^{2}_{q}}=\left\|a\right\|_{\ell^{q}(\mathbb{Z}^{3})}, where

u=(uk)k3q(3),uk=v0L2(B1(k)),u=(u_{k})_{k\in\mathbb{Z}^{3}}\in\ell^{q}(\mathbb{Z}^{3}),\quad u_{k}=\left\|v_{0}\right\|_{L^{2}(B_{1}(k))},
a=(ak)k3q(3),ak=b0L2(B1(k)).a=(a_{k})_{k\in\mathbb{Z}^{3}}\in\ell^{q}(\mathbb{Z}^{3}),\quad a_{k}=\left\|b_{0}\right\|_{L^{2}(B_{1}(k))}.

For any R1R\geq 1, we have the estimate

Nq,R0(v0,b0)CR[k3(|ikR|<R(ui2+ai2))q2]2q\begin{split}N_{q,R}^{0}(v_{0},b_{0})\leq\frac{C}{R}\left[\sum_{k\in\mathbb{Z}^{3}}\left(\sum_{|i-kR|<R}(u_{i}^{2}+a_{i}^{2})\right)^{\frac{q}{2}}\right]^{\frac{2}{q}}\end{split} (3.16)

Now fixe any δ>0\delta>0. Since u,aq(3)u,a\in\ell^{q}(\mathbb{Z}^{3}), we can choose M>1M>1 large enough such that u>Mqδ\left\|u^{>M}\right\|_{\ell^{q}}\leq\delta and a>Mqδ\left\|a^{>M}\right\|_{\ell^{q}}\leq\delta, where

uk>M={0 if |k|M,uk if |k|>M, and ak>M={0 if |k|M,ak if |k|>M.u^{>M}_{k}=\begin{cases}0&\ \text{ if }|k|\leq M,\\ u_{k}&\ \text{ if }|k|>M,\end{cases}\quad\text{ and }\quad a^{>M}_{k}=\begin{cases}0&\ \text{ if }|k|\leq M,\\ a_{k}&\ \text{ if }|k|>M.\end{cases}

Set uM=uu>Mu^{\leq M}=u-u^{>M} and aM=aa>Ma^{\leq M}=a-a^{>M}. Following the argument leading to [4, (2.4)] by Hölder’s inequality we obtain, for all R>MR>M,

[R6q2Nq,R0(v0,b0)]q2C(u>Mqq+a>Mqq)+CR33q2M3(q2)2(uMqq+aMqq).\left[R^{\frac{6}{q}-2}N_{q,R}^{0}(v_{0},b_{0})\right]^{\frac{q}{2}}\leq C\left(\left\|u^{>M}\right\|_{\ell^{q}}^{q}+\left\|a^{>M}\right\|_{\ell^{q}}^{q}\right)+CR^{3-\frac{3q}{2}}M^{\frac{3(q-2)}{2}}\left(\left\|u^{\leq M}\right\|_{\ell^{q}}^{q}+\left\|a^{\leq M}\right\|_{\ell^{q}}^{q}\right).

To conclude, we first choose δ>0\delta>0 small enough so that C(u>Mqq+a>Mqq)<ϵ/2C\left(\left\|u^{>M}\right\|_{\ell^{q}}^{q}+\left\|a^{>M}\right\|_{\ell^{q}}^{q}\right)<\epsilon/2. Then, for this fixed M=M(δ)M=M(\delta), we choose RR sufficiently large to ensure that CR33q2M(δ)3(q2)2(uMqq+aMqq)<ϵ/2CR^{3-\frac{3q}{2}}M(\delta)^{\frac{3(q-2)}{2}}\left(\left\|u^{\leq M}\right\|_{\ell^{q}}^{q}+\left\|a^{\leq M}\right\|_{\ell^{q}}^{q}\right)<\epsilon/2, which is possible provided q>2q>2.

To prove the final assertions, we begin by noting that NR0(v0,b0)Nq,R0(v0,b0)N_{R}^{0}(v_{0},b_{0})\leq N_{q,R}^{0}(v_{0},b_{0}). Moreover, if v0,b0Eq2v_{0},b_{0}\in E^{2}_{q} for q3q\leq 3, then in particular v0,b0E32v_{0},b_{0}\in E^{2}_{3}. Therefore,

limRNR0(v0,b0)limRN3,R0(v0,b0)=0.\lim_{R\to\infty}N_{R}^{0}(v_{0},b_{0})\leq\lim_{R\to\infty}N_{3,R}^{0}(v_{0},b_{0})=0.

For the final part, observe that when q6q\leq 6 and R1R\geq 1, we have R1R6q2R^{-1}\leq R^{\frac{6}{q}-2}. Thus,

limRR1Nq,R0(v0,b0)=0.\lim_{R\to\infty}R^{-1}N^{0}_{q,R}(v_{0},b_{0})=0.

Proof of Theorem 1.8.

For 2q32\leq q\leq 3, Lemma 3.6 implies

limRNR0(v0,b0)=0.\lim_{R\to\infty}N^{0}_{R}(v_{0},b_{0})=0.

Applying Theorem 3.3 then yields the desired result.

For 1q<21\leq q<2, the same conclusion follows from the fact that v0,b0Eq2L2v_{0},b_{0}\in E^{2}_{q}\subset L^{2} since Eq2E^{2}_{q} embeds into L2L^{2} when q<2q<2. This completes the proof of Theorem 1.8. ∎

3.2 A priori bounds and explicit growth rate

In this section we prove new a priori bounds for data (v0,b0)Eq2×Eq2(v_{0},b_{0})\in E^{2}_{q}\times E^{2}_{q} and use it to prove Theorem 1.9.

Lemma 3.7.

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q} for some q1q\geq 1 are divergence free and that (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}) satisfies, for some T2>0T_{2}>0,

esssup0TT1B1(x0)(|v|2+|b|2)𝑑x+0T1B1(x0)(|v|2+|b|2)𝑑x𝑑tq2(x03)< for all T1(0,T2).\begin{split}\left\|\mathop{\rm ess\,sup}_{0\leq T\leq T_{1}}\int_{B_{1}(x_{0})}\left(|v|^{2}+|b|^{2}\right)dx+\int_{0}^{T_{1}}\int_{B_{1}(x_{0})}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}<\infty\ \text{ for all }T_{1}\in(0,T_{2}).\end{split} (3.17)

Then there are positive constants C1C_{1} and λ0<1\lambda_{0}<1, both independent of qq and RR such that, for all R>0R>0 with λRR2T2\lambda_{R}R^{2}\leq T_{2},

esssup0tλRR2BR(x0R)|v|2+|b|22𝑑x+0λRR2BR(x0R)(|v|2+|b|2)𝑑x𝑑tq2(x03)C1A0,q(R),\begin{split}\left\|\mathop{\rm ess\,sup}_{0\leq t\leq\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}\frac{|v|^{2}+|b|^{2}}{2}\,dx+\int_{0}^{\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}\leq C_{1}A_{0,q}(R),\end{split} (3.18)

where

A0,q(R)=RNq,R0=BR(x0R)(|v0|2+|b0|2)𝑑xq2(x03),λR=min(λ0,λ0R2(A0,q(R))2).A_{0,q}(R)=RN^{0}_{q,R}=\left\|\int_{B_{R}(x_{0}R)}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})},\quad\lambda_{R}=\min\left(\lambda_{0},\,\frac{\lambda_{0}R^{2}}{\left(A_{0,q}(R)\right)^{2}}\right).

Furthermore, for all R>0R>0,

0λRR2BR(x0R)|v|103+|b|103+|πcRx0,R(t)|53dxdt3q10(x03)C(A0,q(R))53,q2,\begin{split}\left\|\int_{0}^{\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}+\left|\pi-c_{Rx_{0},R}(t)\right|^{\frac{5}{3}}dxdt\right\|_{\ell^{\frac{3q}{10}}(x_{0}\in\mathbb{Z}^{3})}\leq C\left(A_{0,q}(R)\right)^{\frac{5}{3}},\quad q\geq 2,\end{split} (3.19)

and

0λRR2BR(x0R)|v|103+|b|103+|πcRx0,R(t)|53dxdtσ(x03)C(A0,q(R))53,1q<2.\begin{split}\left\|\int_{0}^{\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}+\left|\pi-c_{Rx_{0},R}(t)\right|^{\frac{5}{3}}dxdt\right\|_{\ell^{\sigma}(x_{0}\in\mathbb{Z}^{3})}\leq C\left(A_{0,q}(R)\right)^{\frac{5}{3}},\quad 1\leq q<2.\end{split} (3.20)

for all σ35\sigma\geq\frac{3}{5}.

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 3.1] for the Navier–Stokes equations to the MHD equations.

Let ϕ0Cc(3)\phi_{0}\in C^{\infty}_{c}(\mathbb{R}^{3}) be radial, non-increasing cutoff function such that ϕ01\phi_{0}\equiv 1 on B1(0)B_{1}(0), suppϕ0B2(0)\mathop{\mathrm{supp}}\nolimits\phi_{0}\subset B_{2}(0), and |ϕ0(x)| 1|\nabla\phi_{0}(x)|{\ \lesssim\ }1, |ϕ01/2(x)| 1|\nabla\phi_{0}^{1/2}(x)|{\ \lesssim\ }1. Let R>0R>0 be as in the statement of the lemma, and define the scaled cutoff ϕ(x):=ϕ0(x/R)\phi(x):=\phi_{0}(x/R). Fix 0<λ10<\lambda\leq 1.

For each κR3\kappa\in R\mathbb{Z}^{3}, define the localized energy quantity

eR,λ(κ):=esssup0tλR2(|v(t)|2+|b(t)|2)ϕ(xκ)𝑑x+0λR2(|v|2+|b|2)ϕ(xκ)𝑑x𝑑t.e_{R,\lambda}(\kappa):=\mathop{\rm ess\,sup}_{0\leq t\leq\lambda R^{2}}\int\left(|v(t)|^{2}+|b(t)|^{2}\right)\phi(x-\kappa)\,dx+\int_{0}^{\lambda R^{2}}\int\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi(x-\kappa)\,dxdt.

We begin by deriving bounds on eR,λ(κ)e_{R,\lambda}(\kappa), which will then be used to control the quantity

ER,q,λ:=esssup0tλR2(|v(t)|2+|b(t)|2)ϕ(xRk)𝑑x+0λR2(|v|2+|b|2)ϕ(xRk)𝑑x𝑑tq2(k3)q2E_{R,q,\lambda}:=\left\|\mathop{\rm ess\,sup}_{0\leq t\leq\lambda R^{2}}\int\left(|v(t)|^{2}+|b(t)|^{2}\right)\phi(x-Rk)\,dx+\int_{0}^{\lambda R^{2}}\int\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi(x-Rk)\,dxdt\right\|_{\ell^{\frac{q}{2}}(k\in\mathbb{Z}^{3})}^{\frac{q}{2}}

in terms of A0,q(R)A_{0,q}(R) for sufficiently small λ\lambda. By assumption, ER,q,λ<E_{R,q,\lambda}<\infty. To estimate eR,λ(κ)e_{R,\lambda}(\kappa), we apply the local energy inequality (1.33):

(|v(t)|2+|b(t)|2)ϕ(xκ)dx+20t(|v|2+|b|2)ϕ(xκ)𝑑x𝑑s(|v0|2+|b0|2)ϕ(xκ)𝑑x+0t(|v|2+|b|2)Δϕ(xκ)𝑑x𝑑s+0t(|v|2+|b|2)(vϕ(xκ))𝑑x𝑑s+0t2π(vϕ(xκ))𝑑x𝑑s20t(bv)(bϕ(xκ))𝑑x𝑑s.\begin{split}\int&\left(|v(t)|^{2}+|b(t)|^{2}\right)\phi(x-\kappa)\,dx+2\int_{0}^{t}\int\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi(x-\kappa)\,dxds\\ &\leq\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi(x-\kappa)\,dx+\int_{0}^{t}\int\left(|v|^{2}+|b|^{2}\right)\Delta\phi(x-\kappa)\,dxds\\ &\quad+\int_{0}^{t}\int\left(|v|^{2}+|b|^{2}\right)\left(v\cdot\nabla\phi(x-\kappa)\right)dxds+\int_{0}^{t}\int 2\pi\left(v\cdot\nabla\phi(x-\kappa)\right)dxds\\ &\quad-2\int_{0}^{t}\int(b\cdot v)\left(b\cdot\nabla\phi(x-\kappa)\right)dxds.\end{split}

We now estimate each term on the right-hand side, beginning with the second term. Using the properties of ϕ\phi, we have:

0λR2(|v|2+|b|2)|Δϕ(xκ)|𝑑x𝑑sCR20λR2B2R(κ)(|v|2+|b|2)𝑑x𝑑sCλκR3;|κκ|2Resssup0tλR2(|v|2+|b|2)ϕ(xκ)𝑑xCλκR3;|κκ|2ReR,λ(κ).\begin{split}\int_{0}^{\lambda R^{2}}\int\left(|v|^{2}+|b|^{2}\right)\left|\Delta\phi(x-\kappa)\right|dxds&\leq\frac{C}{R^{2}}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}\left(|v|^{2}+|b|^{2}\right)dxds\\ &\leq C\lambda\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}\mathop{\rm ess\,sup}_{0\leq t\leq\lambda R^{2}}\int\left(|v|^{2}+|b|^{2}\right)\phi(x-\kappa^{\prime})\,dx\\ &\leq C\lambda\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}e_{R,\lambda}(\kappa^{\prime}).\end{split}

For the cubic terms, we apply the Gagliardo–Nirenberg inequality:

B2R|v|3𝑑x(B2R|v|2)34(B2R|v|2)34+R32(B2R|v|2)32,\int_{B_{2R}}|v|^{3}\,dx{\ \lesssim\ }\left(\int_{B_{2R}}|\nabla v|^{2}\right)^{\frac{3}{4}}\left(\int_{B_{2R}}|v|^{2}\right)^{\frac{3}{4}}+R^{-\frac{3}{2}}\left(\int_{B_{2R}}|v|^{2}\right)^{\frac{3}{2}},
B2R|b|3𝑑x(B2R|b|2)34(B2R|b|2)34+R32(B2R|b|2)32.\int_{B_{2R}}|b|^{3}\,dx{\ \lesssim\ }\left(\int_{B_{2R}}|\nabla b|^{2}\right)^{\frac{3}{4}}\left(\int_{B_{2R}}|b|^{2}\right)^{\frac{3}{4}}+R^{-\frac{3}{2}}\left(\int_{B_{2R}}|b|^{2}\right)^{\frac{3}{2}}.

Let N=sup0tλR2B2R(|v(t)|2+|b(t)|2)𝑑x+20λR2B2R(|v|2+|b|2)𝑑x𝑑tN=\sup_{0\leq t\leq\lambda R^{2}}\int_{B_{2R}}\left(|v(t)|^{2}+|b(t)|^{2}\right)dx+2\int_{0}^{\lambda R^{2}}\int_{B_{2R}}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt. Then, integrating the above estimates over time yields

0λR2B2R(|v|3+|b|3)𝑑x𝑑sN340λR2[(B2R|v|2)34+(B2R|b|2)34]𝑑s+R32N32λR2N32(λR2)14+N32λR12N32λ14R12,\begin{split}\int_{0}^{\lambda R^{2}}\int_{B_{2R}}\left(|v|^{3}+|b|^{3}\right)\,dxds&{\ \lesssim\ }N^{\frac{3}{4}}\int_{0}^{\lambda R^{2}}\left[\left(\int_{B_{2R}}|\nabla v|^{2}\right)^{\frac{3}{4}}+\left(\int_{B_{2R}}|\nabla b|^{2}\right)^{\frac{3}{4}}\right]ds+R^{-\frac{3}{2}}N^{\frac{3}{2}}\lambda R^{2}\\ &{\ \lesssim\ }N^{\frac{3}{2}}(\lambda R^{2})^{\frac{1}{4}}+N^{\frac{3}{2}}\lambda R^{\frac{1}{2}}\\ &{\ \lesssim\ }N^{\frac{3}{2}}\lambda^{\frac{1}{4}}R^{\frac{1}{2}},\end{split} (3.21)

where we’ve used λ1\lambda\leq 1 in the final step. As a consequence, we have

0λR2(|v|2+|b|2)(vϕ(xκ))𝑑x𝑑sCR0λR2B2R(κ)(|v|3+|b|3)𝑑x𝑑sCR12λ14κR3;|κκ|4R(eR,λ(κ))32,\begin{split}\int_{0}^{\lambda R^{2}}\int\left(|v|^{2}+|b|^{2}\right)\left(v\cdot\nabla\phi(x-\kappa)\right)dxds&\leq\frac{C}{R}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}\left(|v|^{3}+|b|^{3}\right)dxds\\ &\leq CR^{-\frac{1}{2}}\lambda^{\frac{1}{4}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 4R}(e_{R,\lambda}(\kappa^{\prime}))^{\frac{3}{2}},\end{split}

and

20λR2(bv)(bϕ(xκ))𝑑x𝑑sCR0λR2B2R(κ)|v||b|2𝑑x𝑑sCR0λR2B2R(κ)(|v|3+|b|3)𝑑x𝑑sCR12λ14κR3;|κκ|4R(eR,λ(κ))32.\begin{split}-2\int_{0}^{\lambda R^{2}}\int(b\cdot v)\left(b\cdot\nabla\phi(x-\kappa)\right)dxds&\leq\frac{C}{R}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}|v||b|^{2}\,dxds\\ &\leq\frac{C}{R}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}\left(|v|^{3}+|b|^{3}\right)dxds\\ &\leq CR^{-\frac{1}{2}}\lambda^{\frac{1}{4}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 4R}(e_{R,\lambda}(\kappa^{\prime}))^{\frac{3}{2}}.\end{split}

Now, the only remaining term is the pressure term. To estimate it, we use the local pressure expansion (1.27) to write π(x,t)\pi(x,t) for xB2R(κ)x\in B_{2R}(\kappa) as

π(x,t)=Δ1divdiv[(vvbb)χ4R(xκ)]3(K(xy)K(κy))(vvbb)(y,t)(1χ4R(yκ))𝑑y+cκ,R(t)=:π1(x,t)+π2(x,t)+cκ,R(t).\begin{split}\pi(x,t)&=-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[(v\otimes v-b\otimes b)\chi_{4R}(x-\kappa)\right]\\ &\quad-\int_{\mathbb{R}^{3}}\left(K(x-y)-K(\kappa-y)\right)(v\otimes v-b\otimes b)(y,t)\left(1-\chi_{4R}(y-\kappa)\right)dy+c_{\kappa,R}(t)\\ &=:\pi_{1}(x,t)+\pi_{2}(x,t)+c_{\kappa,R}(t).\end{split}

Note that

|K(xy)K(κy)|CR|κy|4\begin{split}|K(x-y)-K(\kappa-y)|\leq\frac{CR}{|\kappa-y|^{4}}\end{split} (3.22)

for |xκ|2R|x-\kappa|\leq 2R and |κy|4R|\kappa-y|\geq 4R. This ensures that π2\pi_{2} is well-defined even if vv and bb lack decay at infinity.

For the localized term π1\pi_{1}, standard Calderon–Zygmund theory gives

π1L32(B2R(κ))vχ4R1/2(κ)L32+bχ4R1/2(κ)L32CκR3;|κκ|9R(vϕ1/2(κ)L32+bϕ1/2(κ)L32),\begin{split}\left\|\pi_{1}\right\|_{L^{\frac{3}{2}}(B_{2R}(\kappa))}&\leq\left\|v\chi_{4R}^{1/2}(\cdot-\kappa)\right\|_{L^{3}}^{2}+\left\|b\chi_{4R}^{1/2}(\cdot-\kappa)\right\|_{L^{3}}^{2}\\ &\leq C\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 9R}\left(\left\|v\phi^{1/2}(\cdot-\kappa)\right\|_{L^{3}}^{2}+\left\|b\phi^{1/2}(\cdot-\kappa)\right\|_{L^{3}}^{2}\right),\end{split}

where we used the support and scaling properties of the cutoff functions. Then, applying Hölder’s inequality and the inequality (j=1naj2)(j=1n|aj|)nj=1n|aj|3\left(\sum_{j=1}^{n}a_{j}^{2}\right)\left(\sum_{j=1}^{n}|a_{j}|\right)\leq n\sum_{j=1}^{n}|a_{j}|^{3} along with the bound from (3.21), we obtain

0λR22π1(vϕ(xκ))dxdsCR0λR2κR3;|κκ|9R(vϕ1/2(κ)L33+bϕ1/2(κ)L33)dsCR12λ14κR3;|κκ|10R(eR,λ(κ))32.\begin{split}\int_{0}^{\lambda R^{2}}\int&2\pi_{1}\left(v\cdot\nabla\phi(x-\kappa)\right)dxds\\ &\leq\frac{C}{R}\int_{0}^{\lambda R^{2}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 9R}\left(\left\|v\phi^{1/2}(\cdot-\kappa^{\prime})\right\|_{L^{3}}^{3}+\left\|b\phi^{1/2}(\cdot-\kappa^{\prime})\right\|_{L^{3}}^{3}\right)ds\\ &\leq CR^{-\frac{1}{2}}\lambda^{\frac{1}{4}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 10R}(e_{R,\lambda}(\kappa^{\prime}))^{\frac{3}{2}}.\end{split}

To estimate π2\pi_{2}, we begin with the following pointwise bound for xB2R(κ)x\in B_{2R}(\kappa):

|π2(x,t)|CR|κy|4(|v|2+|b|2)(y,t)(1χ4R(yκ))𝑑yCκR3;|κκ|>4RB2R(κ)R|κy|4(|v(y,t)|2+|b(y,t)|2)ϕ(yκ)𝑑yCR3κR3;|κκ|>4R1|κ/Rκ/R|4B2R(κ)(|v(y,t)|2+|b(y,t)|2)ϕ(yκ)𝑑yCR3(K¯eR,λ)(κ),\begin{split}|\pi_{2}(x,t)|&\leq C\int\frac{R}{|\kappa-y|^{4}}\left(|v|^{2}+|b|^{2}\right)(y,t)\left(1-\chi_{4R}(y-\kappa)\right)dy\\ &\leq C\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|>4R}\int_{B_{2R}(\kappa^{\prime})}\frac{R}{|\kappa-y|^{4}}\left(|v(y,t)|^{2}+|b(y,t)|^{2}\right)\phi(y-\kappa^{\prime})\,dy\\ &\leq\frac{C}{R^{3}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|>4R}\frac{1}{|\kappa/R-\kappa^{\prime}/R|^{4}}\int_{B_{2R}(\kappa^{\prime})}\left(|v(y,t)|^{2}+|b(y,t)|^{2}\right)\phi(y-\kappa^{\prime})\,dy\\ &\leq\frac{C}{R^{3}}\left(\overline{K}*e_{R,\lambda}\right)(\kappa),\end{split}

where we used the estimate (3.22), and the convolution K¯eR,λ\overline{K}*e_{R,\lambda} is taken over the lattice R3R\mathbb{Z}^{3}, with

K¯(x)={1|x/R|4 if |x|>4R,K¯(x)=0 otherwise, for xR3.\overline{K}(x)=\begin{cases}\dfrac{1}{|x/R|^{4}}\ \text{ if }|x|>4R,\\ \overline{K}(x)=0\ \text{ otherwise,}\end{cases}\ \text{ for $x\in R\mathbb{Z}^{3}$.}

Assume now that q2q\geq 2. Using λ1\lambda\leq 1, we estimate

0λR22π2(x,s)(v(x,s)ϕ(xκ))𝑑x𝑑sCR0λR2B2R(κ)|π2|32𝑑x𝑑s+CR0λR2B2R(κ)|v|3𝑑x𝑑sCλ14R12((K¯eR,λ)(κ))32+Cλ14R12|κκ|4R(eR,λ(κ))32.\begin{split}\int_{0}^{\lambda R^{2}}&\int 2\pi_{2}(x,s)\left(v(x,s)\cdot\nabla\phi(x-\kappa)\right)dxds\\ &\leq\frac{C}{R}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}|\pi_{2}|^{\frac{3}{2}}\,dxds+\frac{C}{R}\int_{0}^{\lambda R^{2}}\int_{B_{2R}(\kappa)}|v|^{3}\,dxds\\ &\leq C\lambda^{\frac{1}{4}}R^{-\frac{1}{2}}\left((\overline{K}*e_{R,\lambda})(\kappa)\right)^{\frac{3}{2}}+C\lambda^{\frac{1}{4}}R^{-\frac{1}{2}}\sum_{|\kappa^{\prime}-\kappa|\leq 4R}\left(e_{R,\lambda}(\kappa^{\prime})\right)^{\frac{3}{2}}.\end{split} (3.23)

Note that 0λR22cx0,R(s)vϕ(xκ)𝑑x𝑑s=0\int_{0}^{\lambda R^{2}}\int 2c_{x_{0},R}(s)v\cdot\nabla\phi(x-\kappa)\,dxds=0.

Combining all estimates, we obtain the key bound:

eR,λ(κ)(|v0|2+|b0|2)ϕ(xκ)𝑑x+CλκR3;|κκ|2ReR,λ(κ)+Cλ14R12κR3;|κκ|2R(eR,λ(κ))32+Cλ14R12((K¯eR,λ)(κ))32,\begin{split}e_{R,\lambda}(\kappa)&\leq\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi(x-\kappa)\,dx+C\lambda\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}e_{R,\lambda}(\kappa^{\prime})\\ &\quad+C\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}\left(e_{R,\lambda}(\kappa^{\prime})\right)^{\frac{3}{2}}+C\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\left((\overline{K}*e_{R,\lambda})(\kappa)\right)^{\frac{3}{2}},\end{split} (3.24)

provided λ1\lambda\leq 1. Note that all constants here are independent of qq. We now raise both sides of (3.24) to the power q/2q/2 and sum over κR3\kappa\in R\mathbb{Z}^{3}. The left-hand side yields ER,q,λE_{R,q,\lambda}. For the right-hand side:

  • The initial data term is controlled by

    κR3[(|v0|2+|b0|2)ϕ(xκ)𝑑x]q2Cq(A0,q(R))q2,\sum_{\kappa\in R\mathbb{Z}^{3}}\left[\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi(x-\kappa)\,dx\right]^{\frac{q}{2}}\leq C^{q}\left(A_{0,q}(R)\right)^{\frac{q}{2}},
  • The linear term gives

    κR3[CλκR3;|κκ|2ReR,λ(κ)]q2Cqλq2ER,q,λ,\sum_{\kappa\in R\mathbb{Z}^{3}}\left[C\lambda\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}e_{R,\lambda}(\kappa^{\prime})\right]^{\frac{q}{2}}\leq C^{q}\lambda^{\frac{q}{2}}E_{R,q,\lambda},
  • For the nonlinear cubic term, using (j=1naj)pnpj=1najp\left(\sum_{j=1}^{n}a_{j}\right)^{p}\leq n^{p}\sum_{j=1}^{n}a_{j}^{p} for p1p\geq 1 and aj0a_{j}\geq 0, we have

    κR3[Cλ14R12κR3;|κκ|2R(eR,λ(κ))32]q2Cq(λ14R12)q2κR3(eR,λ(κ))3q4.\sum_{\kappa\in R\mathbb{Z}^{3}}\left[C\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};\,|\kappa^{\prime}-\kappa|\leq 2R}\left(e_{R,\lambda}(\kappa^{\prime})\right)^{\frac{3}{2}}\right]^{\frac{q}{2}}\leq C^{q}\left(\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\right)^{\frac{q}{2}}\sum_{\kappa\in R\mathbb{Z}^{3}}\left(e_{R,\lambda}(\kappa)\right)^{\frac{3q}{4}}.
  • For the convolution term we use Young’s convolution inequality to find

    κR3(Cλ14R12((K¯eR,λ)(κ))32)q2Cq(λ14R12)q2κR3((K¯eR,λ)(κ))3q4Cq(λ14R12)q2K¯1(R3)3q4eR,λ3q43q4,\begin{split}\sum_{\kappa\in R\mathbb{Z}^{3}}\left(C\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\left((\overline{K}*e_{R,\lambda})(\kappa)\right)^{\frac{3}{2}}\right)^{\frac{q}{2}}&\leq C^{q}\left(\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\right)^{\frac{q}{2}}\sum_{\kappa\in R\mathbb{Z}^{3}}\left((\overline{K}*e_{R,\lambda})(\kappa)\right)^{\frac{3q}{4}}\\ &\leq C^{q}\left(\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\right)^{\frac{q}{2}}\left\|\overline{K}\right\|_{\ell^{1}(R\mathbb{Z}^{3})}^{\frac{3q}{4}}\left\|e_{R,\lambda}\right\|_{\ell^{\frac{3q}{4}}}^{\frac{3q}{4}},\end{split}

    where K¯1(R3)\left\|\overline{K}\right\|_{\ell^{1}(R\mathbb{Z}^{3})} is bounded independently of RR.

Now, since eR,λ3q4eR,λq2\left\|e_{R,\lambda}\right\|_{\ell^{\frac{3q}{4}}}\leq\left\|e_{R,\lambda}\right\|_{\ell^{\frac{q}{2}}}, we conclude for E=ER,q,λE=E_{R,q,\lambda} and some constant C21C_{2}\geq 1 independent of qq, RR that

EC2q(A0,q(R))q2+C2qλq2E+C2q(λ14R12)q2E32.\begin{split}E\leq C_{2}^{q}\left(A_{0,q}(R)\right)^{\frac{q}{2}}+C_{2}^{q}\lambda^{\frac{q}{2}}E+C_{2}^{q}\left(\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\right)^{\frac{q}{2}}E^{\frac{3}{2}}.\end{split} (3.25)

The right side is finite for λ<R2T2\lambda<R^{-2}T_{2} by assumption (3.17). It follows from the same argument as in [4, p. 2005], ER,q,λE_{R,q,\lambda} is continuous in λ\lambda. So, from (3.25) we conclude that

E2E0,E0=C2q(A0,q(R))q2,E\leq 2E_{0},\qquad E_{0}=C_{2}^{q}\left(A_{0,q}(R)\right)^{\frac{q}{2}},

provided C2qλq21/4C_{2}^{q}\lambda^{\frac{q}{2}}\leq 1/4 and C2q(λ14R12)q2(2E0)121/4C_{2}^{q}\left(\frac{\lambda^{\frac{1}{4}}}{R^{\frac{1}{2}}}\right)^{\frac{q}{2}}(2E_{0})^{\frac{1}{2}}\leq 1/4, which is achieved if (using q2q\geq 2)

λλR:=min(λ0,λ0R2(A0,q(R))2),\begin{split}\lambda\leq\lambda_{R}:=\min\left(\lambda_{0},\,\frac{\lambda_{0}R^{2}}{\left(A_{0,q}(R)\right)^{2}}\right),\end{split}

where λ0=min((2C2)2,(2C2)12)\lambda_{0}=\min\left((2C_{2})^{-2},(2C_{2})^{-12}\right). This shows the first estimate (3.18) of Lemma 3.7, for q2q\geq 2, with C1=CC22C_{1}=CC_{2}^{2}. Note that the constants C2C_{2}, λ0\lambda_{0}, and C1C_{1} do not depend on qq and RR.

For 1q<21\leq q<2, we replace (3.23) by

0λR22π2vϕ(xκ)𝑑x𝑑s1R4λR72vL(0,λR2;L2(B2R(κ)))|K¯eR,λ(κ)|λR1/2(vL(0,λR2;L2(B2R(κ)))2+|K¯eR,λ(κ)|2)λR1/2(κR3;|κκ|2ReR,λ(κ)+|K¯eR,λ(κ)|2).\begin{split}\int_{0}^{\lambda R^{2}}\int 2\pi_{2}v\cdot\nabla\phi(x-\kappa)\,dx\,ds&\lesssim\frac{1}{R^{4}}{\lambda R^{\frac{7}{2}}}\|v\|_{L^{\infty}(0,\lambda R^{2};L^{2}(B_{2R}(\kappa)))}|\overline{K}*e_{R,\lambda}(\kappa)|\\ &\lesssim{\frac{\lambda}{R^{1/2}}}\left(\|v\|_{L^{\infty}(0,\lambda R^{2};L^{2}(B_{2R}(\kappa)))}^{2}+|\overline{K}*e_{R,\lambda}(\kappa)|^{2}\right)\\ &\lesssim{\frac{\lambda}{R^{1/2}}}\left(\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};|\kappa^{\prime}-\kappa|\leq 2R}e_{R,\lambda}(\kappa^{\prime})+|\overline{K}*e_{R,\lambda}(\kappa)|^{2}\right).\end{split}

Raising both sides of the above inequality to the power q/2q/2 and sum over κR3\kappa\in R\mathbb{Z}^{3}, we get

κR3(C(λ+λR1/2)κR3;|κκ|2ReR,λ(κ))q/2Cq(λ+λR1/2)q/2ER,q,λ,\sum_{\kappa\in R\mathbb{Z}^{3}}\bigg{(}C\big{(}\lambda+{\frac{\lambda}{R^{1/2}}}\big{)}\sum_{\kappa^{\prime}\in R\mathbb{Z}^{3};|\kappa^{\prime}-\kappa|\leq 2R}e_{R,\lambda}(\kappa^{\prime})\bigg{)}^{q/2}\leq C^{q}\bigg{(}\lambda+{\frac{\lambda}{R^{1/2}}}\bigg{)}^{q/2}E_{R,q,\lambda},

Above we have used (j=1naj)pj=1najp\left(\sum_{j=1}^{n}a_{j}\right)^{p}\leq\sum_{j=1}^{n}a_{j}^{p} for 0<p<10<p<1 and aj0a_{j}\geq 0. For the convolution term we use Young’s convolution inequality to obtain

κR3(CλR1/2|K¯eR,λ(κ)|2)q/2=(CλR1/2)q2κR3|K¯eR,λ(κ)|q(CλR1/2)q2K¯1qeR,λqq(CλR1/2)q2eR,λq/2q(CλR1/2)q2ER,q,λ2,\begin{split}&\sum_{\kappa\in R\mathbb{Z}^{3}}\bigg{(}{\frac{C\lambda}{R^{1/2}}}|\overline{K}*e_{R,\lambda}(\kappa)|^{2}\bigg{)}^{q/2}=\left(\frac{C\lambda}{R^{1/2}}\right)^{\frac{q}{2}}\sum_{\kappa\in R\mathbb{Z}^{3}}|\overline{K}*e_{R,\lambda}(\kappa)|^{q}\\ &\quad\leq\left(\frac{C\lambda}{R^{1/2}}\right)^{\frac{q}{2}}\|\overline{K}\|_{\ell^{1}}^{q}\|e_{R,\lambda}\|_{\ell^{q}}^{q}\leq\left(\frac{C\lambda}{R^{1/2}}\right)^{\frac{q}{2}}\|e_{R,\lambda}\|_{\ell^{q/2}}^{q}\leq\bigg{(}\frac{C\lambda}{R^{1/2}}\bigg{)}^{\frac{q}{2}}{E_{R,q,\lambda}^{2}},\end{split}

where we used the fact that K¯1(R3)\|\overline{K}\|_{\ell^{1}(R\mathbb{Z}^{3})} is bounded independently of RR. We conclude that, for some constant C21C_{2}\geq 1 independent of q,Rq,R, we have

ER,q,λC2qA0,q(R)q/2+C2q(λ+λR1/2)q/2ER,q,λ+C2q(λR2)q8ER,q,λ3/2+C2q(λR1/2)q2ER,q,λ2\begin{split}E_{R,q,\lambda}&\leq C_{2}^{q}A_{0,q}(R)^{q/2}+C_{2}^{q}\left(\lambda+\frac{\lambda}{R^{1/2}}\right)^{q/2}E_{R,q,\lambda}+C_{2}^{q}\bigg{(}\frac{\lambda}{R^{2}}\bigg{)}^{\frac{q}{8}}E_{R,q,\lambda}^{3/2}\\ &\quad+C_{2}^{q}\bigg{(}\frac{\lambda}{R^{1/2}}\bigg{)}^{\frac{q}{2}}{E_{R,q,\lambda}^{2}}\end{split} (3.26)

The same argument as in [4, p. 2005] shows that ER,q,λE_{R,q,\lambda} is continuous in λ\lambda. Therefore, we conclude from (3.26) and a continuity argument that the estimate (3.18) also holds for 1<q<21<q<2.

We now show (3.19) and (3.20). By the Gagliardo–Nirenberg inequality,

BR|v|103𝑑x(BR|v|2)(BR|v|2)23+R2(BR|v|2)53,\int_{B_{R}}|v|^{\frac{10}{3}}\,dx{\ \lesssim\ }\left(\int_{B_{R}}|\nabla v|^{2}\right)\left(\int_{B_{R}}|v|^{2}\right)^{\frac{2}{3}}+R^{-2}\left(\int_{B_{R}}|v|^{2}\right)^{\frac{5}{3}},
BR|b|103𝑑x(BR|b|2)(BR|b|2)23+R2(BR|b|2)53.\int_{B_{R}}|b|^{\frac{10}{3}}\,dx{\ \lesssim\ }\left(\int_{B_{R}}|\nabla b|^{2}\right)\left(\int_{B_{R}}|b|^{2}\right)^{\frac{2}{3}}+R^{-2}\left(\int_{B_{R}}|b|^{2}\right)^{\frac{5}{3}}.

Denoting N=sup0tλR2BR(|v(t)|2+|b(t)|2)𝑑x+20λR2BR(|v|2+|b|2)𝑑x𝑑sN=\sup_{0\leq t\leq\lambda R^{2}}\int_{B_{R}}\left(|v(t)|^{2}+|b(t)|^{2}\right)dx+2\int_{0}^{\lambda R^{2}}\int_{B_{R}}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxds with λ=λR\lambda=\lambda_{R}, we have

0λR2BR(|v|103+|b|103)𝑑x𝑑sN230λR2(BR|v|2+|b|2)𝑑s+R2N53λR2N53+λN53N53,\begin{split}\int_{0}^{\lambda R^{2}}\int_{B_{R}}\left(|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}\right)dxds&{\ \lesssim\ }N^{\frac{2}{3}}\int_{0}^{\lambda R^{2}}\left(\int_{B_{R}}|\nabla v|^{2}+|\nabla b|^{2}\right)ds+R^{-2}N^{\frac{5}{3}}\lambda R^{2}\\ &{\ \lesssim\ }N^{\frac{5}{3}}+\lambda N^{\frac{5}{3}}{\ \lesssim\ }N^{\frac{5}{3}},\end{split} (3.27)

using λ1\lambda\leq 1. For kR3k\in R\mathbb{Z}^{3} and Q(k)=BR(k)×(0,λRR2)Q(k)=B_{R}(k)\times(0,\lambda_{R}R^{2}), by (3.27) with BRB_{R} replaced by BR(k)B_{R}(k), we have NeR,λ(k)N\leq e_{R,\lambda}(k) and hence

kR3[Q(k)(|v|103+|b|103)𝑑x𝑑t]3q10CkR3((eR,λ(k))53)3q10CE0,q2,\begin{split}\sum_{k\in R\mathbb{Z}^{3}}\left[\int_{Q(k)}\left(|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}\right)dxdt\right]^{\frac{3q}{10}}\leq C\sum_{k\in R\mathbb{Z}^{3}}\left((e_{R,\lambda}(k))^{\frac{5}{3}}\right)^{\frac{3q}{10}}\leq CE_{0},\quad q\geq 2,\end{split}

and, for σ>0\sigma>0 satisfying 5σ3q2\frac{5\sigma}{3}\geq\frac{q}{2} (which is implied if σ35\sigma\geq\frac{3}{5}), we have

kR3[Q(k)(|v|103+|b|103)𝑑x𝑑t]σCkR3(eR,λ(k))5σ3C[kR3(eR,λ(k))q2]2q5σ3CE02q5σ3,1q<2.\begin{split}\sum_{k\in R\mathbb{Z}^{3}}&\left[\int_{Q(k)}\left(|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}\right)dxdt\right]^{\sigma}\\ &\leq C\sum_{k\in R\mathbb{Z}^{3}}\left(e_{R,\lambda}(k)\right)^{\frac{5\sigma}{3}}\leq C\left[\sum_{k\in R\mathbb{Z}^{3}}\left(e_{R,\lambda}(k)\right)^{\frac{q}{2}}\right]^{\frac{2}{q}\cdot\frac{5\sigma}{3}}\leq CE_{0}^{\frac{2}{q}\cdot\frac{5\sigma}{3}},\quad 1\leq q<2.\end{split}

By Calderon–Zygmund estimates,

kR3(Q(k)|π1|53𝑑x𝑑t)3q10CkR3kR3;|kk|<10R[(Q(k)|v|103𝑑x𝑑t)3q10+(Q(k)|b|103𝑑x𝑑t)3q10]CE0,q2,\begin{split}\sum_{k\in R\mathbb{Z}^{3}}&\left(\int_{Q(k)}|\pi_{1}|^{\frac{5}{3}}dxdt\right)^{\frac{3q}{10}}\\ &\leq C\sum_{k\in R\mathbb{Z}^{3}}\sum_{k^{\prime}\in R\mathbb{Z}^{3};\,|k-k^{\prime}|<10R}\left[\left(\int_{Q(k^{\prime})}|v|^{\frac{10}{3}}\,dxdt\right)^{\frac{3q}{10}}+\left(\int_{Q(k^{\prime})}|b|^{\frac{10}{3}}\,dxdt\right)^{\frac{3q}{10}}\right]\\ &\leq CE_{0},\quad q\geq 2,\end{split}

and

kR3(Q(k)|π1|53𝑑x𝑑t)σCkR3kR3;|kk|<10R[(Q(k)|v|103𝑑x𝑑t)σ+(Q(k)|b|103𝑑x𝑑t)σ]CE02q5σ3,1q<2.\begin{split}\sum_{k\in R\mathbb{Z}^{3}}&\left(\int_{Q(k)}|\pi_{1}|^{\frac{5}{3}}dxdt\right)^{\sigma}\\ &\leq C\sum_{k\in R\mathbb{Z}^{3}}\sum_{k^{\prime}\in R\mathbb{Z}^{3};\,|k-k^{\prime}|<10R}\left[\left(\int_{Q(k^{\prime})}|v|^{\frac{10}{3}}\,dxdt\right)^{\sigma}+\left(\int_{Q(k^{\prime})}|b|^{\frac{10}{3}}\,dxdt\right)^{\sigma}\right]\\ &\leq CE_{0}^{\frac{2}{q}\cdot\frac{5\sigma}{3}},\quad 1\leq q<2.\end{split}

For π2\pi_{2}, recall π2\pi_{2} in BR(k)B_{R}(k) is bounded by R3K¯eR,λ(k)R^{-3}\overline{K}*e_{R,\lambda}(k) and hence

Q(k)|π2|53𝑑x𝑑tCλ((K¯eR,λ)(k))53.\int_{Q(k)}|\pi_{2}|^{\frac{5}{3}}dxdt\leq C\lambda\left((\overline{K}*e_{R,\lambda})(k)\right)^{\frac{5}{3}}.

Thus,

kR3(Q(k)|π2|53𝑑x𝑑t)3q10Cλ3q10kR3((K¯eR,λ)(k))q2Cλ3q10K¯1q2kR3(eR,λ(k))q2CE0,q2,\begin{split}\sum_{k\in R\mathbb{Z}^{3}}\left(\int_{Q(k)}|\pi_{2}|^{\frac{5}{3}}dxdt\right)^{\frac{3q}{10}}&\leq C\lambda^{\frac{3q}{10}}\sum_{k\in R\mathbb{Z}^{3}}\left((\overline{K}*e_{R,\lambda})(k)\right)^{\frac{q}{2}}\\ &\leq C\lambda^{\frac{3q}{10}}\left\|\overline{K}\right\|_{\ell^{1}}^{\frac{q}{2}}\sum_{k\in R\mathbb{Z}^{3}}\left(e_{R,\lambda}(k)\right)^{\frac{q}{2}}\leq CE_{0},\quad q\geq 2,\end{split}

and, if 5σ31\frac{5\sigma}{3}\geq 1, which is implied by our assumptions, we have

kR3(Q(k)|π2|53𝑑x𝑑t)σCλσkR3((K¯eR,λ)(k))5σ3CλσK¯15σ3kR3(eR,λ(k))5σ3CE02q5σ3,1q<2.\begin{split}\sum_{k\in R\mathbb{Z}^{3}}\left(\int_{Q(k)}|\pi_{2}|^{\frac{5}{3}}dxdt\right)^{\sigma}&\leq C\lambda^{\sigma}\sum_{k\in R\mathbb{Z}^{3}}\left((\overline{K}*e_{R,\lambda})(k)\right)^{\frac{5\sigma}{3}}\\ &\leq C\lambda^{\sigma}\left\|\overline{K}\right\|_{\ell^{1}}^{\frac{5\sigma}{3}}\sum_{k\in R\mathbb{Z}^{3}}\left(e_{R,\lambda}(k)\right)^{\frac{5\sigma}{3}}\leq CE_{0}^{\frac{2}{q}\cdot\frac{5\sigma}{3}},\quad 1\leq q<2.\end{split}

We conclude that

Q(k)|v|103+|b|103+|π1+π2|53dxdt3q10(kR3)CE0103q=C(A0,q(R))53,q2,\left\|\int_{Q(k)}|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}+|\pi_{1}+\pi_{2}|^{\frac{5}{3}}dxdt\right\|_{\ell^{\frac{3q}{10}}(k\in R\mathbb{Z}^{3})}\leq CE_{0}^{\frac{10}{3q}}=C\left(A_{0,q}(R)\right)^{\frac{5}{3}},\quad q\geq 2,

and, for 5σ31\frac{5\sigma}{3}\geq 1,

Q(k)|v|103+|b|103+|π1+π2|53dxdtσ(kR3)CE0103q=C(A0,q(R))53,1q<2.\left\|\int_{Q(k)}|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}+|\pi_{1}+\pi_{2}|^{\frac{5}{3}}dxdt\right\|_{\ell^{\sigma}(k\in R\mathbb{Z}^{3})}\leq CE_{0}^{\frac{10}{3q}}=C\left(A_{0,q}(R)\right)^{\frac{5}{3}},\quad 1\leq q<2.

This shows (3.19) and (3.20) and completes the proof. ∎

We now prove Theorem 1.9, which follows directly as a simple consequence of Lemma 3.7.

Proof of Theorem 1.9.

The proof of Theorem 1.9 is an adaption of the proof of [4, Theorem 1.4] for the Navier–Stokes equations to the MHD equations.

Observe that, by the assumption R1R\geq 1, the estimate (3.16), and Hölder’s inequality, we have

Nq,R0(v0,b0)CR[k3(|ikR|<Rui2+ai2)q2]2qCR[(k3|ikR|<RuiqR(36q)q2)2q+(k3|ikR|<RaiqR(36q)q2)2q]CR26q(v0Eq22+b0Eq22),\begin{split}N_{q,R}^{0}(v_{0},b_{0})&\leq\frac{C}{R}\left[\sum_{k\in\mathbb{Z}^{3}}\left(\sum_{|i-kR|<R}u_{i}^{2}+a_{i}^{2}\right)^{\frac{q}{2}}\right]^{\frac{2}{q}}\\ &\leq\frac{C}{R}\left[\left(\sum_{k\in\mathbb{Z}^{3}}\sum_{|i-kR|<R}u_{i}^{q}R^{\left(3-\frac{6}{q}\right)\frac{q}{2}}\right)^{\frac{2}{q}}+\left(\sum_{k\in\mathbb{Z}^{3}}\sum_{|i-kR|<R}a_{i}^{q}R^{\left(3-\frac{6}{q}\right)\frac{q}{2}}\right)^{\frac{2}{q}}\right]\\ &\leq CR^{2-\frac{6}{q}}\left(\left\|v_{0}\right\|_{E^{2}_{q}}^{2}+\left\|b_{0}\right\|_{E^{2}_{q}}^{2}\right),\end{split} (3.28)

where ui=v0L2(B1(i))u_{i}=\left\|v_{0}\right\|_{L^{2}(B_{1}(i))} and ai=b0L2(B1(i))a_{i}=\left\|b_{0}\right\|_{L^{2}(B_{1}(i))} for i3i\in\mathbb{Z}^{3}. Now, applying Lemma 3.7, we obtain the bound

esssup0tλRR2BR(x0R)|v|2+|b|22𝑑x+0λRR2BR(x0R)(|v|2+|b|2)𝑑x𝑑tq2(x03)C1A0,q(R).\left\|\mathop{\rm ess\,sup}_{0\leq t\leq\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}\frac{|v|^{2}+|b|^{2}}{2}\,dx+\int_{0}^{\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}\leq C_{1}A_{0,q}(R).

Next, using the definition of λR\lambda_{R} from Lemma 3.7 and the estimate (3.28), we find

λRR2=min(λ0R2,λ0R2(Nq,R0(v0,b0))2)min(λ0R2,λ0R12q2C2(v0Eq24+b0Eq24))λ1Rmin(2,12q2)(1+(v0,b0)Eq2×Eq2)4,\begin{split}\lambda_{R}R^{2}=\min\left(\lambda_{0}R^{2},\,\frac{\lambda_{0}R^{2}}{\left(N_{q,R}^{0}(v_{0},b_{0})\right)^{2}}\right)&\geq\min\left(\lambda_{0}R^{2},\,\frac{\lambda_{0}R^{\frac{12}{q}-2}}{C^{2}\left(\left\|v_{0}\right\|_{E_{q}^{2}}^{4}+\left\|b_{0}\right\|_{E_{q}^{2}}^{4}\right)}\right)\\ &\geq\frac{\lambda_{1}R^{\min\left(2,\,\frac{12}{q}-2\right)}}{\left(1+\left\|(v_{0},b_{0})\right\|_{E_{q}^{2}\times E_{q}^{2}}\right)^{4}},\end{split}

where λ1:=λ0(1+C)2\lambda_{1}:=\lambda_{0}(1+C)^{-2}. Furthermore, from (3.28), we also have A0,q(R)=RNq,R0(v0,b0)CR36q((v0,b0)Eq2×Eq22)A_{0,q}(R)=RN_{q,R}^{0}(v_{0},b_{0})\leq CR^{3-\frac{6}{q}}\left(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}^{2}\right). This yields the desired upper bound in the statement of Theorem 1.9. ∎

Lemma 3.8 (Far-field regularity of local energy solutions with data in Eq2E^{2}_{q}).

Assume v0,b0Eq2v_{0},b_{0}\in E^{2}_{q} for some q1q\geq 1 are divergence free. If (v,b)(v,b) is a local energy solution on 3×(0,T0)\mathbb{R}^{3}\times(0,T_{0}) evolving from (v0,b0)(v_{0},b_{0}) with (v,b)𝐋𝐄q(0,T0)(v,b)\in{\bf LE}_{q}(0,T_{0}). Then

(v,b)(t)Eq4×Eq4 for a.e. t(0,T0].(v,b)(t)\in E^{4}_{q}\times E^{4}_{q}\ \text{ for a.e. }t\in(0,T_{0}].
Proof.

By Lemma 3.7 with R=1R=1, we have

ekq/2(k3)CA0,q,fkq/3(k3)CA0,q3/2,\left\|e_{k}\right\|_{\ell^{q/2}(k\in\mathbb{Z}^{3})}\leq CA_{0,q},\quad\left\|f_{k}\right\|_{\ell^{q/3}(k\in\mathbb{Z}^{3})}\leq CA_{0,q}^{3/2}, (3.29)

where

ek=esssup0tT0B1(k)|v|2+|b|22𝑑x+0T0B1(k)|v|2+|b|2dxdt,e_{k}=\mathop{\rm ess\,sup}_{0\leq t\leq T_{0}}\int_{B_{1}(k)}\frac{|v|^{2}+|b|^{2}}{2}\,dx+\int_{0}^{T_{0}}\int_{B_{1}(k)}|\nabla v|^{2}+|\nabla b|^{2}\,dxdt,
fk=0T0B1(k)|v|3+|b|3+|πck,1(t)|3/2dxdt,f_{k}=\int_{0}^{T_{0}}\int_{B_{1}(k)}|v|^{3}+|b|^{3}+|\pi-c_{k,1}(t)|^{3/2}dxdt,

and

A0,q=B1(k)|v0(x)|2+|b0(x)|2dxq/2(k3)=(v0,b0)Eq2×Eq22.A_{0,q}=\left\|\int_{B_{1}(k)}|v_{0}(x)|^{2}+|b_{0}(x)|^{2}\,dx\right\|_{\ell^{q/2}(k\in\mathbb{Z}^{3})}=\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}^{2}.

Since

limRfkq/3(k3;|k|>R)=0,\lim_{R\to\infty}\left\|f_{k}\right\|_{\ell^{q/3}(k\in\mathbb{Z}^{3};|k|>R)}=0,

by Lemma 3.2, there exists R0>0R_{0}>0 such that

v,bLCloc(BR0c×[T0/2,T0])v,b\in L^{\infty}\cap C_{\mathrm{loc}}(B_{R_{0}}^{c}\times[T_{0}/2,T_{0}])

and

limRvL(BR0c×[T0/2,T0])+bL(BR0c×[T0/2,T0])=0.\lim_{R\to\infty}\left\|v\right\|_{L^{\infty}(B_{R_{0}}^{c}\times[T_{0}/2,T_{0}])}+\left\|b\right\|_{L^{\infty}(B_{R_{0}}^{c}\times[T_{0}/2,T_{0}])}=0.

In fact, for |k|>R0|k|>R_{0},

vL(B1(k)×[T0/2,T0])3+bL(B1(k)×[T0/2,T0])3C|kk|2fk.\left\|v\right\|_{L^{\infty}(B_{1}(k)\times[T_{0}/2,T_{0}])}^{3}+\left\|b\right\|_{L^{\infty}(B_{1}(k)\times[T_{0}/2,T_{0}])}^{3}\leq C\sum_{|k^{\prime}-k|\leq 2}f_{k^{\prime}}.

Thus

vL(B1(k)×[T0/2,T0])+bL(B1(k)×[T0/2,T0])q(k3;|k|>R0)Cfkq/3(k3;|k|>R02)1/31.\left\|\left\|v\right\|_{L^{\infty}(B_{1}(k)\times[T_{0}/2,T_{0}])}+\left\|b\right\|_{L^{\infty}(B_{1}(k)\times[T_{0}/2,T_{0}])}\right\|_{\ell^{q}(k\in\mathbb{Z}^{3};|k|>R_{0})}\leq C\left\|f_{k}\right\|_{\ell^{q/3}(k\in\mathbb{Z}^{3};|k|>R_{0}-2)}^{1/3}\ll 1.

By (3.29) and Sobolev imbedding,

v,bL8/3(0,T0;L4(BR0)).v,b\in L^{8/3}(0,T_{0};L^{4}(B_{R_{0}})).

Thus

(v,b)(t)Eq4×Eq4 for a.e. t(0,T0],(v,b)(t)\in E^{4}_{q}\times E^{4}_{q}\ \text{ for a.e. }t\in(0,T_{0}],

completing the proof of the lemma. ∎

3.3 Global existence

In this section, we prove Theorem 1.10.

For q2q\geq 2, we achieve this by considering the perturbed MHD equations

tvΔv+vv+uv+vubbabba+π=0,tbΔb+vb+ub+vabvavbu=0,v=b=0.\begin{split}\partial_{t}v-\Delta v+v\cdot\nabla v+u\cdot\nabla v+v\cdot\nabla u-b\cdot\nabla b-a\cdot\nabla b-b\cdot\nabla a+\nabla\pi&=0,\\ \partial_{t}b-\Delta b+v\cdot\nabla b+u\cdot\nabla b+v\cdot\nabla a-b\cdot\nabla v-a\cdot\nabla v-b\cdot\nabla u\qquad\ \ &=0,\\ \nabla\cdot v=\nabla\cdot b&=0.\end{split} (3.30)

where uu and aa are given divergence-free vector fields. A local energy solution to the perturbed MHD equations, (3.30), is a weak solution (v,b)(v,b) satisfying Definition 1.7 with the obvious modifications, namely, (v,b)(v,b) and π\pi satisfy the perturbed system as distributions and also satisfy the perturbed local energy inequality.

For 1q<21\leq q<2, we achieve this via the localized and regularized MHD equations:

tvϵΔvϵ+(𝒥ϵ(vϵ))(vϵΦϵ)(𝒥ϵ(bϵ))(bϵΦϵ)+πϵ=0,tbϵΔbϵ+(𝒥ϵ(vϵ))(bϵΦϵ)(𝒥ϵ(bϵ))(vϵΦϵ)=0,vϵ=bϵ=0,\begin{split}\partial_{t}v^{\epsilon}-\Delta v^{\epsilon}+\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\cdot\nabla\right)(v^{\epsilon}\Phi_{\epsilon})-\left(\mathcal{J}_{\epsilon}(b^{\epsilon})\cdot\nabla\right)(b^{\epsilon}\Phi_{\epsilon})+\nabla\pi^{\epsilon}&=0,\\ \partial_{t}b^{\epsilon}-\Delta b^{\epsilon}+\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\cdot\nabla\right)(b^{\epsilon}\Phi_{\epsilon})-\left(\mathcal{J}_{\epsilon}(b^{\epsilon})\cdot\nabla\right)(v^{\epsilon}\Phi_{\epsilon})\qquad\quad&=0,\\ \nabla\cdot v^{\epsilon}=\nabla\cdot b^{\epsilon}&=0,\end{split} (3.31)

where 𝒥ϵ(f)=ηϵf\mathcal{J}_{\epsilon}(f)=\eta_{\epsilon}*f for a spatial mollifier ηϵ(x)=ϵ3η(x/ϵ)\eta_{\epsilon}(x)=\epsilon^{-3}\eta(x/\epsilon) and Φϵ(x)=Φ(ϵx)\Phi_{\epsilon}(x)=\Phi(\epsilon x) for a fixed radially decreasing cutoff function Φ\Phi satisfying Φ=1\Phi=1 on B1(0)B_{1}(0) and supp(Φ)B3/2(0)\mathop{\mathrm{supp}}\nolimits(\Phi)\subset B_{3/2}(0).

3.3.1 The case q2q\geq 2

Lemma 3.9.

Let ϵ(0,1]\epsilon\in(0,1] and δ>0\delta>0 be given, let T0>0T_{0}>0, and let η\eta be a spatial mollifier in 3\mathbb{R}^{3}. Assume that v0,b0L2v_{0},b_{0}\in L^{2} are divergence free and that u,a:3×[0,T0]3u,a:\mathbb{R}^{3}\times[0,T_{0}]\to\mathbb{R}^{3} satisfies u=a=0\nabla\cdot u=\nabla\cdot a=0 and

esssup0<tT0(u,a)(t)Luloc3×Luloc3<δ.\mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left\|(u,a)(t)\right\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}<\delta.

Then there exist Tϵ,δ=min(T0,C(ϵ)((v0,b0)Eq2×Eq2+δ)2)T_{\epsilon,\delta}=\min\left(T_{0},\,C(\epsilon)(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+\delta)^{-2}\right) and a mild solution (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) of the integral equation

vϵ(t)=etΔv0+0te(ts)Δ((ηϵvϵ)vϵ(ηϵbϵ)bϵ)𝑑s+Lt(1)(vϵ,bϵ),bϵ(t)=etΔb0+0te(ts)Δ((ηϵvϵ)bϵ(ηϵbϵ)vϵ)𝑑s+Lt(2)(vϵ,bϵ),\begin{split}v_{\epsilon}(t)&=e^{t\Delta}v_{0}+\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon})\otimes v_{\epsilon}-(\eta_{\epsilon}*b_{\epsilon})\otimes b_{\epsilon}\right)ds+L_{t}^{(1)}(v_{\epsilon},b_{\epsilon}),\\ b_{\epsilon}(t)&=e^{t\Delta}b_{0}+\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon})\otimes b_{\epsilon}-(\eta_{\epsilon}*b_{\epsilon})\otimes v_{\epsilon}\right)ds+L_{t}^{(2)}(v_{\epsilon},b_{\epsilon}),\end{split} (3.32)

for 0<t<Tϵ,δ0<t<T_{\epsilon,\delta}, where

Lt(1)(vϵ,bϵ)=0te(ts)Δ((ηϵu)vϵ+vϵ(ηϵu)(ηϵa)bϵbϵ(ηϵa))ds,L_{t}^{(1)}(v_{\epsilon},b_{\epsilon})=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*u)\otimes v_{\epsilon}+v_{\epsilon}\otimes(\eta_{\epsilon}*u)-(\eta_{\epsilon}*a)\otimes b_{\epsilon}-b_{\epsilon}\otimes(\eta_{\epsilon}*a)\right)ds,
Lt(2)(vϵ,bϵ)=0te(ts)Δ((ηϵu)bϵ+bϵ(ηϵu)(ηϵa)vϵbϵ(ηϵu))ds,L_{t}^{(2)}(v_{\epsilon},b_{\epsilon})=\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*u)\otimes b_{\epsilon}+b_{\epsilon}\otimes(\eta_{\epsilon}*u)-(\eta_{\epsilon}*a)\otimes v_{\epsilon}-b_{\epsilon}\otimes(\eta_{\epsilon}*u)\right)ds,

with (vϵ,bϵ)𝐋𝐄q(0,Tϵ,δ)C([0,Tϵ,δ];L2×L2)(v_{\epsilon},b_{\epsilon})\in{\bf LE}_{q}(0,T_{\epsilon,\delta})\cap C([0,T_{\epsilon,\delta}];L^{2}\times L^{2}), and (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) satisfies

(esssup0<t<Tϵ,δB1(k)(|vϵ(x,t)|2+|bϵ(x,t)|2)dx)12q2C(v0,b0)E2q×E2q and sup0<t<Tϵ,δ(vϵ,bϵ)(t)L2×L22C(v0,b0)L2×L2\begin{split}\left\|\left(\mathop{\rm ess\,sup}_{0<t<T_{\epsilon,\delta}}\int_{B_{1}(k)}\left(|v_{\epsilon}(x,t)|^{2}+|b_{\epsilon}(x,t)|^{2}\right)dx\right)^{\frac{1}{2}}\right\|_{\ell^{q}}&\leq 2C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}\quad\text{ and }\\ \sup_{0<t<T_{\epsilon,\delta}}\left\|(v_{\epsilon},b_{\epsilon})(t)\right\|_{L^{2}\times L^{2}}&\leq 2C\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}\end{split} (3.33)

for a universal constant C>0C>0. This is the unique mild solution of (3.32) in the class (3.33). There exists a pressure πϵ\pi_{\epsilon} so that (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) and πϵ\pi_{\epsilon} solve

tvϵΔvϵ+(ηϵvϵ)vϵ+(ηϵu)vϵ+vϵ(ηϵu)(ηϵbϵ)bϵ(ηϵa)bϵbϵ(ηϵa)+πϵ=0,tbϵΔbϵ+(ηϵvϵ)bϵ+(ηϵu)bϵ+vϵ(ηϵa)(ηϵbϵ)vϵ(ηϵa)vϵbϵ(ηϵu)=0,vϵ=bϵ=0,\begin{split}\partial_{t}v_{\epsilon}-\Delta v_{\epsilon}&+(\eta_{\epsilon}*v_{\epsilon})\cdot\nabla v_{\epsilon}+(\eta_{\epsilon}*u)\cdot\nabla v_{\epsilon}+v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u)\\ &-(\eta_{\epsilon}*b_{\epsilon})\cdot\nabla b_{\epsilon}-(\eta_{\epsilon}*a)\cdot\nabla b_{\epsilon}-b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a)+\nabla\pi_{\epsilon}=0,\\ \partial_{t}b_{\epsilon}-\Delta b_{\epsilon}&+(\eta_{\epsilon}*v_{\epsilon})\cdot\nabla b_{\epsilon}+(\eta_{\epsilon}*u)\cdot\nabla b_{\epsilon}+v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a)\\ &-(\eta_{\epsilon}*b_{\epsilon})\cdot\nabla v_{\epsilon}-(\eta_{\epsilon}*a)\cdot\nabla v_{\epsilon}-b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u)\qquad\ \ =0,\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\,\nabla\cdot v_{\epsilon}=\nabla\cdot b_{\epsilon}=0,\end{split} (3.34)

in the weak sense on 3×(0,Tϵ,δ)\mathbb{R}^{3}\times(0,T_{\epsilon,\delta}). Finally, (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) and πϵ\pi_{\epsilon} are smooth by the interior regularity of the Stokes equations with smooth coefficients.

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 4.1] for the Navier–Stokes equations to the MHD equations.

Note that v0,b0E2qv_{0},b_{0}\in E^{2}_{q} since L2E2qL^{2}\subset E^{2}_{q}. We begin by establishing estimates for the iterates of the Picard scheme. Define the initial iterates as

(vϵ1,bϵ1)=(etΔv0,etΔb0),(v_{\epsilon}^{1},b_{\epsilon}^{1})=(e^{t\Delta}v_{0},e^{t\Delta}b_{0}),

and for n>1n>1, set

vϵn(t)=etΔv0+0te(ts)Δ((ηϵvϵn1)vϵn1(ηϵbϵn1)bϵn1)ds+Lt(1)(vϵn1,bϵn1),bϵn(t)=etΔb0+0te(ts)Δ((ηϵvϵn1)bϵn1(ηϵbϵn1)vϵn1)ds+Lt(2)(vϵn1,bϵn1).\begin{split}v_{\epsilon}^{n}(t)&=e^{t\Delta}v_{0}+\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon}^{n-1})\otimes v_{\epsilon}^{n-1}-(\eta_{\epsilon}*b_{\epsilon}^{n-1})\otimes b_{\epsilon}^{n-1}\right)ds+L_{t}^{(1)}(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1}),\\ b_{\epsilon}^{n}(t)&=e^{t\Delta}b_{0}+\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon}^{n-1})\otimes b_{\epsilon}^{n-1}-(\eta_{\epsilon}*b_{\epsilon}^{n-1})\otimes v_{\epsilon}^{n-1}\right)ds+L_{t}^{(2)}(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1}).\end{split}

For the initial iterates, it follows from the same approach obtaining [4, (4.7)] that

(sup0<t<1B1(k)|vϵ1(x,t)|2dx)12qCv0E2q and (sup0<t<1B1(k)|bϵ1(x,t)|2dx)12qCb0E2q.\begin{split}\left\|\left(\sup_{0<t<1}\int_{B_{1}(k)}\left|v_{\epsilon}^{1}(x,t)\right|^{2}dx\right)^{\frac{1}{2}}\right\|_{\ell^{q}}\leq C\left\|v_{0}\right\|_{E^{2}_{q}}\ \text{ and }\ \left\|\left(\sup_{0<t<1}\int_{B_{1}(k)}\left|b_{\epsilon}^{1}(x,t)\right|^{2}dx\right)^{\frac{1}{2}}\right\|_{\ell^{q}}\leq C\left\|b_{0}\right\|_{E^{2}_{q}}.\end{split} (3.35)

For the nnth iterates, we use the assumption that

(sup0<t<TϵB1(k)(|vϵn1(x,t)|2+|bϵn1(x,t)|2)dx)12q<2C(v0,b0)E2q×E2q.\begin{split}\left\|\left(\sup_{0<t<T_{\epsilon}}\int_{B_{1}(k)}\left(|v_{\epsilon}^{n-1}(x,t)|^{2}+|b_{\epsilon}^{n-1}(x,t)|^{2}\right)dx\right)^{\frac{1}{2}}\right\|_{\ell^{q}}<2C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}.\end{split}

We have

B1(k)|vϵn(x,t)|2dxB1(k)|etΔv0(x)|2dx+I(1)(k)+J(1)(k),B1(k)|bϵn(x,t)|2dxB1(k)|etΔb0(x)|2dx+I(2)(k)+J(2)(k),\begin{split}\int_{B_{1}(k)}\left|v_{\epsilon}^{n}(x,t)\right|^{2}dx&\leq\int_{B_{1}(k)}\left|e^{t\Delta}v_{0}(x)\right|^{2}dx+I^{(1)}(k)+J^{(1)}(k),\\ \int_{B_{1}(k)}\left|b_{\epsilon}^{n}(x,t)\right|^{2}dx&\leq\int_{B_{1}(k)}\left|e^{t\Delta}b_{0}(x)\right|^{2}dx+I^{(2)}(k)+J^{(2)}(k),\end{split} (3.36)

where

I(1)(k)=B1(k)|0te(ts)Δ((ηϵvϵn1)vϵn1(ηϵbϵn1)bϵn1)ds|2dx,J(1)(k)=B1(k)|Lt(1)(vϵn1,bϵn1)|2dx,I(2)(k)=B1(k)|0te(ts)Δ((ηϵvϵn1)bϵn1(ηϵbϵn1)vϵn1)ds|2dx,J(2)(k)=B1(k)|Lt(2)(vϵn1,bϵn1)|2dx.\begin{split}I^{(1)}(k)&=\int_{B_{1}(k)}\left|\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon}^{n-1})\otimes v_{\epsilon}^{n-1}-(\eta_{\epsilon}*b_{\epsilon}^{n-1})\otimes b_{\epsilon}^{n-1}\right)ds\right|^{2}dx,\\ J^{(1)}(k)&=\int_{B_{1}(k)}\left|L_{t}^{(1)}(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1})\right|^{2}dx,\\ I^{(2)}(k)&=\int_{B_{1}(k)}\left|\int_{0}^{t}e^{(t-s)\Delta}\mathbb{P}\nabla\cdot\left((\eta_{\epsilon}*v_{\epsilon}^{n-1})\otimes b_{\epsilon}^{n-1}-(\eta_{\epsilon}*b_{\epsilon}^{n-1})\otimes v_{\epsilon}^{n-1}\right)ds\right|^{2}dx,\\ J^{(2)}(k)&=\int_{B_{1}(k)}\left|L_{t}^{(2)}(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1})\right|^{2}dx.\end{split}

The first two terms on the right-hand side of (3.36) have already been estimated in (3.35). For I(1)(k)I^{(1)}(k) and I(2)(k)I^{(2)}(k), using the same technique deriving [4, (4.11), (4.12)], we get

I(1)(k)+I(2)(k)C(ϵ)t2[(sup0<s<tvϵn1E2q2)((K~un1)(k))2+(sup0<s<tbϵn1E2q2)((K~an1)(k))2]+C(ϵ)t[sup0<s<tvϵn1E2q2sup0<s<t|kk|<8vϵn1L2(B1(k))2+sup0<s<tbϵn1E2q2sup0<s<t|kk|<8bϵn1L2(B1(k))2],\begin{split}I^{(1)}(k)&+I^{(2)}(k)\\ &\leq C(\epsilon)t^{2}\left[\left(\sup_{0<s<t}\left\|v_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\right)\left((\widetilde{K}*u^{n-1})(k)\right)^{2}+\left(\sup_{0<s<t}\left\|b_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\right)\left((\widetilde{K}*a^{n-1})(k)\right)^{2}\right]\\ &\quad+C(\epsilon)t\left[\sup_{0<s<t}\left\|v_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\sup_{0<s<t}\sum_{|k-k^{\prime}|<8}\left\|v_{\epsilon}^{n-1}\right\|_{L^{2}(B_{1}(k^{\prime}))}^{2}\right.\\ &\qquad\qquad\qquad\left.+\sup_{0<s<t}\left\|b_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\sup_{0<s<t}\sum_{|k-k^{\prime}|<8}\left\|b_{\epsilon}^{n-1}\right\|_{L^{2}(B_{1}(k^{\prime}))}^{2}\right],\end{split} (3.37)

where

ukn1=(sup0<s<tB1(k)|vϵn1(y)|2dy)1/2,akn1=(sup0<s<tB1(k)|bϵn1(y)|2dy)1/2,u_{k}^{n-1}=\left(\sup_{0<s<t}\int_{B_{1}(k)}\left|v_{\epsilon}^{n-1}(y)\right|^{2}dy\right)^{1/2},\qquad a_{k}^{n-1}=\left(\sup_{0<s<t}\int_{B_{1}(k)}\left|b_{\epsilon}^{n-1}(y)\right|^{2}dy\right)^{1/2},

and

K~(k)=|k|4 if |k|4,K~(k)=0 otherwise.\widetilde{K}(k)=|k|^{-4}\ \text{ if }|k|\geq 4,\qquad\widetilde{K}(k)=0\ \text{ otherwise.} (3.38)

We next estimate the terms J(1)(k)J^{(1)}(k) and J(2)(k)J^{(2)}(k). Using the same argument for [4, (4.13), (4.14)] yields

J(1)(k)+J(2)(k)Cδ2t2((K~un1)(k)+(K~an1)(k))2+C(ϵ)tδ2(vϵn1L2(B4(k))2+bϵn1L2(B4(k))2).\begin{split}J^{(1)}(k)+J^{(2)}(k)&\leq C\delta^{2}t^{2}\left((\widetilde{K}*u^{n-1})(k)+(\widetilde{K}*a^{n-1})(k)\right)^{2}\\ &\quad+C(\epsilon)t\delta^{2}\left(\left\|v_{\epsilon}^{n-1}\right\|_{L^{2}(B_{4}(k))}^{2}+\left\|b_{\epsilon}^{n-1}\right\|_{L^{2}(B_{4}(k))}^{2}\right).\end{split} (3.39)

Combining (3.35), (3.37), and (3.39), we have

B1(k)(|vϵn(x,t)|2+|bϵn(x,t)|2)dxC|kk|4B1(k)(|v0(x)|2+|b0(x)|2)dx+C[(K~(u+a))(k)]2+C(ϵ)t2[(sup0<s<tvϵn1E2q2)((K~un1)(k))2+(sup0<s<tbϵn1E2q2)((K~an1)(k))2]+C(ϵ)t[sup0<s<tvϵn1E2q2sup0<s<t|kk|<8vϵn1L2(B1(k))2+sup0<s<tbϵn1E2q2sup0<s<t|kk|<8bϵn1L2(B1(k))2]+Cδ2t2((K~un1)(k)+(K~an1)(k))2+C(ϵ)tδ2(vϵn1L2(B4(k))2+bϵn1L2(B4(k))2).\begin{split}&\int_{B_{1}(k)}\left(\left|v_{\epsilon}^{n}(x,t)\right|^{2}+\left|b_{\epsilon}^{n}(x,t)\right|^{2}\right)dx\\ &\ \leq C\sum_{|k-k^{\prime}|\leq 4}\int_{B_{1}(k^{\prime})}\left(|v_{0}(x)|^{2}+|b_{0}(x)|^{2}\right)dx+C\left[\left(\widetilde{K}*(u+a)\right)(k)\right]^{2}\\ &\ \quad+C(\epsilon)t^{2}\left[\left(\sup_{0<s<t}\left\|v_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\right)\left((\widetilde{K}*u^{n-1})(k)\right)^{2}+\left(\sup_{0<s<t}\left\|b_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\right)\left((\widetilde{K}*a^{n-1})(k)\right)^{2}\right]\\ &\ \quad+C(\epsilon)t\left[\sup_{0<s<t}\left\|v_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\sup_{0<s<t}\sum_{|k-k^{\prime}|<8}\left\|v_{\epsilon}^{n-1}\right\|_{L^{2}(B_{1}(k^{\prime}))}^{2}\right.\\ &\ \qquad\qquad\qquad\left.+\sup_{0<s<t}\left\|b_{\epsilon}^{n-1}\right\|_{E^{2}_{q}}^{2}\sup_{0<s<t}\sum_{|k-k^{\prime}|<8}\left\|b_{\epsilon}^{n-1}\right\|_{L^{2}(B_{1}(k^{\prime}))}^{2}\right]\\ &\ \quad+C\delta^{2}t^{2}\left((\widetilde{K}*u^{n-1})(k)+(\widetilde{K}*a^{n-1})(k)\right)^{2}+C(\epsilon)t\delta^{2}\left(\left\|v_{\epsilon}^{n-1}\right\|_{L^{2}(B_{4}(k))}^{2}+\left\|b_{\epsilon}^{n-1}\right\|_{L^{2}(B_{4}(k))}^{2}\right).\end{split} (3.40)

Taking the supremum in time of the left-hand side of (LABEL:eq-4.15-BT-SIMA2021), applying the q2\ell^{\frac{q}{2}} norm, using Young’s convolution inequality, and raising everything to the 1/21/2 power yields

(vϵn,bϵn)𝐋𝐄q(0,t)C(v0,b0)E2q×E2q+C(ϵ)t12(vϵn1,bϵn1)𝐋𝐄q(0,t)2+C(ϵ)δt12(vϵn1,bϵn1)𝐋𝐄q(0,t),\begin{split}\left\|(v_{\epsilon}^{n},b_{\epsilon}^{n})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}&\leq C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+C(\epsilon)t^{\frac{1}{2}}\left\|(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}^{2}\\ &\quad+C(\epsilon)\delta t^{\frac{1}{2}}\left\|(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1})\right\|_{{\bf LE}_{q}^{\flat}(0,t)},\end{split} (3.41)

where 𝐋𝐄q(I)\left\|\,\cdot\,\right\|_{{\bf LE}_{q}^{\flat}(I)} is the first part of the norm 𝐋𝐄q(I)\left\|\,\cdot\,\right\|_{{\bf LE}_{q}(I)} defined by (v,b)𝐋𝐄q(I)=(v,b)E,2T,q×E,2T,q\left\|(v,b)\right\|_{{\bf LE}_{q}^{\flat}(I)}=\left\|(v,b)\right\|_{E^{\infty,2}_{T,q}\times E^{\infty,2}_{T,q}}. So, if tt is small as determined by C(ϵ)C(\epsilon), ϵ\epsilon, and (v0,b0)E2q×E2q\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}} (but independently of nn), tC(ϵ)/((v0,b0)E2q×E2q2+δ2)t\leq C(\epsilon)/\left(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}^{2}+\delta^{2}\right), then the right-hand side of (3.41) is controlled by 2C(v0,b0)E2q×E2q2C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}. This establishes a uniform-in-nn bound for (vϵn,bϵn)(v_{\epsilon}^{n},b_{\epsilon}^{n}).

These uniform bounds and the estimation methods above allow us to show the difference estimate

(vϵn+1,bϵn+1)(vϵn,bϵn)𝐋𝐄q(0,t)C(ϵ)t((v0,b0)E2q×E2q+δ)(vϵn,bϵn)(vϵn1,bϵn1)𝐋𝐄q(0,t).\left\|(v_{\epsilon}^{n+1},b_{\epsilon}^{n+1})-(v_{\epsilon}^{n},b_{\epsilon}^{n})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}\leq C(\epsilon)\sqrt{t}\left(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+\delta\right)\left\|(v_{\epsilon}^{n},b_{\epsilon}^{n})-(v_{\epsilon}^{n-1},b_{\epsilon}^{n-1})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}.

Thus, if tt is sufficiently small, then (vϵn,bϵn)(v_{\epsilon}^{n},b_{\epsilon}^{n}) is a Cauchy sequence in 𝐋𝐄q(0,t){\bf LE}_{q}^{\flat}(0,t) norm and converges to a limit (vϵn,bϵn)(v_{\epsilon}^{n},b_{\epsilon}^{n}) in the sense that

(vϵn,bϵn)(vϵ,bϵ)𝐋𝐄q(0,t)0, as n.\left\|(v_{\epsilon}^{n},b_{\epsilon}^{n})-(v_{\epsilon},b_{\epsilon})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}\to 0,\quad\text{ as }n\to\infty.

This convergence implies (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) satisfies (3.32) and (3.33).

Uniqueness in the class (3.33) follows from the same difference estimates as before. Suppose (v1,b1)(v_{1},b_{1}) and (v2,b2)(v_{2},b_{2}) are two mild solutions of (3.32) satisfying (3.33). Then

(v1,b1)(v2,b2)𝐋𝐄q(0,t)C(ϵ)t((v0,b0)E2q×E2q+δ)(v1,b1)(v2,b2)𝐋𝐄q(0,t).\left\|(v_{1},b_{1})-(v_{2},b_{2})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}\leq C(\epsilon)\sqrt{t}(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+\delta)\left\|(v_{1},b_{1})-(v_{2},b_{2})\right\|_{{\bf LE}_{q}^{\flat}(0,t)}.

Hence, if t>0t>0 is sufficiently small, we conclude that (v1,b1)=(v2,b2)(v_{1},b_{1})=(v_{2},b_{2}).

We now recover a pressure associated to (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}). It is known that (vϵ,bϵ)L(0,Tϵ,δ;L2×L2)(v_{\epsilon},b_{\epsilon})\in L^{\infty}(0,T_{\epsilon,\delta};L^{2}\times L^{2}) for some Tϵ,δ>0T_{\epsilon,\delta}>0, with (vϵ,bϵ)L(0,Tϵ,δ;L2×L2)2C(v0,b0)L2×L2\left\|(v_{\epsilon},b_{\epsilon})\right\|_{L^{\infty}(0,T_{\epsilon,\delta};L^{2}\times L^{2})}\leq 2C\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}. Therefore, the following nonlinear terms belongs to L(0,Tϵ,δ;L2)L^{\infty}(0,T_{\epsilon,\delta};L^{2}):

ηϵvϵvϵ+ηϵuvϵ+vϵηϵuηϵbϵbηϵabϵbϵηϵa.\eta_{\epsilon}*v_{\epsilon}\otimes v_{\epsilon}+\eta_{\epsilon}*u\otimes v_{\epsilon}+v_{\epsilon}\otimes\eta_{\epsilon}*u-\eta_{\epsilon}*b_{\epsilon}\otimes b-\eta_{\epsilon}*a\otimes b_{\epsilon}-b_{\epsilon}\otimes\eta_{\epsilon}*a.

Consequently πϵ=(Δ)1ij(ηϵvϵvϵ+ηϵuvϵ+vϵηϵuηϵbϵbηϵabϵbϵηϵa)\pi_{\epsilon}=(-\Delta)^{-1}\partial_{i}\partial_{j}(\eta_{\epsilon}*v_{\epsilon}\otimes v_{\epsilon}+\eta_{\epsilon}*u\otimes v_{\epsilon}+v_{\epsilon}\otimes\eta_{\epsilon}*u-\eta_{\epsilon}*b_{\epsilon}\otimes b-\eta_{\epsilon}*a\otimes b_{\epsilon}-b_{\epsilon}\otimes\eta_{\epsilon}*a) is well-defined. It follows that vϵetΔv0v_{\epsilon}-e^{t\Delta}v_{0} solves the Stokes system with pressure πϵ\pi_{\epsilon} and forcing term equal to (ηϵvϵvϵ+ηϵuvϵ+vϵηϵuηϵbϵbηϵabϵbϵηϵa)\nabla\cdot(\eta_{\epsilon}*v_{\epsilon}\otimes v_{\epsilon}+\eta_{\epsilon}*u\otimes v_{\epsilon}+v_{\epsilon}\otimes\eta_{\epsilon}*u-\eta_{\epsilon}*b_{\epsilon}\otimes b-\eta_{\epsilon}*a\otimes b_{\epsilon}-b_{\epsilon}\otimes\eta_{\epsilon}*a). Adding back the linear term etΔv0e^{t\Delta}v_{0}, we see that (vϵ,πϵ)(v_{\epsilon},\pi_{\epsilon}) solve the perturbed, regularized Navier–Stokes equations. The local pressure expansion (1.27) follows from the definition of πϵ\pi_{\epsilon}.

We now establish the estimate

0tB1(k)(|vϵ|2+|bϵ|2)dxdsq/2(k)<.\left\|\int_{0}^{t}\int_{B_{1}(k)}\left(|\nabla v_{\epsilon}|^{2}+|\nabla b_{\epsilon}|^{2}\right)dxds\right\|_{\ell^{q/2}(k)}<\infty. (3.42)

This follow from the local energy equality satisfied by (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) and the associated pressure πϵ\pi_{\epsilon}, valid for tTϵ,δt\leq T_{\epsilon,\delta} due to the regularity of the solutions due to smoothness and convergence to the data in L2locL^{2}_{\mathrm{loc}}:

B1(k)(|vϵ|2+|bϵ|2)(x,t)ϕ(xk)dx+20t(|vϵ|2+|bϵ|2)ϕ(xk)dxds=(|v0|2+|b0|2)ϕ(xk)dx+0t(|vϵ|2+|bϵ|2)Δϕ(xk)dxds+0t(|vϵ|2+|bϵ|2)(ηϵvϵ+ηϵu)ϕ(xk)dxds20t(vϵ(ηϵu))vϵϕ(xk)dxds+20t(bϵ(ηϵa))vϵϕ(xk)dxds20t(vϵ(ηϵa))bϵϕ(xk)dxds+20t(bϵ(ηϵu))bϵϕ(xk)dxds+20tπϵ(vϵϕ(xk))dxds20t(vϵbϵ)(ηϵbϵ+ηϵa)ϕ(xk)dxds.\begin{split}&\int_{B_{1}(k)}\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(x,t)\phi(x-k)\,dx+2\int_{0}^{t}\int\left(|\nabla v_{\epsilon}|^{2}+|\nabla b_{\epsilon}|^{2}\right)\phi(x-k)\,dxds\\ &\quad=\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi(x-k)\,dx+\int_{0}^{t}\int\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)\Delta\phi(x-k)\,dxds\\ &\quad\quad+\int_{0}^{t}\int\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(\eta_{\epsilon}*v_{\epsilon}+\eta_{\epsilon}*u)\cdot\nabla\phi(x-k)\,dxds\\ &\quad\quad-2\int_{0}^{t}\int(v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u))\cdot v_{\epsilon}\phi(x-k)\,dxds+2\int_{0}^{t}\int(b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a))\cdot v_{\epsilon}\phi(x-k)\,dxds\\ &\quad\quad-2\int_{0}^{t}\int(v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a))\cdot b_{\epsilon}\phi(x-k)\,dxds+2\int_{0}^{t}\int(b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u))\cdot b_{\epsilon}\phi(x-k)\,dxds\\ &\quad\quad+2\int_{0}^{t}\int\pi_{\epsilon}(v_{\epsilon}\cdot\nabla\phi(x-k))\,dxds-2\int_{0}^{t}\int(v_{\epsilon}\cdot b_{\epsilon})(\eta_{\epsilon}*b_{\epsilon}+\eta_{\epsilon}*a)\cdot\nabla\phi(x-k)\,dxds.\end{split} (3.43)

To estimate the nonlinear terms, we use the bound ηϵfL(B2(k))C(ϵ)fL2uloc\left\|\eta_{\epsilon}*f\right\|_{L^{\infty}(B_{2}(k))}\leq C(\epsilon)\left\|f\right\|_{L^{2}_{\mathrm{uloc}}} for ϵ1\epsilon\leq 1. Hence, we have

0t(|vϵ|2+|bϵ|2)(ηϵvϵ)ϕ(xk)dxdsC(ϵ)vϵL2L2ulocesssup0<s<tkkB1(k)(|vϵ|2+|bϵ|2)dx,\begin{split}\int_{0}^{t}\int&\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(\eta_{\epsilon}*v_{\epsilon})\cdot\nabla\phi(x-k)\,dxds\\ &\leq C(\epsilon)\left\|v_{\epsilon}\right\|_{L^{2}\infty L^{2}_{\mathrm{uloc}}}\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)dx,\end{split} (3.44)

and

0t(vϵbϵ)(ηϵbϵ)ϕ(xk)dxds0t(|vϵ|2+|bϵ|2)(ηϵbϵ)ϕ(xk)dxdsC(ϵ)bϵL2L2ulocesssup0<s<tkkB1(k)(|vϵ|2+|bϵ|2)dx.\begin{split}\int_{0}^{t}\int&(v_{\epsilon}\cdot b_{\epsilon})(\eta_{\epsilon}*b_{\epsilon})\cdot\nabla\phi(x-k)\,dxds\\ &{\ \lesssim\ }\int_{0}^{t}\int\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(\eta_{\epsilon}*b_{\epsilon})\cdot\nabla\phi(x-k)\,dxds\\ &\leq C(\epsilon)\left\|b_{\epsilon}\right\|_{L^{2}\infty L^{2}_{\mathrm{uloc}}}\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)dx.\end{split} (3.45)

From the assumptions on uu and aa, we also have

ηϵuL+(ηϵu)LC(ϵ)uL(0,T0;L3uloc)C(ϵ)δ,\left\|\eta_{\epsilon}*u\right\|_{L^{\infty}}+\left\|\nabla(\eta_{\epsilon}*u)\right\|_{L^{\infty}}\leq C(\epsilon)\left\|u\right\|_{L^{\infty}(0,T_{0};L^{3}_{\mathrm{uloc}})}\leq C(\epsilon)\delta,

and

ηϵaL+(ηϵa)LC(ϵ)aL(0,T0;L3uloc)C(ϵ)δ.\left\|\eta_{\epsilon}*a\right\|_{L^{\infty}}+\left\|\nabla(\eta_{\epsilon}*a)\right\|_{L^{\infty}}\leq C(\epsilon)\left\|a\right\|_{L^{\infty}(0,T_{0};L^{3}_{\mathrm{uloc}})}\leq C(\epsilon)\delta.

These imply the following estimates:

0t(|vϵ|2+|bϵ|2)(ηϵu)ϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)(|vϵ(x,s)|2+|bϵ(x,s)|2)dx,\begin{split}\int_{0}^{t}\int&\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(\eta_{\epsilon}*u)\cdot\nabla\phi(x-k)\,dxds\\ &\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|v_{\epsilon}(x,s)|^{2}+|b_{\epsilon}(x,s)|^{2}\right)dx,\end{split} (3.46)
0t(vϵbϵ)(ηϵa)ϕ(xk)dxds0t(vϵ|2+|bϵ|2)(ηϵa)ϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)(|vϵ(x,s)|2+|bϵ(x,s)|2)dx,\begin{split}\int_{0}^{t}\int&(v_{\epsilon}\cdot b_{\epsilon})(\eta_{\epsilon}*a)\cdot\nabla\phi(x-k)\,dxds\\ &{\ \lesssim\ }\int_{0}^{t}\int\left(v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)(\eta_{\epsilon}*a)\cdot\nabla\phi(x-k)\,dxds\\ &\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|v_{\epsilon}(x,s)|^{2}+|b_{\epsilon}(x,s)|^{2}\right)dx,\end{split} (3.47)
0t(vϵ(ηϵu))vϵϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)|vϵ|2dx,\begin{split}\int_{0}^{t}\int(v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u))\cdot v_{\epsilon}\phi(x-k)\,dxds\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}|v_{\epsilon}|^{2}\,dx,\end{split} (3.48)
0t(bϵ(ηϵa))vϵϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)|bϵ||vϵ|dxC(ϵ)δesssup0<s<tkkB1(k)(|bϵ|2+|vϵ|2)dx,\begin{split}\int_{0}^{t}\int(b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a))\cdot v_{\epsilon}\phi(x-k)\,dxds&\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}|b_{\epsilon}||v_{\epsilon}|\,dx\\ &\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|b_{\epsilon}|^{2}+|v_{\epsilon}|^{2}\right)dx,\end{split} (3.49)
0t(vϵ(ηϵa))bϵϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)|vϵ||bϵ|dxC(ϵ)δesssup0<s<tkkB1(k)(|vϵ|2+|bϵ|2)dx,\begin{split}\int_{0}^{t}\int(v_{\epsilon}\cdot\nabla(\eta_{\epsilon}*a))\cdot b_{\epsilon}\phi(x-k)\,dxds&\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}|v_{\epsilon}||b_{\epsilon}|\,dx\\ &\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)dx,\end{split} (3.50)

and

0t(bϵ(ηϵu))bϵϕ(xk)dxdsC(ϵ)δesssup0<s<tkkB1(k)|bϵ|2dx.\begin{split}\int_{0}^{t}\int(b_{\epsilon}\cdot\nabla(\eta_{\epsilon}*u))\cdot b_{\epsilon}\phi(x-k)\,dxds\leq C(\epsilon)\delta\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\sim k}\int_{B_{1}(k^{\prime})}|b_{\epsilon}|^{2}\,dx.\end{split} (3.51)

The pressure satisfies the local pressure expansion (1.27), which allows us–after incorporating an additive constant–to express it as a sum of two components: πϵ(x,t)+c=πϵ,near+πϵ,far\pi_{\epsilon}(x,t)+c=\pi_{\epsilon,{\rm near}}+\pi_{\epsilon,{\rm far}}, where πϵ,near\pi_{\epsilon,{\rm near}} is a Calderon–Zygmund operator applied to a localized term, and πϵ,far\pi_{\epsilon,{\rm far}} is a nonsingular integral operator acting on data supported away from the ball B2(k)B_{2}(k). Due to the structure of the pressure term in the local energy inequality, the additive constant cc plays no role and may be disregarded. By applying the Calderon–Zygmund inequality, the contribution from πϵ,near\pi_{\epsilon,{\rm near}} to the local energy inequality can be estimated in the same way as the nonlinear and perturbative terms discussed earlier. Specifically, it is controlled by the right-hand sides of estimates (3.44) through (3.51). We are thus left to estimate only the far-filed component πϵ,far\pi_{\epsilon,{\rm far}}. In B2(k)×(0,T0)B_{2}(k)\times(0,T_{0}),

|πϵ,far|Ck3;|kk|>41|kk|4B2(k)(|vϵ||ηϵvϵ|+|ηϵu||vϵ|+|bϵ||ηϵbϵ|+|ηϵa||bϵ|)dyC(ϵ)k3;|kk|>41|kk|4(vϵL2(B3(k))2+vϵL2(B3(k))uL(0,T0;L3uloc)+bϵL2(B3(k))2+bϵL2(B3(k))aL(0,T0;L3uloc)).\begin{split}|\pi_{\epsilon,{\rm far}}|&\leq C\sum_{k^{\prime}\in\mathbb{Z}^{3};|k^{\prime}-k|>4}\frac{1}{|k-k^{\prime}|^{4}}\int_{B_{2}(k^{\prime})}\left(|v_{\epsilon}||\eta_{\epsilon}*v_{\epsilon}|+|\eta_{\epsilon}*u||v_{\epsilon}|+|b_{\epsilon}||\eta_{\epsilon}*b_{\epsilon}|+|\eta_{\epsilon}*a||b_{\epsilon}|\right)dy\\ &\leq C(\epsilon)\sum_{k^{\prime}\in\mathbb{Z}^{3};|k^{\prime}-k|>4}\frac{1}{|k-k^{\prime}|^{4}}\left(\left\|v_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}^{2}+\left\|v_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}\left\|u\right\|_{L^{\infty}(0,T_{0};L^{3}_{\mathrm{uloc}})}\right.\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\left.+\left\|b_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}^{2}+\left\|b_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}\left\|a\right\|_{L^{\infty}(0,T_{0};L^{3}_{\mathrm{uloc}})}\right).\end{split}

Therefore, the contribution of πϵ,far\pi_{\epsilon,{\rm far}} to the local energy equation satisfies

0tB2(k)πϵ,far(vϵϕ(xk))dxdsC(ϵ)T0vϵLL2ulocesssup0<s<tk3;|kk|>41|kk|4B3(k)(|vϵ|2+|bϵ|2)dy+C(ϵ)δT0vϵLL2(B2(k))esssup0<s<tk3;|kk|>41|kk|4(vϵL2(B3(k))+bϵL2(B3(k))).\begin{split}\int_{0}^{t}&\int_{B_{2}(k)}\pi_{\epsilon,{\rm far}}(v_{\epsilon}\cdot\nabla\phi(x-k))\,dxds\\ &\leq C(\epsilon)T_{0}\left\|v_{\epsilon}\right\|_{L^{\infty}L^{2}_{\mathrm{uloc}}}\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\in\mathbb{Z}^{3};|k^{\prime}-k|>4}\frac{1}{|k-k^{\prime}|^{4}}\int_{B_{3}(k^{\prime})}\left(|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}\right)dy\\ &\quad+C(\epsilon)\delta T_{0}\left\|v_{\epsilon}\right\|_{L^{\infty}L^{2}(B_{2}(k))}\mathop{\rm ess\,sup}_{0<s<t}\sum_{k^{\prime}\in\mathbb{Z}^{3};|k^{\prime}-k|>4}\frac{1}{|k-k^{\prime}|^{4}}\left(\left\|v_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}+\left\|b_{\epsilon}\right\|_{L^{2}(B_{3}(k^{\prime}))}\right).\end{split} (3.52)

Taking the essential supremum in tt, raising both sides of (3.52) to the power q/2q/2, and summing over k3k\in\mathbb{Z}^{3}, we apply Hölder’s and Young’s inequalities to control the far-field pressure term. This establishes the estimate (3.42), completing the proof of Lemma 3.9. ∎

Lemma 3.10.

Assume that v0,b0E2qv_{0},b_{0}\in E^{2}_{q}, for some 2q<2\leq q<\infty, are divergence free. There exists a small universal constant c0c_{0} so that for all δ(0,c0]\delta\in(0,c_{0}] and for all divergence free vector fields u,a:3×[0,T0]3u,a:\mathbb{R}^{3}\times[0,T_{0}]\to\mathbb{R}^{3} satisfying

esssup0<tT0(u,a)(t)L3uloc×L3uloc<δ\mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left\|(u,a)(t)\right\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}<\delta

for some T0>0T_{0}>0 and if, additionally, a given local energy solution (v,b)(v,b) to the perturbed MHD equations, (3.30), satisfies

esssup0tT0B1(x0)(|v|2+|b|2)dx+0T0B1(x0)(|v|2+|b|2)dxdtq2(x03)<,\begin{split}\left\|\mathop{\rm ess\,sup}_{0\leq t\leq T_{0}}\int_{B_{1}(x_{0})}\left(|v|^{2}+|b|^{2}\right)dx+\int_{0}^{T_{0}}\int_{B_{1}(x_{0})}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}<\infty,\end{split}

then there are positive universal constants C1C_{1} and λ0<1\lambda_{0}<1 such that

esssup0tλB1(x0)|v|2+|b|22dx+0λB1(x0)(|v|2+|b|2)dxdtq2(x03)<C1A0,q,\begin{split}\left\|\mathop{\rm ess\,sup}_{0\leq t\leq\lambda}\int_{B_{1}(x_{0})}\frac{|v|^{2}+|b|^{2}}{2}dx+\int_{0}^{\lambda}\int_{B_{1}(x_{0})}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}<C_{1}A_{0,q},\end{split}

where

A0,q=B1(x0)(|v0|2+|b0|2)dxq2(x03),λ=min(T0,λ0,λ0A0,q2).A_{0,q}=\left\|\int_{B_{1}(x_{0})}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})},\quad\lambda=\min\left(T_{0},\,\lambda_{0},\,\frac{\lambda_{0}}{A_{0,q}^{2}}\right).

Consequently,

0λB1(x0)|v|103+|b|103+|πcx0(t)|53dxdt3q10(x03)CA0,q53.\begin{split}\left\|\int_{0}^{\lambda}\int_{B_{1}(x_{0})}|v|^{\frac{10}{3}}+|b|^{\frac{10}{3}}+\left|\pi-c_{x_{0}}(t)\right|^{\frac{5}{3}}dxdt\right\|_{\ell^{\frac{3q}{10}}(x_{0}\in\mathbb{Z}^{3})}\leq CA_{0,q}^{\frac{5}{3}}.\end{split}
Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 4.2] for the Navier–Stokes equations to the MHD equations.

Once the perturbation terms in the local energy inequality for (v,b)(v,b) are estimated, the proof proceeds identically to that of Lemma 3.7 with R=1R=1 and λR=λ\lambda_{R}=\lambda.

The linear terms in the perturbed local energy inequality can be estimated as follows:

0λ(uv+vuabba)(ϕ(xκ)v)dxdt=0λ[uv(ϕ(xκ)v)vu:(ϕ(xκ)v)ab(ϕ(xκ)v)+ba:(ϕ(xκ)v)]dxdtC0λB2(κ)[|u|(|v|2+|v||b|)+|a|(|v||b|+|v||b|+|b||v|)]dxdtCuLL3uloc0λvL6(B2(κ))(vL2(B2(κ))+vL2(B2(κ)))dt+CaLL3uloc0λvL6(B2(κ))(bL2(B2(κ))+bL2(B2(κ)))dt+CaLL3uloc0λbL6(B2(κ))vL2(B2(κ))dtCλδesssup0<t<λκκB1(κ)(|v(x,t)|2+|b(x,t)|2)dx+Cδesssup0<t<λκκ0λB1(κ)(|v(x,t)|2+|b(x,t)|2)dxdt,\begin{split}\int_{0}^{\lambda}&\int\left(u\cdot\nabla v+v\cdot\nabla u-a\cdot\nabla b-b\cdot\nabla a\right)\cdot\left(\phi(x-\kappa)v\right)dxdt\\ &=\int_{0}^{\lambda}\int\left[u\cdot\nabla v\cdot(\phi(x-\kappa)v)-v\otimes u:\nabla(\phi(x-\kappa)v)\right.\\ &\qquad\qquad\ \left.-a\cdot\nabla b\cdot(\phi(x-\kappa)v)+b\otimes a:\nabla(\phi(x-\kappa)v)\right]dxdt\\ &\leq C\int_{0}^{\lambda}\int_{B_{2}(\kappa)}\left[|u|(|v|^{2}+|v||\nabla b|)+|a|(|v||b|+|v||\nabla b|+|b||\nabla v|)\right]dxdt\\ &\leq C\left\|u\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|v\right\|_{L^{6}(B_{2}(\kappa))}\left(\left\|v\right\|_{L^{2}(B_{2}(\kappa))}+\left\|\nabla v\right\|_{L^{2}(B_{2}(\kappa))}\right)dt\\ &\quad+C\left\|a\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|v\right\|_{L^{6}(B_{2}(\kappa))}\left(\left\|b\right\|_{L^{2}(B_{2}(\kappa))}+\left\|\nabla b\right\|_{L^{2}(B_{2}(\kappa))}\right)dt\\ &\quad+C\left\|a\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|b\right\|_{L^{6}(B_{2}(\kappa))}\left\|\nabla v\right\|_{L^{2}(B_{2}(\kappa))}dt\\ &\leq C\lambda\delta\mathop{\rm ess\,sup}_{0<t<\lambda}\sum_{\kappa^{\prime}\sim\kappa}\int_{B_{1}(\kappa^{\prime})}\left(|v(x,t)|^{2}+|b(x,t)|^{2}\right)dx\\ &\quad+C\delta\mathop{\rm ess\,sup}_{0<t<\lambda}\sum_{\kappa^{\prime}\sim\kappa}\int_{0}^{\lambda}\int_{B_{1}(\kappa^{\prime})}\left(|\nabla v(x,t)|^{2}+|\nabla b(x,t)|^{2}\right)dxdt,\end{split}

and

0λ(ub+vaavbu)(ϕ(xκ)b)dxdt=0λ[ub(ϕ(xκ)b)va:(ϕ(xκ)b)av(ϕ(xκ)b)+bu:(ϕ(xκ)b)]dxdtC0λB2(κ)[|u|(|b|2+|b||b|)+|a|(|v||b|+|v||b|+|b||v|)]dxdtCuLL3uloc0λbL6(B2(κ))(bL2(B2(κ))+bL2(B2(κ)))dt+CaLL3uloc0λvL6(B2(κ))(bL2(B2(κ))+bL2(B2(κ)))dt+CaLL3uloc0λbL6(B2(κ))vL2(B2(κ))dtCλδesssup0<t<λκκB1(κ)(|v(x,t)|2+|b(x,t)|2)dx+Cδesssup0<t<λκκ0λB1(κ)(|v(x,t)|2+|b(x,t)|2)dxdt.\begin{split}\int_{0}^{\lambda}&\int\left(u\cdot\nabla b+v\cdot\nabla a-a\cdot\nabla v-b\cdot\nabla u\right)\cdot\left(\phi(x-\kappa)b\right)dxdt\\ &=\int_{0}^{\lambda}\int\left[u\cdot\nabla b\cdot(\phi(x-\kappa)b)-v\otimes a:\nabla(\phi(x-\kappa)b)\right.\\ &\qquad\qquad\ \left.-a\cdot\nabla v\cdot(\phi(x-\kappa)b)+b\otimes u:\nabla(\phi(x-\kappa)b)\right]dxdt\\ &\leq C\int_{0}^{\lambda}\int_{B_{2}(\kappa)}\left[|u|(|b|^{2}+|b||\nabla b|)+|a|(|v||b|+|v||\nabla b|+|b||\nabla v|)\right]dxdt\\ &\leq C\left\|u\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|b\right\|_{L^{6}(B_{2}(\kappa))}\left(\left\|b\right\|_{L^{2}(B_{2}(\kappa))}+\left\|\nabla b\right\|_{L^{2}(B_{2}(\kappa))}\right)dt\\ &\quad+C\left\|a\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|v\right\|_{L^{6}(B_{2}(\kappa))}\left(\left\|b\right\|_{L^{2}(B_{2}(\kappa))}+\left\|\nabla b\right\|_{L^{2}(B_{2}(\kappa))}\right)dt\\ &\quad+C\left\|a\right\|_{L^{\infty}L^{3}_{\mathrm{uloc}}}\int_{0}^{\lambda}\left\|b\right\|_{L^{6}(B_{2}(\kappa))}\left\|\nabla v\right\|_{L^{2}(B_{2}(\kappa))}dt\\ &\leq C\lambda\delta\mathop{\rm ess\,sup}_{0<t<\lambda}\sum_{\kappa^{\prime}\sim\kappa}\int_{B_{1}(\kappa^{\prime})}\left(|v(x,t)|^{2}+|b(x,t)|^{2}\right)dx\\ &\quad+C\delta\mathop{\rm ess\,sup}_{0<t<\lambda}\sum_{\kappa^{\prime}\sim\kappa}\int_{0}^{\lambda}\int_{B_{1}(\kappa^{\prime})}\left(|\nabla v(x,t)|^{2}+|\nabla b(x,t)|^{2}\right)dxdt.\end{split}

The pressure can be decomposed into local and far-field contributions. The local part contains new terms that are handled exactly as in the previous estimates, once the Calderon–Zygmund inequality is applied. The far-field pressure splits as πfar=πfar,(v,b)+πfar,(u,a)\pi_{\rm far}=\pi_{{\rm far},(v,b)}+\pi_{{\rm far},(u,a)}, where πfar,(v,b)\pi_{{\rm far},(v,b)} matches the far-field term treated in the proof of Lemma 3.7, and πfar,(u,a)\pi_{{\rm far},(u,a)} is the remaining contribution. Since the estimate for πfar,(v,b)\pi_{{\rm far},(v,b)} is already established in the proof of Lemma 3.7, we focus on bounding πfar,(u,a)\pi_{{\rm far},(u,a)} in B2(κ)×(0,T)B_{2}(\kappa)\times(0,T). Specifically, we have

|πfar,(u,a)(x,t)|C1|κy|4(|v(y,t)||u(y,t)|+|b(y,t)||a(y,t)|)(1χ4(yκ))dyCδK~eλ1/2(κ),\begin{split}\left|\pi_{{\rm far},(u,a)}(x,t)\right|&\leq C\int\frac{1}{|\kappa-y|^{4}}\left(|v(y,t)||u(y,t)|+|b(y,t)||a(y,t)|\right)\left(1-\chi_{4}(y-\kappa)\right)dy\\ &\leq C\delta\widetilde{K}*e_{\lambda}^{1/2}(\kappa),\end{split}

where K~\widetilde{K} is defined in (3.38), and

eλ(κ)=esssup0tλB1(κ)(|v(x,t)|2+|b(x,t)|2)dx+0λB1(κ)(|v(x,t)|2+|b(x,t)|2)dxdt.e_{\lambda}(\kappa)=\mathop{\rm ess\,sup}_{0\leq t\leq\lambda}\int_{B_{1}(\kappa)}\left(|v(x,t)|^{2}+|b(x,t)|^{2}\right)dx+\int_{0}^{\lambda}\int_{B_{1}(\kappa)}\left(|\nabla v(x,t)|^{2}+|\nabla b(x,t)|^{2}\right)dxdt.

This yields the estimate:

0λπfar,(u,a)(x,s)v(x,s)ϕ(xκ)dxdsC0λB2(κ)δ1/2(K~eλ1/2)δ1/2|v|dxdsCδ0λB2(κ)(K~eλ1/2)2dxds+δ0λB2(κ)|v|2dxdsCδλ((K~eλ1/2)(κ))2+δλesssup0<t<λ|κκ|<4B1(κ)|v|2dx.\begin{split}\int_{0}^{\lambda}\int&\pi_{{\rm far},(u,a)}(x,s)v(x,s)\cdot\nabla\phi(x-\kappa)\,dxds\\ &\leq C\int_{0}^{\lambda}\int_{B_{2}(\kappa)}\delta^{1/2}(\widetilde{K}*e_{\lambda}^{1/2})\delta^{1/2}|v|\,dxds\\ &\leq C\delta\int_{0}^{\lambda}\int_{B_{2}(\kappa)}(\widetilde{K}*e_{\lambda}^{1/2})^{2}\,dxds+\delta\int_{0}^{\lambda}\int_{B_{2}(\kappa)}|v|^{2}\,dxds\\ &\leq C\delta\lambda\left((\widetilde{K}*e_{\lambda}^{1/2})(\kappa)\right)^{2}+\delta\lambda\mathop{\rm ess\,sup}_{0<t<\lambda}\sum_{|\kappa-\kappa^{\prime}|<4}\int_{B_{1}(\kappa^{\prime})}|v|^{2}\,dx.\end{split}

Combining the above estimates with the argument in the proof of Lemma 3.7 (see (3.24)), we obtain

eλ(κ)(|v0|2+|b0|2)ϕ(xκ)dx+Cλκ3;|κκ|2eλ(κ)+Cλ1/4κ3;|κκ|10(eλ(κ))3/2+Cλ1/4((K~eλ)(κ))3/2+Cδ|κκ|<10eλ(κ)+Cδλ1/4((K~eλ1/2)(κ))2,\begin{split}e_{\lambda}(\kappa)&\leq\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi(x-\kappa)\,dx+C\lambda\sum_{\kappa^{\prime}\in\mathbb{Z}^{3};|\kappa^{\prime}-\kappa|\leq 2}e_{\lambda}(\kappa^{\prime})\\ &\quad+C\lambda^{1/4}\sum_{\kappa^{\prime}\in\mathbb{Z}^{3};|\kappa^{\prime}-\kappa|\leq 10}\left(e_{\lambda}(\kappa^{\prime})\right)^{3/2}+C\lambda^{1/4}\left((\widetilde{K}*e_{\lambda})(\kappa)\right)^{3/2}\\ &\quad+C\delta\sum_{|\kappa-\kappa^{\prime}|<10}e_{\lambda}(\kappa^{\prime})+C\delta\lambda^{1/4}\left((\widetilde{K}*e_{\lambda}^{1/2})(\kappa)\right)^{2},\end{split} (3.53)

where we are using λλ01\lambda\leq\lambda_{0}\leq 1. The first two lines above coincide exactly with the estimates in the proof of Lemma 3.7, so we focus on the two additional terms in the last line. To control the final term, we apply the q/2\ell^{q/2} norm:

((K~eλ1/2)(κ))2q/q=(K~eλ1/2)(κ)q/q2CK~1eλq/2.\left\|\left((\widetilde{K}*e_{\lambda}^{1/2})(\kappa)\right)^{2}\right\|_{\ell^{q/q}}=\left\|(\widetilde{K}*e_{\lambda}^{1/2})(\kappa)\right\|_{\ell^{q/q}}^{2}\leq C\left\|\widetilde{K}\right\|_{\ell^{1}}\left\|e_{\lambda}\right\|_{\ell^{q/2}}.

We now choose c0c_{0} (which bounds δ\delta) sufficiently small so that, after taking the q/2\ell^{q/2} norm of both sides of the inequality, the δ\delta-weighted terms on the right can be absorbed into the left-hand side. With this absorption, the remaining terms are exactly as in the proof of Lemma 3.7, and the conclusion follows by the same argument. ∎

Lemma 3.11.

Let ϵ,δ>0\epsilon,\delta>0 be given and assume δc0\delta\leq c_{0}, where c0c_{0} is given in Lemma 3.11. Assume that v0,b0L2v_{0},b_{0}\in L^{2} are divergence free and that u,a:3×[0,T0]3u,a:\mathbb{R}^{3}\times[0,T_{0}]\to\mathbb{R}^{3} are divergence free that satisfy

esssup0<tT0(u(t)L3uloc+b(t)L3uloc)<δ.\mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left(\left\|u(t)\right\|_{L^{3}_{\mathrm{uloc}}}+\left\|b(t)\right\|_{L^{3}_{\mathrm{uloc}}}\right)<\delta.

Then there exist T(0,T0]T\in(0,T_{0}] and a weak solution (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) and pressure πϵ\pi_{\epsilon} to (3.34) on 3×[0,T]\mathbb{R}^{3}\times[0,T]. Furthermore, we have that (vϵ,bϵ)L(0,T;E2q×E2q)(v_{\epsilon},b_{\epsilon})\in L^{\infty}(0,T;E^{2}_{q}\times E^{2}_{q}) and satisfies

esssup0tT0B1(x0)|vϵ|2+|bϵ|22dx+0T0B1(x0)(|vϵ|2+|bϵ|2)dxdtq2(x03)22C(v0,b0)E2q×E2q\begin{split}\left\|\mathop{\rm ess\,sup}_{0\leq t\leq T_{0}}\int_{B_{1}(x_{0})}\frac{|v_{\epsilon}|^{2}+|b_{\epsilon}|^{2}}{2}\,dx+\int_{0}^{T_{0}}\int_{B_{1}(x_{0})}\left(|\nabla v_{\epsilon}|^{2}+|\nabla b_{\epsilon}|^{2}\right)dxdt\right\|_{\ell^{\frac{q}{2}}(x_{0}\in\mathbb{Z}^{3})}^{2}\leq 2C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}\end{split}

for some positive constant CC independent of ϵ,δ,(u,a)\epsilon,\delta,(u,a) and (v0,b0)(v_{0},b_{0}). Here, T=min(T0,λ0,λ0A0,q2)T=\min\left(T_{0},\lambda_{0},\lambda_{0}A_{0,q}^{-2}\right) depends on (v0,b0)E2q×E2q\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}} but not on (v0,b0)L2×L2\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}, (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}), ϵ,δ\epsilon,\delta, or (u,a)(u,a).

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 4.3] for the Navier–Stokes equations to the MHD equations.

Let (vϵ,bϵ,πϵ)(v_{\epsilon},b_{\epsilon},\pi_{\epsilon}) be a smooth solution of (3.34) on 3×[0,T0]\mathbb{R}^{3}\times[0,T_{0}] with initial data(v0,b0)L2(v_{0},b_{0})\in L^{2}. The energy equality for the regularized perturbed problem reads:

vϵ(t)L22+bϵ(t)L22+20t0(|vϵ|2+|bϵ|2)dxds=v0L22+b0L22+20t[(vϵvϵbϵbϵ)(ηϵu)+(vϵbϵbϵvϵ)(ηϵa)]dxds\begin{split}&\left\|v_{\epsilon}(t)\right\|_{L^{2}}^{2}+\left\|b_{\epsilon}(t)\right\|_{L^{2}}^{2}+2\int_{0}^{t}\int_{0}(|\nabla v_{\epsilon}|^{2}+|\nabla b_{\epsilon}|^{2})\,dxds\\ &=\left\|v_{0}\right\|_{L^{2}}^{2}+\left\|b_{0}\right\|_{L^{2}}^{2}+2\int_{0}^{t}\int\left[(v_{\epsilon}\cdot\nabla v_{\epsilon}-b_{\epsilon}\cdot\nabla b_{\epsilon})\cdot(\eta_{\epsilon}*u)+(v_{\epsilon}\cdot\nabla b_{\epsilon}-b_{\epsilon}\cdot\nabla v_{\epsilon})\cdot(\eta_{\epsilon}*a)\right]dxds\end{split}

By the estimate from [27, p. 217], the right-hand side is uniformly bounded in ϵ\epsilon, implying that (vϵ,bϵ)(t)L2×L2M1\left\|(v_{\epsilon},b_{\epsilon})(t)\right\|_{L^{2}\times L^{2}}\leq M_{1} for all t[0,T0]t\in[0,T_{0}] for some constant M1M_{1} independent of ϵ\epsilon. Furthermore, if (vϵ,bϵ)𝐋𝐄q(0,T0)(v_{\epsilon},b_{\epsilon})\in{\bf LE}_{q}(0,T_{0}), then by Lemma 3.10, the local energy estimates extend up to time T=min(T0,λ0,λ0A0,q2)T=\min\left(T_{0},\lambda_{0},\lambda_{0}A_{0,q}^{-2}\right), yielding (vϵ,bϵ)𝐋𝐄q(0,T0)<M2\left\|(v_{\epsilon},b_{\epsilon})\right\|_{{\bf LE}_{q}(0,T_{0})}<M_{2} for some constant M2M_{2} independent of ϵ\epsilon.

Now, let Tϵ,δT_{\epsilon,\delta} be the time-scale provided in Lemma 3.9 corresponding to initial data of size M1M_{1} in L2L^{2}. Then, Lemma 3.9 ensures that the solution (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) exists on 3×[0,Tϵ,δ]\mathbb{R}^{3}\times[0,T_{\epsilon,\delta}] and belongs to 𝐋𝐄q(0,Tϵ,δ){\bf LE}_{q}(0,T_{\epsilon,\delta}). Since Lemma 3.10 also applies to the regularized system, we further conclude that (vϵ,bϵ)𝐋𝐄q(0,Tϵ,δ)<M2\left\|(v_{\epsilon},b_{\epsilon})\right\|_{{\bf LE}_{q}(0,T_{\epsilon,\delta})}<M_{2} and hence esssup0<tTϵ,δ(v,b)(t)E2q×E2qM2\mathop{\rm ess\,sup}_{0<t\leq T_{\epsilon,\delta}}\left\|(v,b)(t)\right\|_{E^{2}_{q}\times E^{2}_{q}}\leq M_{2}. In addition, the energy estimate gives esssup0<tTϵ,δ(v,b)(t)L2×L2M1\mathop{\rm ess\,sup}_{0<t\leq T_{\epsilon,\delta}}\left\|(v,b)(t)\right\|_{L^{2}\times L^{2}}\leq M_{1}. This allows us to restart the solution at any time t[Tϵ,δ/2,3Tϵ,δ/4]t_{*}\in[T_{\epsilon,\delta}/2,3T_{\epsilon,\delta}/4], and apply Lemma 3.10 again with the same bounds. By uniqueness, the extended solution coincides with the original one, and hence we obtain a solution on [0,3Tϵ,δ/2][0,3T_{\epsilon,\delta}/2] that remains in 𝐋𝐄q(0,3Tϵ,δ/2){\bf LE}_{q}(0,3T_{\epsilon,\delta}/2). Repeating this argument and iterating the solution step-by-step, we reach the full time interval [0,T][0,T]. Throughout the iteration, the 𝐋𝐄q{\bf LE}_{q} and L2L^{2} norms remain bounded uniformly by M2M_{2} and M1M_{1}, respectively. Therefore, for each ϵ>0\epsilon>0, the solution (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) to the regularized system exists on [0,T][0,T] and satisfies (vϵ,bϵ)𝐋𝐄q(0,T)(v_{\epsilon},b_{\epsilon})\in{\bf LE}_{q}(0,T) with bounds independent of ϵ\epsilon. ∎

Lemma 3.12.

Let c0c_{0} and λ0\lambda_{0} be the constants in Lemma 3.10. Assume that v0,b0E2qv_{0},b_{0}\in E^{2}_{q} are divergence free and that u,a:3×[0,T0]3u,a:\mathbb{R}^{3}\times[0,T_{0}]\to\mathbb{R}^{3} are divergence free and satisfy

esssup0<tT0(u(t)L3uloc+a(t)L3uloc)<δc0 and esssup0<tT0(u(t)L4uloc+a(t)L4uloc)<.\mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left(\left\|u(t)\right\|_{L^{3}_{\mathrm{uloc}}}+\left\|a(t)\right\|_{L^{3}_{\mathrm{uloc}}}\right)<\delta\leq c_{0}\ \text{ and }\ \mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left(\left\|u(t)\right\|_{L^{4}_{\mathrm{uloc}}}+\left\|a(t)\right\|_{L^{4}_{\mathrm{uloc}}}\right)<\infty.

Let T=min(T0,λ0,λ0A0,q2)T=\min\left(T_{0},\,\lambda_{0},\,\lambda_{0}A_{0,q}^{-2}\right). Then there exist a local energy solution (v,b)(v,b) and π\pi to the perturbed MHD equations, (3.30), satisfying

(v,b)𝐋𝐄q(0,T)C(v0,b0)E2q×E2q\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}\leq C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}

for some constant C>0C>0.

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 4.4] for the Navier–Stokes equations to the MHD equations.

Fix (v0,b0)E2q×E2q(v_{0},b_{0})\in E^{2}_{q}\times E^{2}_{q}. For each ϵ>0\epsilon>0, approximate the data by divergence-free vector fields (v0(ϵ),b0(ϵ))L2×L2(v_{0}^{(\epsilon)},b_{0}^{(\epsilon)})\in L^{2}\times L^{2} satisfying v0v0(ϵ)E2q+b0b0(ϵ)E2q<ϵ\left\|v_{0}-v_{0}^{(\epsilon)}\right\|_{E^{2}_{q}}+\left\|b_{0}-b_{0}^{(\epsilon)}\right\|_{E^{2}_{q}}<\epsilon. Such approximations can be constructed using the Bogovskii map (see [40]). Let (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) denote the solutions constructed in Lemma 3.12 corresponding to the initial data (v0(ϵ),b0(ϵ))(v_{0}^{(\epsilon)},b_{0}^{(\epsilon)}). By the uniform estimates from Lemma 3.12, we obtain bounds on tvϵ\partial_{t}v_{\epsilon} and tbϵ\partial_{t}b_{\epsilon} in the dual of L3(0,T;W01,3(BM(0)))L^{3}(0,T;W_{0}^{1,3}(B_{M}(0))), which allow us to extract a subsequence (vn,bn):=(vϵn,bϵn)(v_{n},b_{n}):=(v_{\epsilon_{n}},b_{\epsilon_{n}}) of (vϵ,bϵ)(v_{\epsilon},b_{\epsilon}) and πn:=πϵn\pi_{n}:=\pi_{\epsilon_{n}} of πϵ\pi_{\epsilon}, such that, as nn\to\infty,

(vn,bn)(v,b) in L(0,T;L2loc×L2loc),(vn,bn)(v,b) in L2(0,T;H1loc×H1loc),(vn,bn),(ηϵnvn,ηϵnbn)(v,b) in L3(0,T;L3loc×L3loc),(ηϵnu,ηϵna)(u,a) in L3(0,T;L3loc×L3loc),πn(k)π(k) in L3/2(0,T;L3/2(Bk(0))),\begin{split}(v_{n},b_{n})\overset{*}{\rightharpoonup}(v,b)&\ \text{ in }L^{\infty}(0,T;L^{2}_{\mathrm{loc}}\times L^{2}_{\mathrm{loc}}),\\ (v_{n},b_{n})\rightharpoonup(v,b)&\ \text{ in }L^{2}(0,T;H^{1}_{\mathrm{loc}}\times H^{1}_{\mathrm{loc}}),\\ (v_{n},b_{n}),(\eta_{\epsilon_{n}}*v_{n},\eta_{\epsilon_{n}}*b_{n})\rightarrow(v,b)&\ \text{ in }L^{3}(0,T;L^{3}_{\mathrm{loc}}\times L^{3}_{\mathrm{loc}}),\\ (\eta_{\epsilon_{n}}*u,\eta_{\epsilon_{n}}*a)\rightarrow(u,a)&\ \text{ in }L^{3}(0,T;L^{3}_{\mathrm{loc}}\times L^{3}_{\mathrm{loc}}),\\ \pi_{n}^{(k)}\rightharpoonup\pi^{(k)}&\ \text{ in }L^{3/2}(0,T;L^{3/2}(B_{k}(0))),\end{split}

where π(k)(x,t)=π(x,t)ck(t)\pi^{(k)}(x,t)=\pi(x,t)-c_{k}(t) for xBk(0)x\in B_{k}(0) and t(0,T0]t\in(0,T_{0}] for some ckL3/2(0,T0)c_{k}\in L^{3/2}(0,T_{0}), and πn(k)\pi_{n}^{(k)} is the local pressure expansion for πn\pi_{n} on ball Bk(0)B_{k}(0). The limit (v,b,π)(v,b,\pi) is a local energy solution to the perturbed MHD equations with initial data (v0,b0)(v_{0},b_{0}). We claim that (v,b,π)(v,b,\pi) satisfies the perturbed local energy inequality: for all nonnegative ϕCc(3×[0,T))\phi\in C^{\infty}_{c}(\mathbb{R}^{3}\times[0,T)),

2(|v|2+|b|2)ϕdxdt(|v0|2+|b0|2)ϕdx+(|v|2+|b|2)(tϕ+Δϕ)dxdt+(|v|2+|b|2+2π)(vϕ)dxdt+(|v|2+|b|2)(uϕ)dxdt+2(vvbb)(uϕ)dxdt+2(vu+ba)(vϕ)dxdt+2(vbbv)(aϕ)dxdt2(va+bu)(bϕ)dxdt2(vb)((b+a)ϕ)dxdt.\begin{split}2&\iint\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\phi\,dxdt\\ &\leq\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\phi\,dx+\iint\left(|v|^{2}+|b|^{2}\right)(\partial_{t}\phi+\Delta\phi)\,dxdt+\iint\left(|v|^{2}+|b|^{2}+2\pi\right)(v\cdot\nabla\phi)\,dxdt\\ &\quad+\iint\left(|v|^{2}+|b|^{2}\right)(u\cdot\nabla\phi)\,dxdt+2\iint(v\cdot\nabla v-b\cdot\nabla b)\cdot(u\phi)\,dxdt\\ &\quad+2\iint(v\cdot u+b\cdot a)(v\cdot\nabla\phi)\,dxdt+2\iint(v\cdot\nabla b-b\cdot\nabla v)\cdot(a\phi)\,dxdt\\ &\quad-2\iint(v\cdot a+b\cdot u)(b\cdot\nabla\phi)\,dxdt-2\iint(v\cdot b)((b+a)\cdot\nabla\phi)\,dxdt.\end{split} (3.54)

The first two lines are inherited via standard compactness arguments. We now focus on the convergence of the remaining terms, especially those not involving ϕ\nabla\phi, which are of higher order. We have

|(vnvn)(ηϵnu)ϕ(vv)uϕdxdt||((vnv)vn)(uϕ)dxdt|+|(v(vnv))(uϕ)dxdt|+|(vnvn)(ηϵnuu)ϕdxdt|=:I1,n+I2,n+I3,n,\begin{split}&\left|\iint(v_{n}\cdot\nabla v_{n})\cdot(\eta_{\epsilon_{n}}*u)\phi-(v\cdot\nabla v)\cdot u\phi\,dxdt\right|\\ &\qquad\leq\left|\iint((v_{n}-v)\cdot\nabla v_{n})\cdot(u\phi)\,dxdt\right|+\left|\iint(v\cdot\nabla(v_{n}-v))\cdot(u\phi)\,dxdt\right|\\ &\qquad\quad+\left|\iint(v_{n}\cdot\nabla v_{n})\cdot(\eta_{\epsilon_{n}}*u-u)\phi\,dxdt\right|\\ &\qquad=:I_{1,n}+I_{2,n}+I_{3,n},\end{split} (3.55)
|(bnbn)(ηϵnu)ϕ(bb)uϕdxdt||((bnb)bn)(uϕ)dxdt|+|(b(bnb))(uϕ)dxdt|+|(bnbn)(ηϵnuu)ϕdxdt|=:I4,n+I5,n+I6,n,\begin{split}&\left|\iint(b_{n}\cdot\nabla b_{n})\cdot(\eta_{\epsilon_{n}}*u)\phi-(b\cdot\nabla b)\cdot u\phi\,dxdt\right|\\ &\qquad\leq\left|\iint((b_{n}-b)\cdot\nabla b_{n})\cdot(u\phi)\,dxdt\right|+\left|\iint(b\cdot\nabla(b_{n}-b))\cdot(u\phi)\,dxdt\right|\\ &\qquad\quad+\left|\iint(b_{n}\cdot\nabla b_{n})\cdot(\eta_{\epsilon_{n}}*u-u)\phi\,dxdt\right|\\ &\qquad=:I_{4,n}+I_{5,n}+I_{6,n},\end{split} (3.56)
|(vnbn)(ηϵna)ϕ(vb)aϕdxdt||((vnv)bn)(aϕ)dxdt|+|(v(bnb))(aϕ)dxdt|+|(vnbn)(ηϵnaa)ϕdxdt|=:I7,n+I8,n+I9,n,\begin{split}&\left|\iint(v_{n}\cdot\nabla b_{n})\cdot(\eta_{\epsilon_{n}}*a)\phi-(v\cdot\nabla b)\cdot a\phi\,dxdt\right|\\ &\qquad\leq\left|\iint((v_{n}-v)\cdot\nabla b_{n})\cdot(a\phi)\,dxdt\right|+\left|\iint(v\cdot\nabla(b_{n}-b))\cdot(a\phi)\,dxdt\right|\\ &\qquad\quad+\left|\iint(v_{n}\cdot\nabla b_{n})\cdot(\eta_{\epsilon_{n}}*a-a)\phi\,dxdt\right|\\ &\qquad=:I_{7,n}+I_{8,n}+I_{9,n},\end{split} (3.57)

and

|(bnvn)(ηϵna)ϕ(bv)aϕdxdt||((bnb)vn)(aϕ)dxdt|+|(b(vnv))(aϕ)dxdt|+|(bnvn)(ηϵnaa)ϕdxdt|=:I10,n+I11,n+I12,n.\begin{split}&\left|\iint(b_{n}\cdot\nabla v_{n})\cdot(\eta_{\epsilon_{n}}*a)\phi-(b\cdot\nabla v)\cdot a\phi\,dxdt\right|\\ &\qquad\leq\left|\iint((b_{n}-b)\cdot\nabla v_{n})\cdot(a\phi)\,dxdt\right|+\left|\iint(b\cdot\nabla(v_{n}-v))\cdot(a\phi)\,dxdt\right|\\ &\qquad\quad+\left|\iint(b_{n}\cdot\nabla v_{n})\cdot(\eta_{\epsilon_{n}}*a-a)\phi\,dxdt\right|\\ &\qquad=:I_{10,n}+I_{11,n}+I_{12,n}.\end{split} (3.58)

Our goals is to show that the above twelve quantities vanishes as nn\to\infty. Let BB be a ball containing suppϕ\mathop{\mathrm{supp}}\nolimits\,\phi. Then, using Hölder’s inequality and log-convexity of LpL^{p} norms, we have

I1,nuLL4ulocvnvL2(0,T;L2(B))1/4vnL2(0,T;H1(B))7/40,I_{1,n}{\ \lesssim\ }\left\|u\right\|_{L^{\infty}L^{4}_{\mathrm{uloc}}}\left\|v_{n}-v\right\|_{L^{2}(0,T;L^{2}(B))}^{1/4}\left\|v_{n}\right\|_{L^{2}(0,T;H^{1}(B))}^{7/4}\to 0,
I4,nuLL4ulocbnbL2(0,T;L2(B))1/4bnL2(0,T;H1(B))7/40,I_{4,n}{\ \lesssim\ }\left\|u\right\|_{L^{\infty}L^{4}_{\mathrm{uloc}}}\left\|b_{n}-b\right\|_{L^{2}(0,T;L^{2}(B))}^{1/4}\left\|b_{n}\right\|_{L^{2}(0,T;H^{1}(B))}^{7/4}\to 0,
I7,naLL4ulocvnvL2(0,T;L2(B))1/4bnL2(0,T;H1(B))7/40,I_{7,n}{\ \lesssim\ }\left\|a\right\|_{L^{\infty}L^{4}_{\mathrm{uloc}}}\left\|v_{n}-v\right\|_{L^{2}(0,T;L^{2}(B))}^{1/4}\left\|b_{n}\right\|_{L^{2}(0,T;H^{1}(B))}^{7/4}\to 0,

and

I10,naLL4ulocbnbL2(0,T;L2(B))1/4vnL2(0,T;H1(B))7/40,I_{10,n}{\ \lesssim\ }\left\|a\right\|_{L^{\infty}L^{4}_{\mathrm{uloc}}}\left\|b_{n}-b\right\|_{L^{2}(0,T;L^{2}(B))}^{1/4}\left\|v_{n}\right\|_{L^{2}(0,T;H^{1}(B))}^{7/4}\to 0,

as nn\to\infty by strong convergence of (vn,bn)(v_{n},b_{n}) to (v,b)(v,b) in L2(0,T;L2(B)×L2(B))L^{2}(0,T;L^{2}(B)\times L^{2}(B)). Next, weak convergence of (vn,bn)(v_{n},b_{n}) to (v,b)(v,b) in L2(0,T;H1(B)×H1(B))L^{2}(0,T;H^{1}(B)\times H^{1}(B)) ensures that I2,nI_{2,n}, I5,nI_{5,n}, I8,nI_{8,n}, I11,n0I_{11,n}\to 0 as nn\to\infty, since the products viujϕ,biujϕ,viajϕv_{i}u_{j}\phi,b_{i}u_{j}\phi,v_{i}a_{j}\phi, and biajϕb_{i}a_{j}\phi all belong to L2(B×(0,T))L^{2}(B\times(0,T)). Finally, for the mollifier terms, we use strong convergence of the mollified quantities in L(0,T;L3(B))L^{\infty}(0,T;L^{3}(B)) and uniform bounds on vnv_{n}, bnb_{n} in L2(0,T;H1(B))L^{2}(0,T;H^{1}(B)), to deduce I3,nI_{3,n}, I6,nI_{6,n}, I9,nI_{9,n}, and I12,n0I_{12,n}\to 0 as nn\to\infty. Hence, all error terms vanish in the limit, and (3.54) holds. Moreover, following the argument in [23, (3.28)-(3.29)], we derive the time-slice version of the perturbed local energy inequality: for any nonnegative ψCc(3)\psi\in C^{\infty}_{c}(\mathbb{R}^{3}) and any t(0,T)t\in(0,T),

(|v(t)|2+|b(t)|2)ψdx+20t(|v|2+|b|2)ψdxdt(|v0|2+|b0|2)ψdx+0t(|v|2+|b|2)Δψdxdt+(|v|2+|b|2+2π)(vψ)dxdt+0t(|v|2+|b|2)(uψ)dxdt+20t(vvbb)(uψ)dxdt+2(vu+ba)(vψ)dxdt+20t(vbbv)(aψ)dxdt20t(va+bu)(bψ)dxdt20t(vb)((b+a)ψ)dxdt.\begin{split}\int&\left(|v(t)|^{2}+|b(t)|^{2}\right)\psi\,dx+2\int_{0}^{t}\int\left(|\nabla v|^{2}+|\nabla b|^{2}\right)\psi\,dxdt\\ &\leq\int\left(|v_{0}|^{2}+|b_{0}|^{2}\right)\psi\,dx+\int_{0}^{t}\int\left(|v|^{2}+|b|^{2}\right)\Delta\psi\,dxdt+\iint\left(|v|^{2}+|b|^{2}+2\pi\right)(v\cdot\nabla\psi)\,dxdt\\ &\quad+\int_{0}^{t}\int\left(|v|^{2}+|b|^{2}\right)(u\cdot\nabla\psi)\,dxdt+2\int_{0}^{t}\int(v\cdot\nabla v-b\cdot\nabla b)\cdot(u\psi)\,dxdt\\ &\quad+2\iint(v\cdot u+b\cdot a)(v\cdot\nabla\psi)\,dxdt+2\int_{0}^{t}\int(v\cdot\nabla b-b\cdot\nabla v)\cdot(a\psi)\,dxdt\\ &\quad-2\int_{0}^{t}\int(v\cdot a+b\cdot u)(b\cdot\nabla\psi)\,dxdt-2\int_{0}^{t}\int(v\cdot b)((b+a)\cdot\nabla\psi)\,dxdt.\end{split} (3.59)

We now establish the bound for (v,b)𝐋𝐄q(0,T)\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}. Let ϕ\phi the cutoff function used in the proof of Lemma 3.7, and define ψ(x):=ϕ(xk)\psi(x):=\phi(x-k) for each k3k\in\mathbb{Z}^{3}. Fix a large integer K>0K>0, and restrict to |k|K|k|\leq K. Applying (3.59) with this choice of ψ\psi, the right-hand side can be approximated by the corresponding terms for the sequence {(vn,bn)}\{(v_{n},b_{n})\}, since all such quantities converge in the limit. In particular, for all |k|K|k|\leq K, we can ensure that the difference between the terms involving (v,b)(v,b) and those for (vn,bn)(v_{n},b_{n}) is less than 2KK32^{-K}K^{-3} uniformly, provided nNKn\geq N_{K} for some sufficiently large NKN_{K}. Taking these approximate terms, applying standard estimates (which can be derived similar to the proof of (3.19))

0λRR2BR(x0R)|v|3+|b|3+|πcRx0,R(t)|3/2dxdtq3(x03)CλR110R12(A0,q(R))32,R>0.\begin{split}\left\|\int_{0}^{\lambda_{R}R^{2}}\int_{B_{R}(x_{0}R)}|v|^{3}+|b|^{3}+\left|\pi-c_{Rx_{0},R}(t)\right|^{3/2}dxdt\right\|_{\ell^{\frac{q}{3}}(x_{0}\in\mathbb{Z}^{3})}\leq C\lambda_{R}^{\frac{1}{10}}R^{\frac{1}{2}}\left(A_{0,q}(R)\right)^{\frac{3}{2}},\quad R>0.\end{split} (3.60)

Taking the essential supremum in time followed by the q/2\ell^{q/2}-sum over |k|K|k|\leq K, we obtain a uniform bound

[|k|K(esssup0<t<TB1(k)(|v(x,t)|2+|b(x,t)|2)dx+0TB1(k)(|v|2+|b|2)dxdt)q/2]1/qC((v0,b0)E2q×E2q+C2K).\begin{split}&\left[\sum_{|k|\leq K}\left(\mathop{\rm ess\,sup}_{0<t<T}\int_{B_{1}(k)}\left(|v(x,t)|^{2}+|b(x,t)|^{2}\right)dx+\int_{0}^{T}\int_{B_{1}(k)}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dxdt\right)^{q/2}\right]^{1/q}\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\leq C\left(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+\frac{C}{2^{K}}\right).\end{split}

Since this bound holds for all KK\in\mathbb{N}, it follows that (v,b)𝐋𝐄q(0,T)(v,b)\in{\bf LE}_{q}(0,T), with the norm estimate (v,b)𝐋𝐄q(0,T)C(v0,b0)E2q×E2q\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}\leq C\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}. ∎

Lemma 3.13.

Let 2q<2\leq q<\infty. Assume v0,b0L2v_{0},b_{0}\in L^{2} are divergence free, and assume u,a:3×[0,T0]3u,a:\mathbb{R}^{3}\times[0,T_{0}]\to\mathbb{R}^{3} are divergence free and satisfy

esssup0<tT0(u(t)L3uloc+a(t)L3uloc)<δc0 and esssup0<tT0(u(t)L4uloc+a(t)L4uloc)<.\mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left(\left\|u(t)\right\|_{L^{3}_{\mathrm{uloc}}}+\left\|a(t)\right\|_{L^{3}_{\mathrm{uloc}}}\right)<\delta\leq c_{0}\ \text{ and }\ \mathop{\rm ess\,sup}_{0<t\leq T_{0}}\left(\left\|u(t)\right\|_{L^{4}_{\mathrm{uloc}}}+\left\|a(t)\right\|_{L^{4}_{\mathrm{uloc}}}\right)<\infty.

where c0c_{0} is from Lemma 3.10. For any T(0,T0]T\in(0,T_{0}], if δδ0(T)c0\delta\leq\delta_{0}(T)\leq c_{0} is sufficiently small, then there exists a local energy solution (v,b)(v,b) to the perturbed MHD equations, (3.30), so that (v,b)𝐋𝐄q(0,T)(v,b)\in{\bf LE}_{q}(0,T). In particular, this is true when (u,a)(0,0)(u,a)\equiv(0,0).

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma 5.1] for the Navier–Stokes equations to the MHD equations.

We begin with the special case (u,a)=(0,0)(u,a)=(0,0) to highlight the key ideas. Suppose the initial data (v0,b0)L2×L2(v_{0},b_{0})\in L^{2}\times L^{2}, and define ak=B1(k)(|v0|2+|b0|2)dxa_{k}=\int_{B_{1}(k)}\left(|v_{0}|^{2}+|b_{0}|^{2}\right)dx for k3k\in\mathbb{Z}^{3}. Then,

kakq/2(maxkak)q/21kak(kak)q/2,\sum_{k}a_{k}^{q/2}\leq(\max_{k}a_{k})^{q/2-1}\sum_{k}a_{k}\leq\left(\sum_{k}a_{k}\right)^{q/2},

which shows that (v0,b0)E2q×E2q(v_{0},b_{0})\in E^{2}_{q}\times E^{2}_{q} with (v0,b0)E2q×E2qM2:=C(v0,b0)L2×L2\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}\leq M_{2}:=C\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}. Let (v,b)(v,b) be the solution of the perturbed MHD equations with (u,a)=(0,0)(u,a)=(0,0) (so that it is a solution of (MHD)) constructed via Lemma 3.12 with initial data (v0,b0)(v_{0},b_{0}). Then (v,b)𝐋𝐄q(0,T0)(v,b)\in{\bf LE}_{q}(0,T_{0}), where T0=T0(M2)T_{0}=T_{0}(M_{2}) is the existence time depending on the size of the initial data in E2q×E2qE^{2}_{q}\times E^{2}_{q}. For almost every t(0,T)t\in(0,T), we have that (v,b)(t)E3×E3(v,b)(t)\in E^{3}\times E^{3} and that (v,b)(t)L2×L2(v0,b0)L2×L2\left\|(v,b)(t)\right\|_{L^{2}\times L^{2}}\leq\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}. The inclusion (v,b)(t)E3(v,b)(t)\in E^{3} follows from Lemma 3.8 and the embedding E4qE3E^{4}_{q}\subset E^{3}. Hence, (v,b)(t)E2q×E2qM2\left\|(v,b)(t)\right\|_{E^{2}_{q}\times E^{2}_{q}}\leq M_{2} for almost every t(0,T0)t\in(0,T_{0}). In particular, these bounds hold at some time t0(T0/2,T0)t_{0}\in(T_{0}/2,T_{0}). We now restart the MHD equations at time t0t_{0}, treating (v,b)(t0)(v,b)(t_{0}) as new initial data in E3×E3E^{3}\times E^{3}. By Lemma 3.12, there exists a local energy solution (v1,b1)(v_{1},b_{1}) in 𝐋𝐄q(t0,t0+T0){\bf LE}_{q}(t_{0},t_{0}+T_{0}). By uniqueness of local energy solution with E3×E3E^{3}\times E^{3} (Corollary 3.5), we have (v1,b1)=(v,b)(v_{1},b_{1})=(v,b) on some short interval [t0,t0+Δt1][t_{0},t_{0}+\Delta t_{1}]. This allows us to glue (v1,b1)(v_{1},b_{1}) to (v,b)(v,b), yielding a local energy solution (still denoted (v,b)(v,b)) on [0,3T0/2][0,3T_{0}/2] that lies in 𝐋𝐄q(t0,t0+T0)𝐋𝐄q(0,T0){\bf LE}_{q}(t_{0},t_{0}+T_{0})\cap{\bf LE}_{q}(0,T_{0}). Hence, (v,b)𝐋𝐄q(0,3T0/2)(v,b)\in{\bf LE}_{q}(0,3T_{0}/2). Repeating this argument, we obtain a solution (v,b)𝐋𝐄q(0,T)(v,b)\in{\bf LE}_{q}(0,T) for any T>0T>0, using the uniform-in-time control of (v,b)E2q×E2q\left\|(v,b)\right\|_{E^{2}_{q}\times E^{2}_{q}}.

Now consider the general case (u,a)(0,0)(u,a)\neq(0,0). Let (v0,b0)L2×L2E2q×E2q(v_{0},b_{0})\in L^{2}\times L^{2}\subset E^{2}_{q}\times E^{2}_{q}, and let (v,b)(v,b) be the local energy solution to the perturbed MHD equations, (3.30), given by Lemma 3.12. Assuming δ:=(u,a)L(L3uloc×L3uloc)c0\delta:=\left\|(u,a)\right\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}\mathrm{uloc})}\ll c_{0}, we have (v,b)𝐋𝐄q(0,T0)(v,b)\in{\bf LE}_{q}(0,T_{0}) for some T0=T0((v0,b0)E2q×E2q)T_{0}=T_{0}(\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}). We now derive an energy estimate. Using the following bounds:

3|u||v|(|v|+|v|)kB1(k)|u||v|(|v|+|v|)uL3ulockB1(k)(|v|2+|v|2),\int_{\mathbb{R}^{3}}|u||v|(|\nabla v|+|v|)\leq\sum_{k}\int_{B_{1}(k)}|u||v|(|\nabla v|+|v|){\ \lesssim\ }\left\|u\right\|_{L^{3}_{\mathrm{uloc}}}\sum_{k}\int_{B_{1}(k)}(|\nabla v|^{2}+|v|^{2}),
3|u||b|(|b|+|b|)kB1(k)|u||b|(|b|+|b|)uL3ulockB1(k)(|b|2+|b|2),\int_{\mathbb{R}^{3}}|u||b|(|\nabla b|+|b|)\leq\sum_{k}\int_{B_{1}(k)}|u||b|(|\nabla b|+|b|){\ \lesssim\ }\left\|u\right\|_{L^{3}_{\mathrm{uloc}}}\sum_{k}\int_{B_{1}(k)}(|\nabla b|^{2}+|b|^{2}),
3|a||v|(|b|+|b|)kB1(k)|a||v|(|b|+|b|)aL3ulockB1(k)(|b|2+|b|2+|v|2),\int_{\mathbb{R}^{3}}|a||v|(|\nabla b|+|b|)\leq\sum_{k}\int_{B_{1}(k)}|a||v|(|\nabla b|+|b|){\ \lesssim\ }\left\|a\right\|_{L^{3}_{\mathrm{uloc}}}\sum_{k}\int_{B_{1}(k)}(|\nabla b|^{2}+|b|^{2}+|v|^{2}),

and

3|a||b|(|v|+|v|)kB1(k)|a||b|(|v|+|v|)aL3ulockB1(k)(|v|2+|v|2+|b|2),\int_{\mathbb{R}^{3}}|a||b|(|\nabla v|+|v|)\leq\sum_{k}\int_{B_{1}(k)}|a||b|(|\nabla v|+|v|){\ \lesssim\ }\left\|a\right\|_{L^{3}_{\mathrm{uloc}}}\sum_{k}\int_{B_{1}(k)}(|\nabla v|^{2}+|v|^{2}+|b|^{2}),

we obtain the energy inequality:

(v,b)(t)L2×L22+2(v,b)L2(0,t;L2×L2)2(v0,b0)L2×L22+C(u,a)L(L3uloc×L3uloc)((v,b)L2(0,t;L2×L2)2+tsup0<s<t(v,b)(s)L2×L22),\begin{split}&\left\|(v,b)(t)\right\|_{L^{2}\times L^{2}}^{2}+2\left\|(\nabla v,\nabla b)\right\|_{L^{2}(0,t;L^{2}\times L^{2})}^{2}\\ &\qquad\leq\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}^{2}\\ &\qquad\quad+C\left\|(u,a)\right\|_{L^{\infty}(L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}})}\left(\left\|(\nabla v,\nabla b)\right\|_{L^{2}(0,t;L^{2}\times L^{2})}^{2}+t\sup_{0<s<t}\left\|(v,b)(s)\right\|_{L^{2}\times L^{2}}^{2}\right),\end{split} (3.61)

for 0<t<T0<t<T. If TT0T\leq T_{0}, the result follows. Otherwise, we choose δ\delta sufficiently small (depending on TT) to absorb the right-hand side, yielding:

sup0<t<T(v,b)(t)L2×L22+2(v,b)L2(0,T;L2×L2)22(v0,b0)L2×L22.\sup_{0<t<T}\left\|(v,b)(t)\right\|_{L^{2}\times L^{2}}^{2}+2\left\|(\nabla v,\nabla b)\right\|_{L^{2}(0,T;L^{2}\times L^{2})}^{2}\leq 2\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}^{2}.

This in turn implies

sup0<t<T(v,b)(t)E2q×E2qC(v0,b0)L2×L2.\sup_{0<t<T}\left\|(v,b)(t)\right\|_{E^{2}_{q}\times E^{2}_{q}}\leq C\left\|(v_{0},b_{0})\right\|_{L^{2}\times L^{2}}.

Thus, we obtain uniform-in-time control of (v,b)(t)E2q×E2q\left\|(v,b)(t)\right\|_{E^{2}_{q}\times E^{2}_{q}} and the argument from the (u,a)=(0,0)(u,a)=(0,0) case applies to yield the desired result. ∎

Lemma 3.14.

Suppose that v0,b0E4v_{0},b_{0}\in E^{4}_{\infty} are divergence free. Assume also that δ:=(v0,b0)E4×E4<ϵ\delta:=\left\|(v_{0},b_{0})\right\|_{E^{4}_{\infty}\times E^{4}_{\infty}}<\epsilon_{*} for a universal constant ϵ\epsilon_{*}. Then there exists a second universal constant τ0>0\tau_{0}>0 and (v,b)(v,b) and π\pi comprising a local energy solution to (MHD) in 3×(0,τ0)\mathbb{R}^{3}\times(0,\tau_{0}) with initial data (v0,b0)(v_{0},b_{0}) so that (v,b)(v,b) and π\pi are smooth in space and time, (v,b)C([0,τ0];E4×E4)(v,b)\in C([0,\tau_{0}];E^{4}_{\infty}\times E^{4}_{\infty}), and

sup0tτ0(v,b)(t)L4uloc×L4uloc<Cδ.\sup_{0\leq t\leq\tau_{0}}\left\|(v,b)(t)\right\|_{L^{4}_{\mathrm{uloc}}\times L^{4}_{\mathrm{uloc}}}<C\delta.

Furthermore, if (u,a)𝒩MHD(v0,b0)(u,a)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}), then (u,a)=(v,b)(u,a)=(v,b) on 3×[0,τ0]\mathbb{R}^{3}\times[0,\tau_{0}].

Proof.

The proof of the lemma is an adaption of the proof of [4, Lemma A.1] for the Navier–Stokes equations to the MHD equations.

Since (v0,b0)E4×E4(v_{0},b_{0})\in E^{4}_{\infty}\times E^{4}_{\infty}, it follows that (v0,b0)E2×E2(v_{0},b_{0})\in E^{2}\times E^{2} as E4E2E^{4}_{\infty}\subset E^{2}. Viewing the MHD equations as a coupled system of inhomogeneous Stokes systems, we apply the linear theory from [21, §5] and follow the argument of Theorem 1.5 therein to construct a global-in-time local energy solution (v,b)(v,b) evolving from (v0,b0)(v_{0},b_{0}). We may assume (v0,b0)L3uloc×L3uloc<ϵ3\left\|(v_{0},b_{0})\right\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}<\epsilon_{3}, where ϵ3\epsilon_{3} is given in Theorem 3.4. Then, for all x03x_{0}\in\mathbb{R}^{3} and r1r\leq 1,

1rBr(x0)(|v0|2+|b0|2)dxC(Br(x0)(|v0|3+|b0|3)dx)Cϵ3.\frac{1}{r}\int_{B_{r}(x_{0})}(|v_{0}|^{2}+|b_{0}|^{2})\,dx\leq C\left(\int_{B_{r}(x_{0})}(|v_{0}|^{3}+|b_{0}|^{3})\,dx\right)\leq C\epsilon_{3}.

Thus, Theorem 3.4 ensures uniqueness of the local energy solution (v,b)𝒩MHD(v0,b0)(v,b)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}) up to some time τ0\tau_{0}.

Now consider the mild solution constructed in Theorem 1.1 with r=4r=4, q=q=\infty. Since E44ulocE^{4}\subset\mathcal{L}^{4}_{\mathrm{uloc}}, the closure of BUC(3)BUC(\mathbb{R}^{3}) in the L4ulocL^{4}_{\mathrm{uloc}} norm, there exists a time T>0T>0 and a unique mild solution (u,a)C([0,T);L4uloc)(u,a)\in C([0,T);L^{4}_{\mathrm{uloc}}). In the construction of this mild solution (see Section 2.1), we may redefine the space T\mathcal{E}_{T} with the norm (v,b)T=sup0<t<T(v,b)(t)Erq+sup0<t<Tt3/(2r)(v,b)(t)L\left\|(v,b)\right\|_{\mathcal{E}_{T}}=\sup_{0<t<T}\left\|(v,b)(t)\right\|_{E^{r}_{q}}+\sup_{0<t<T}t^{3/(2r)}\left\|(v,b)(t)\right\|_{L^{\infty}} to ensure that the mild solution (u(t),a(t))L×L(u(t),a(t))\in L^{\infty}\times L^{\infty} for all t>0t>0. Then, by adapting the regularity argument from [14, §4], we conclude that (u,a)(u,a) is smooth in both space and time for all t>0t>0. By choosing ϵ\epsilon_{*} sufficiently small, the existence time TT for the mild solution in Theorem 1.1 exceeds τ0\tau_{0}.

We now verify that (u,a)(u,a) defines a local energy solution. By embeddings, the same convergence properties at t=0t=0 hold with L4L^{4} and L4ulocL^{4}_{\mathrm{uloc}} replaced by L2L^{2} and L2ulocL^{2}_{\mathrm{uloc}}, respectively. This implies that if wL2(3)w\in L^{2}(\mathbb{R}^{3}) is compactly supported, then

limt0(u(x,t)u0(x))w(x)dx=limt0(a(x,t)a0(x))w(x)dx=0.\lim_{t\to 0}\int(u(x,t)-u_{0}(x))w(x)\,dx=\lim_{t\to 0}\int(a(x,t)-a_{0}(x))w(x)\,dx=0.

Moreover, the maps

tu(x,t)w(x)dx and ta(x,t)w(x)dxt\mapsto\int u(x,t)\cdot w(x)\,dx\quad\text{ and }\quad t\mapsto\int a(x,t)\cdot w(x)\,dx

are continuous for t>0t>0 due to the smoothness of (u,a)(u,a).

To complete the verification, we adapt the pressure construction from [5, Theorem 1.4], as carried out in [5, §6] for the Navier–Stokes equations. This yields a pressure π\pi such that (u,a,π)(u,a,\pi) satisfies the MHD equations in the distributional sense. The local expansion of π\pi also guarantees that πL3/2loc(3×(0,T))\pi\in L^{3/2}_{\mathrm{loc}}(\mathbb{R}^{3}\times(0,T)). The local energy inequality follows from the space-time smoothness of (u,a)(u,a), and item 2 in the definition of local energy solution is satisfied since (u,a)LL3uloc(u,a)\in L^{\infty}L^{3}_{\mathrm{uloc}}. Hence, (u,a)𝒩MHD(v0,b0)(u,a)\in\mathcal{N}_{\rm MHD}(v_{0},b_{0}), and uniqueness implies (v,b)=(u,a)(v,b)=(u,a) on 3×(0,τ0)\mathbb{R}^{3}\times(0,\tau_{0}). Therefore,

(v,b)(t)L3uloc×L3ulocC(v,b)(t)L4uloc×L4uloc<Cδ\left\|(v,b)(t)\right\|_{L^{3}_{\mathrm{uloc}}\times L^{3}_{\mathrm{uloc}}}\leq C\left\|(v,b)(t)\right\|_{L^{4}_{\mathrm{uloc}}\times L^{4}_{\mathrm{uloc}}}<C\delta

for all t(0,τ0)t\in(0,\tau_{0}).

The mild solution constructed in Theorem 1.1 satisfies (v,b)C([0,τ0];L4uloc×L4uloc)(v,b)\in C([0,\tau_{0}];L^{4}_{\mathrm{uloc}}\times L^{4}_{\mathrm{uloc}}). Since L4uloc=E4E2L^{4}_{\mathrm{uloc}}=E^{4}_{\infty}\subset E^{2}_{\infty}, we may apply Lemma 3.8 with q=q=\infty and data (v0,b0)E4×E4E2×E2(v_{0},b_{0})\in E^{4}_{\infty}\times E^{4}_{\infty}\subset E^{2}_{\infty}\times E^{2}_{\infty} to conclude that for almost every t>0t>0, (v,b)(t)E4×E4(v,b)(t)\in E^{4}_{\infty}\times E^{4}_{\infty}. This further implies (v,b)C([0,τ0];E4×E4)(v,b)\in C([0,\tau_{0}];E^{4}_{\infty}\times E^{4}_{\infty}). ∎

Proof of Theorem 1.10 for q2q\geq 2.

The proof of Theorem 1.10 for q2q\geq 2 is an adaption of the proof of [4, Theorem 1.5] for the Navier–Stokes equations to the MHD equations.

Assume that v0,b0E2qv_{0},b_{0}\in E^{2}_{q} are divergence free. By Lemma 3.12 with (u,a)=(0,0)(u,a)=(0,0), there exists a local energy solution (v,b)(v,b) to the MHD equations on 3×(0,T0)\mathbb{R}^{3}\times(0,T_{0}) such that (v,b)𝐋𝐄q(0,T0)(v,b)\in{\bf LE}_{q}(0,T_{0}). Moreover, by Lemma 3.8, we have

(v,b)(t)E4q×E4q for a.e. t[T0/2,T0].(v,b)(t)\in E^{4}_{q}\times E^{4}_{q}\ \text{ for a.e. }t\in[T_{0}/2,T_{0}].

Choose a time t0>T0/2t_{0}>T_{0}/2 so that (v,b)(t0)E4q×E4q(v,b)(t_{0})\in E^{4}_{q}\times E^{4}_{q}. We aim to construct a local energy solution in 𝐋𝐄q(t0,t0+τ0){\bf LE}_{q}(t_{0},t_{0}+\tau_{0}) with initial data (v,b)(t0)E4q×E4q(v,b)(t_{0})\in E^{4}_{q}\times E^{4}_{q}, where τ0\tau_{0} is the fixed time-scale in Lemma 3.14. Using the Bogovskii map (see [40] for the details), for any δ>0\delta>0, we can decompose the initial data as

(v,b)(t0)=(u0,a0)+(w0,d0),w0=d0=0,(v,b)(t_{0})=(u_{0},a_{0})+(w_{0},d_{0}),\qquad\nabla\cdot w_{0}=\nabla\cdot d_{0}=0,

where

(u0,a0)E4q×E4q<δ,w0,d0L2(3).\left\|(u_{0},a_{0})\right\|_{E^{4}_{q}\times E^{4}_{q}}<\delta,\qquad w_{0},d_{0}\in L^{2}(\mathbb{R}^{3}).

By Lemma 3.14, choosing δ\delta sufficiently small ensures the existence of a local energy solution (u,a)(u,a) with pressure π\pi, defined on 3×(t0,t0+τ0)\mathbb{R}^{3}\times(t_{0},t_{0}+\tau_{0}), evolving from the initial data (u0,a0)(u_{0},a_{0}), which is smooth in both space and time and satisfies

supt0tt0+τ0(u,a)(t)L4uloc×L4uloc<Cδ.\sup_{t_{0}\leq t\leq t_{0}+\tau_{0}}\left\|(u,a)(t)\right\|_{L^{4}_{\mathrm{uloc}}\times L^{4}_{\mathrm{uloc}}}<C\delta.

Furthermore, by the uniqueness result in Lemma 3.14, this solution (u,a)(u,a) coincides with the one given by Lemma 3.12, and hence (u,a)𝐋𝐄q(t0,t0+τ0)(u,a)\in{\bf LE}_{q}(t_{0},t_{0}+\tau_{0}).

Next, we apply Lemma 3.13 with the perturbation factor (u,a)(u,a), again choosing δ\delta sufficiently small to ensure that the time-scale it yields is at least τ0\tau_{0}. This gives a local energy solution (w,d)𝐋𝐄q(t0,t0+τ0)(w,d)\in{\bf LE}_{q}(t_{0},t_{0}+\tau_{0}) to the the perturbed MHD equation with initial data (w0,d0)(w_{0},d_{0}) and associated pressure π\pi. Define (v1,b1):=(u,a)+(w,d)(v_{1},b_{1}):=(u,a)+(w,d). This gives a local energy solution on 3×(t0,t0+τ0)\mathbb{R}^{3}\times(t_{0},t_{0}+\tau_{0}). To very the local energy inequality for (v1,b1)(v_{1},b_{1}), we use approximations (w(n),d(n))(w,d)(w^{(n)},d^{(n)})\to(w,d), as in the proof of Lemma 3.12, and apply the inequality to (u,a)+(w(n),d(n))(u,a)+(w^{(n)},d^{(n)}).

Since (v,b)(v,b) and (v1,b1)(v_{1},b_{1}) coincide at t0t_{0}, and since (v,b)(t0)E3×E3(v,b)(t_{0})\in E^{3}\times E^{3} since E4E3E^{4}_{\infty}\subset E^{3}, Corollary 3.5 implies that (v,b)(x,t)=(v1,b1)(x,t)(v,b)(x,t)=(v_{1},b_{1})(x,t) on 3×(t0,t0+γ)\mathbb{R}^{3}\times(t_{0},t_{0}+\gamma) for some γ>0\gamma>0. Thus, we may glue (v1,b1)(v_{1},b_{1}) to (v,b)(v,b) to extend the solution to 𝐋𝐄q(0,t0+τ0){\bf LE}_{q}(0,t_{0}+\tau_{0}). Repeating this procedure nn times yields a solution (v,b)𝐋𝐄q(0,t0+nτ0)(v,b)\in{\bf LE}_{q}(0,t_{0}+n\tau_{0}). Taking the limit nn\to\infty, we obtain the global-in-time local energy solution asserted in Theorem 1.10 for q2q\geq 2. ∎

3.3.2 The case 1q<21\leq q<2

Now, we consider the case 1q<21\leq q<2 and look at the localized and regularized MHD equations, (3.31). The following lemma corresponds to [1, Lemma 4.6] for the Navier–Stokes equations.

Lemma 3.15.

Let q1q\geq 1. For each 0<ϵ<10<\epsilon<1 and divergence free v0,b0v_{0},b_{0} with v0E2qB\left\|v_{0}\right\|_{E^{2}_{q}}\leq B, b0E2qB\left\|b_{0}\right\|_{E^{2}_{q}}\leq B, if 0<T<min(1,cϵ3B2)0<T<\min\left(1,\,c\epsilon^{3}B^{-2}\right), we can find a unique solution (v,b)=(vϵ,bϵ)(v,b)=(v^{\epsilon},b^{\epsilon}) to the integral form of (3.31)

v(t)=etΔv00te(tτ)Δ(𝒥ϵ(v)vΦϵ𝒥ϵ(b)bΦϵ)(τ)dτ,b(t)=etΔb00te(tτ)Δ(𝒥ϵ(v)bΦϵ𝒥ϵ(b)vΦϵ)(τ)dτ,\begin{split}v(t)&=e^{t\Delta}v_{0}-\int_{0}^{t}e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot\left(\mathcal{J}_{\epsilon}(v)\otimes v\Phi_{\epsilon}-\mathcal{J}_{\epsilon}(b)\otimes b\Phi_{\epsilon}\right)(\tau)\,d\tau,\\ b(t)&=e^{t\Delta}b_{0}-\int_{0}^{t}e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot\left(\mathcal{J}_{\epsilon}(v)\otimes b\Phi_{\epsilon}-\mathcal{J}_{\epsilon}(b)\otimes v\Phi_{\epsilon}\right)(\tau)\,d\tau,\end{split} (3.62)

satisfying

(v,b)𝐋𝐄q(0,T)2C0B,\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}\leq 2C_{0}B,

where c>0c>0 and C0>1C_{0}>1 are absolute constants.

Proof.

The proof of the lemma is an adaption of the proof of [1, Lemma 4.6] for the Navier–Stokes equations to the MHD equations.

Let Ψ(v,b)\Psi(v,b) denote the mapping defined by the right-hand side of (3.62) for (v,b)𝐋𝐄q(0,T)(v,b)\in{\bf LE}_{q}(0,T). By [1, Lemma 2.9] and the assumption T1T\leq 1, we obtain the estimate

Ψ(v,b)𝐋𝐄q(0,T)v0E2q+𝒥ϵ(v)vΦϵET,q2,2+𝒥ϵ(b)bΦϵET,q2,2+b0E2q+𝒥ϵ(v)bΦϵET,q2,2+𝒥ϵ(b)vΦϵET,q2,2v0E2q+𝒥ϵ(v)L(0,T;L(3))vET,q2,2+𝒥ϵ(b)L(0,T;L(3))bET,q2,2+b0E2q+𝒥ϵ(v)L(0,T;L(3))bET,q2,2+𝒥ϵ(b)L(0,T;L(3))vET,q2,2v0E2q+b0E2q+ϵ32T(vET,q,22+bET,q,22).\begin{split}\left\|\Psi(v,b)\right\|_{{\bf LE}_{q}(0,T)}&{\ \lesssim\ }\left\|v_{0}\right\|_{E^{2}_{q}}+\left\|\mathcal{J}_{\epsilon}(v)\otimes v\Phi_{\epsilon}\right\|_{E_{T,q}^{2,2}}+\left\|\mathcal{J}_{\epsilon}(b)\otimes b\Phi_{\epsilon}\right\|_{E_{T,q}^{2,2}}\\ &\quad+\left\|b_{0}\right\|_{E^{2}_{q}}+\left\|\mathcal{J}_{\epsilon}(v)\otimes b\Phi_{\epsilon}\right\|_{E_{T,q}^{2,2}}+\left\|\mathcal{J}_{\epsilon}(b)\otimes v\Phi_{\epsilon}\right\|_{E_{T,q}^{2,2}}\\ &{\ \lesssim\ }\left\|v_{0}\right\|_{E^{2}_{q}}+\left\|\mathcal{J}_{\epsilon}(v)\right\|_{L^{\infty}(0,T;L^{\infty}(\mathbb{R}^{3}))}\left\|v\right\|_{E_{T,q}^{2,2}}+\left\|\mathcal{J}_{\epsilon}(b)\right\|_{L^{\infty}(0,T;L^{\infty}(\mathbb{R}^{3}))}\left\|b\right\|_{E_{T,q}^{2,2}}\\ &\quad+\left\|b_{0}\right\|_{E^{2}_{q}}+\left\|\mathcal{J}_{\epsilon}(v)\right\|_{L^{\infty}(0,T;L^{\infty}(\mathbb{R}^{3}))}\left\|b\right\|_{E_{T,q}^{2,2}}+\left\|\mathcal{J}_{\epsilon}(b)\right\|_{L^{\infty}(0,T;L^{\infty}(\mathbb{R}^{3}))}\left\|v\right\|_{E_{T,q}^{2,2}}\\ &{\ \lesssim\ }\left\|v_{0}\right\|_{E^{2}_{q}}+\left\|b_{0}\right\|_{E^{2}_{q}}+\epsilon^{-\frac{3}{2}}\sqrt{T}\left(\left\|v\right\|_{E_{T,q}^{\infty,2}}^{2}+\left\|b\right\|_{E_{T,q}^{\infty,2}}^{2}\right).\end{split}

Therefore, for some constants C0,C1>0C_{0},C_{1}>0,

Ψ(v,b)𝐋𝐄q(0,T)C0(v0,b0)E2q×E2q+C1ϵ32T(v,b)𝐋𝐄q(0,T)2,\left\|\Psi(v,b)\right\|_{{\bf LE}_{q}(0,T)}\leq C_{0}\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}+C_{1}\epsilon^{-\frac{3}{2}}\sqrt{T}\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}^{2},

To estimate the difference, let (v,b),(u,a)𝐋𝐄q(0,T)(v,b),(u,a)\in{\bf LE}_{q}(0,T). Then

Ψ(v,b)Ψ(u,a)𝐋𝐄q(0,T)C1ϵ32T((v,b)𝐋𝐄q(0,T)+(u,a)𝐋𝐄q(0,T))(v,b)(u,a)𝐋𝐄q(0,T).\begin{split}&\left\|\Psi(v,b)-\Psi(u,a)\right\|_{{\bf LE}_{q}(0,T)}\\ &\qquad\leq C_{1}\epsilon^{-\frac{3}{2}}\sqrt{T}\left(\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}+\left\|(u,a)\right\|_{{\bf LE}_{q}(0,T)}\right)\left\|(v,b)-(u,a)\right\|_{{\bf LE}_{q}(0,T)}.\end{split}

Applying the Picard contraction principle, we see that if the time TT satisfies T<ϵ364(C0C1B)2=cϵ3B2T<\frac{\epsilon^{3}}{64(C_{0}C_{1}B)^{2}}=c\epsilon^{3}B^{-2}, then Ψ\Psi has a unique fixed point (v,b)𝐋𝐄q(0,T)(v,b)\in{\bf LE}_{q}(0,T) with (v,b)𝐋𝐄q(0,T)2C0B\left\|(v,b)\right\|_{{\bf LE}_{q}(0,T)}\leq 2C_{0}B, solving the integral system (3.62). ∎

Lemma 3.16.

Let v0,b0E2qv_{0},b_{0}\in E^{2}_{q}, q1q\geq 1, be divergence free. For each ϵ(0,1)\epsilon\in(0,1), we can find (vϵ,bϵ)(v^{\epsilon},b^{\epsilon}) in 𝐋𝐄q(0,T){\bf LE}_{q}(0,T) and πϵ\pi^{\epsilon} in L(0,T;L2(3))L^{\infty}(0,T;L^{2}(\mathbb{R}^{3})) for some positive T=T(ϵ,(v0,b0)E2q×E2q)T=T(\epsilon,\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}) which solve the localized and regularized MHD equations, (3.31), in the sense of distributions and (vϵ,bϵ)(t)(v0,b0)(v^{\epsilon},b^{\epsilon})(t)\to(v_{0},b_{0}) in L2(E)×L2(E)L^{2}(E)\times L^{2}(E) as t0+t\to 0^{+} for any compact subset EE of 3\mathbb{R}^{3}.

Proof.

The proof of the lemma is an adaption of the proof of [1, Lemma 4.7] for the Navier–Stokes equations to the MHD equations. We provide the corresponding details by following the same logic used in the proof of [23, Lemma 3.4] for the Navier–Stokes equations, adapting it from the L2ulocL^{2}_{\mathrm{uloc}} framework to the E2qE^{2}_{q} setting, and from the Navier–Stokes equations to the MHD equations.

By Lemma 3.15, there is a mild solution (vϵ,bϵ)𝐋𝐄q(0,T)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,T) to (3.62) for T=T(ϵ,(v0,b0)E2q×E2q)T=T(\epsilon,\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}). Apparently,

vϵetΔv0E,2t,q=0te(tτ)Δ(𝒥ϵ(v)vΦϵ𝒥ϵ(b)bΦϵ)(τ)dτE,2t,q𝒥ϵ(v)vΦϵE2,2t,q+𝒥ϵ(b)bΦϵE2,2t,qϵ3/2t(vE,2t,q2+bE,2t,q2),\begin{split}\left\|v^{\epsilon}-e^{t\Delta}v_{0}\right\|_{E^{\infty,2}_{t^{\prime},q}}&=\left\|\int_{0}^{t}e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot\left(\mathcal{J}_{\epsilon}(v)\otimes v\Phi_{\epsilon}-\mathcal{J}_{\epsilon}(b)\otimes b\Phi_{\epsilon}\right)(\tau)\,d\tau\right\|_{E^{\infty,2}_{t^{\prime},q}}\\ &{\ \lesssim\ }\left\|\mathcal{J}_{\epsilon}(v)\otimes v\Phi_{\epsilon}\right\|_{E^{2,2}_{t^{\prime},q}}+\left\|\mathcal{J}_{\epsilon}(b)\otimes b\Phi_{\epsilon}\right\|_{E^{2,2}_{t^{\prime},q}}\\ &{\ \lesssim\ }\epsilon^{-3/2}\sqrt{t^{\prime}}\left(\left\|v\right\|_{E^{\infty,2}_{t^{\prime},q}}^{2}+\left\|b\right\|_{E^{\infty,2}_{t^{\prime},q}}^{2}\right),\end{split}

and

bϵetΔb0E,2t,q=0te(tτ)Δ(𝒥ϵ(v)bΦϵ𝒥ϵ(b)vΦϵ)(τ)dτE,2t,q𝒥ϵ(v)bΦϵE2,2t,q+𝒥ϵ(b)vΦϵE2,2t,qϵ3/2tvbE,2t,qϵ3/2t(vE,2t,q2+bE,2t,q2),\begin{split}\left\|b^{\epsilon}-e^{t\Delta}b_{0}\right\|_{E^{\infty,2}_{t^{\prime},q}}&=\left\|\int_{0}^{t}e^{(t-\tau)\Delta}\mathbb{P}\nabla\cdot\left(\mathcal{J}_{\epsilon}(v)\otimes b\Phi_{\epsilon}-\mathcal{J}_{\epsilon}(b)\otimes v\Phi_{\epsilon}\right)(\tau)\,d\tau\right\|_{E^{\infty,2}_{t^{\prime},q}}\\ &{\ \lesssim\ }\left\|\mathcal{J}_{\epsilon}(v)\otimes b\Phi_{\epsilon}\right\|_{E^{2,2}_{t^{\prime},q}}+\left\|\mathcal{J}_{\epsilon}(b)\otimes v\Phi_{\epsilon}\right\|_{E^{2,2}_{t^{\prime},q}}\\ &{\ \lesssim\ }\epsilon^{-3/2}\sqrt{t^{\prime}}\left\|vb\right\|_{E^{\infty,2}_{t^{\prime},q}}{\ \lesssim\ }\epsilon^{-3/2}\sqrt{t^{\prime}}\left(\left\|v\right\|_{E^{\infty,2}_{t^{\prime},q}}^{2}+\left\|b\right\|_{E^{\infty,2}_{t^{\prime},q}}^{2}\right),\end{split}

where we have used [1, Lemma 2.9] and assumed tT1t^{\prime}\leq T\leq 1. Also, for any compact subset EE of 3\mathbb{R}^{3}, we have etΔv0v0L2(E)\left\|e^{t\Delta}v_{0}-v_{0}\right\|_{L^{2}(E)}, etΔb0b0L2(E)0\left\|e^{t\Delta}b_{0}-b_{0}\right\|_{L^{2}(E)}\to 0 as tt goes to 0 by Legesgue’s convergence theorem. Then, it follows that limt0+vϵ(t)v0L2(E)=0\lim_{t\to 0^{+}}\left\|v^{\epsilon}(t)-v_{0}\right\|_{L^{2}(E)}=0 and limt0+bϵ(t)b0L2(E)=0\lim_{t\to 0^{+}}\left\|b^{\epsilon}(t)-b_{0}\right\|_{L^{2}(E)}=0 for any compact subset EE of 3\mathbb{R}^{3}.

Note that both etΔv0e^{t\Delta}v_{0} and etΔb0e^{t\Delta}b_{0}, with v0,b0E2qv_{0},b_{0}\in E^{2}_{q}, solve the homogeneous heat equation in the distribution sense. Also, using v0=b0=0\nabla\cdot v_{0}=\nabla\cdot b_{0}=0, we can easily see that etΔv0=etΔb0=0\nabla\cdot e^{t\Delta}v_{0}=\nabla\cdot e^{t\Delta}b_{0}=0.

On the other hand, 𝒥ϵ(vϵ),𝒥ϵ(bϵ)L(3×[0,T])\mathcal{J}_{\epsilon}(v^{\epsilon}),\mathcal{J}_{\epsilon}(b^{\epsilon})\in L^{\infty}(\mathbb{R}^{3}\times[0,T]) and (vϵ,bϵ)𝐋𝐄q(0,T)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,T) imply

𝒥ϵ(vϵ)vϵΦϵ,𝒥ϵ(bϵ)bϵΦϵ,𝒥ϵ(vϵ)bϵΦϵ,𝒥ϵ(bϵ)vϵΦϵL(0,T;L2(3)).\mathcal{J}_{\epsilon}(v^{\epsilon})\otimes v^{\epsilon}\Phi_{\epsilon},\quad\mathcal{J}_{\epsilon}(b^{\epsilon})\otimes b^{\epsilon}\Phi_{\epsilon},\quad\mathcal{J}_{\epsilon}(v^{\epsilon})\otimes b^{\epsilon}\Phi_{\epsilon},\quad\mathcal{J}_{\epsilon}(b^{\epsilon})\otimes v^{\epsilon}\Phi_{\epsilon}\in L^{\infty}(0,T;L^{2}(\mathbb{R}^{3})).

Hence by the classical theory, uϵ=vϵetΔv0u^{\epsilon}=v^{\epsilon}-e^{t\Delta}v_{0} and πϵ\pi^{\epsilon} defined by

πϵ=(Δ)1ij[(𝒥ϵ(vϵi)vϵj𝒥ϵ(bϵi)bϵj)Φϵ]L(0,T;L2(3)),\pi^{\epsilon}=(-\Delta)^{-1}\partial_{i}\partial_{j}\left[\left(\mathcal{J}_{\epsilon}(v^{\epsilon}_{i})v^{\epsilon}_{j}-\mathcal{J}_{\epsilon}(b^{\epsilon}_{i})b^{\epsilon}_{j}\right)\Phi_{\epsilon}\right]\in L^{\infty}(0,T;L^{2}(\mathbb{R}^{3})),

solves the Stokes system with the source term [(𝒥ϵ(vϵ)vϵ𝒥ϵ(bϵ)bϵ)Φϵ]\nabla\cdot\left[\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\otimes v^{\epsilon}-\mathcal{J}_{\epsilon}(b^{\epsilon})\otimes b^{\epsilon}\right)\Phi_{\epsilon}\right] in the distribution sense. Moreover, aϵ=bϵetΔb0a^{\epsilon}=b^{\epsilon}-e^{t\Delta}b_{0} solves the forced heat equation with the forcing [(𝒥ϵ(vϵ)bϵ𝒥ϵ(bϵ)vϵ)Φϵ]\nabla\cdot\left[\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\otimes b^{\epsilon}-\mathcal{J}_{\epsilon}(b^{\epsilon})\otimes v^{\epsilon}\right)\Phi_{\epsilon}\right] in the distribution sense. By adding the homogeneous heat equation for etΔv0e^{t\Delta}v_{0} with etΔv0=0\nabla\cdot e^{t\Delta}v_{0}=0 and the Stokes system for uϵu^{\epsilon} and πϵ\pi^{\epsilon}, vϵ=etΔv0+uϵv^{\epsilon}=e^{t\Delta}v_{0}+u^{\epsilon} satisfies

tvϵΔvϵ+(𝒥ϵ(vϵ))(vϵΦϵ)(𝒥ϵ(bϵ))(bϵΦϵ)+πϵ=0\partial_{t}v^{\epsilon}-\Delta v^{\epsilon}+\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\cdot\nabla\right)(v^{\epsilon}\Phi_{\epsilon})-\left(\mathcal{J}_{\epsilon}(b^{\epsilon})\cdot\nabla\right)(b^{\epsilon}\Phi_{\epsilon})+\nabla\pi^{\epsilon}=0

in the sense of distribution. Moreover, by adding the homogeneous heat equation for etΔb0e^{t\Delta}b_{0} with etΔv0=0\nabla\cdot e^{t\Delta}v_{0}=0 and the forced heat equation for aϵa^{\epsilon}, bϵ=etΔb0+aϵb^{\epsilon}=e^{t\Delta}b_{0}+a^{\epsilon} satisfies

tbϵΔbϵ+(𝒥ϵ(vϵ))(bϵΦϵ)(𝒥ϵ(bϵ))(vϵΦϵ)=0\partial_{t}b^{\epsilon}-\Delta b^{\epsilon}+\left(\mathcal{J}_{\epsilon}(v^{\epsilon})\cdot\nabla\right)(b^{\epsilon}\Phi_{\epsilon})-\left(\mathcal{J}_{\epsilon}(b^{\epsilon})\cdot\nabla\right)(v^{\epsilon}\Phi_{\epsilon})=0

in the sense of distribution. ∎

We next show global existence for the localized and regularized MHD equations, (3.31).

Lemma 3.17.

Assume v0,b0E2qv_{0},b_{0}\in E^{2}_{q}, 1q<21\leq q<2, are divergence free, and fix ϵ(0,1)\epsilon\in(0,1). Then, there exists a global solution (vϵ,bϵ,πϵ)(v^{\epsilon},b^{\epsilon},\pi^{\epsilon}) to the localized and regularized MHD equations, (3.31), such that (vϵ,bϵ)𝐋𝐄q(0,T)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,T) for any T<T<\infty, and (vϵ,bϵ,πϵ)(v^{\epsilon},b^{\epsilon},\pi^{\epsilon}) satisfies the a priori bounds (3.18) and (3.20) for all R=nR=n\in\mathbb{N} up to time Tn=λ0n2min(1,n2(c3(v0,b0)E2q×E2q2)2)T_{n}=\lambda_{0}n^{2}\min\left(1,\,n^{2}(c_{3}\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}^{2})^{-2}\right) for some c3>0c_{3}>0, where λ0\lambda_{0} is given in Lemma 3.7.

Proof.

The proof of the lemma is an adaption of the proof of [1, Lemma 4.8] for the Navier–Stokes equations to the MHD equations.

We set the radius R=nR=n\in\mathbb{N}. By (3.28), we have the bound A0,q(n)c3(v0E2q2+b0E2q2)A_{0,q}(n)\leq c_{3}(\left\|v_{0}\right\|_{E^{2}_{q}}^{2}+\left\|b_{0}\right\|_{E^{2}_{q}}^{2}) for all nn\in\mathbb{N}. Define

Tn=λ0n2min{1,n2(c3(v0E2q2+b0E2q2))2}λnn2,T_{n}=\lambda_{0}n^{2}\min\left\{1,n^{2}\left(c_{3}(\left\|v_{0}\right\|_{E^{2}_{q}}^{2}+\left\|b_{0}\right\|_{E^{2}_{q}}^{2})\right)^{-2}\right\}\leq\lambda_{n}n^{2},

where the constants λ0\lambda_{0} and λn\lambda_{n} are as in Lemma 3.7. Note that TnT_{n} is increasing and TnT_{n}\to\infty as nn\to\infty. Now, by the same argument used in the proof of Lemma 3.7, if a solution (vϵ,bϵ,πϵ)(v^{\epsilon},b^{\epsilon},\pi^{\epsilon}) of (3.31) satisfies (vϵ,bϵ)𝐋𝐄q(0,T)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,T), then it satisfies the a priori bounds (3.18) and (3.20) on the interval [0,min(T,Tn)][0,\min(T,T_{n})] with radius R=nR=n.

Since the system (3.31) is a coupled system of inhomogeneous Stokes systems with localized and regularized forcing, standard theory guarantees the existence of unique global solution (vϵ,bϵ,πϵ)(v^{\epsilon},b^{\epsilon},\pi^{\epsilon}). By uniqueness, this solution agrees with the 𝐋𝐄q{\bf LE}_{q}-solution constructed in Lemma 3.16, and thus (vϵ,bϵ)𝐋𝐄q(0,τ)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,\tau) for some τ=τ(ϵ,(v0,b0)E2q×E2q)>0\tau=\tau(\epsilon,\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}})>0. Fix nn\in\mathbb{N}. Applying (3.18) with R=nR=n, we obtain (vϵ,bϵ)(τ)E2q×E2qC(n)(v0,b0)E2q×E2q\left\|(v^{\epsilon},b^{\epsilon})(\tau)\right\|_{E^{2}_{q}\times E^{2}_{q}}\leq C(n)\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}}. Then, by Lemma 3.16, there exists an 𝐋𝐄q{\bf LE}_{q}-solution on (τ,τ+τ1)(\tau,\tau+\tau_{1}) for some τ1=τ1(ϵ,C(n)(v0,b0)E2q×E2q)>0\tau_{1}=\tau_{1}(\epsilon,C(n)\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}})>0. By uniqueness, this solution coincides with (vϵ,bϵ)(v^{\epsilon},b^{\epsilon}), and we conclude that (vϵ,bϵ)𝐋𝐄q(0,τ+τ1)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,\tau+\tau_{1}), with the a priori bound (3.18) valid up to time τ+τ1\tau+\tau_{1}. This argument can be iterated: by repeatedly extending the solution, we obtain (vϵ,bϵ)𝐋𝐄q(0,τ+kτ1)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,\tau+k\tau_{1}) for kk\in\mathbb{N}, until τ+kτ1Tn\tau+k\tau_{1}\geq T_{n}. Thus, for each nn\in\mathbb{N}, we obtain (vϵ,bϵ)𝐋𝐄q(0,Tn)(v^{\epsilon},b^{\epsilon})\in{\bf LE}_{q}(0,T_{n}), with the a priori bound (3.18) holding for R=nR=n up to time TnT_{n}. Since TnT_{n}\to\infty, the lemma follows. ∎

Proof of Theorem 1.10 for 1q<21\leq q<2.

The proof of Theorem 1.10 for 1q<21\leq q<2 is an adaption of the proof of [1, Theorem 1.4] for the Navier–Stokes equations to the MHD equations. We provide the necessary details by following the same stragegy as in the proof of [3, Theorem 1.5] for the Navier–Stokes equations, adapting the argument from the L2ulocL^{2}_{\mathrm{uloc}} framework to the E2qE^{2}_{q} setting, and from the Navier–Stokes to the MHD equations.

For kk\in\mathbb{N}, let (vk,bk,πk)(v^{k},b^{k},\pi^{k}) be the solution of the localized, regularized MHD equations, (3.31), with ϵ=1/k\epsilon=1/k, given in Lemma 3.17. They share the same a priori bound (3.18) for R=nR=n up to time TnT_{n}, thus

supk(vk,bk)𝐋𝐄q(0,Tn)<,n.\sup_{k\in\mathbb{N}}\left\|(v^{k},b^{k})\right\|_{{\bf LE}_{q}(0,T_{n})}<\infty,\qquad\forall n\in\mathbb{N}.

Using this a priori bound, we now construct the desired global solutions as the limit of vkv^{k} defined in (0,Tk)(0,T_{k}), TkT_{k}\to\infty by induction. Lemma 3.17 implies that (vϵ,bϵ)(v^{\epsilon},b^{\epsilon}) are uniformly bounded in the class from inequalities

sup0<t<T1B1(|vk|2+|bk|2)dx+0T1B1(|vk|2+|bk|2)dxdtcA\sup_{0<t<T_{1}}\int_{B_{1}}(|v^{k}|^{2}+|b^{k}|^{2})\,dx+\int_{0}^{T_{1}}\int_{B_{1}}(|\nabla v^{k}|^{2}+|\nabla b^{k}|^{2})\,dxdt\leq cA (3.63)
0T1B1(|vk|10/3+|bk|10/3)dxdtcA5/3,\int_{0}^{T_{1}}\int_{B_{1}}(|v^{k}|^{10/3}+|b^{k}|^{10/3})\,dxdt\leq cA^{5/3}, (3.64)
0T1B1|πk(x,t)c0,2k(t)|3/2dxdtC(T1,A),\int_{0}^{T_{1}}\int_{B_{1}}|\pi^{k}(x,t)-c_{0,2}^{k}(t)|^{3/2}\,dxdt\leq C(T_{1},A), (3.65)

where c0,2k(t)c_{0,2}^{k}(t) is the function of tt in (3.19) with x0=0x_{0}=0 and R=2R=2, and

tvkχ1+tbkχ1C(T1,A),\left\|\partial_{t}v^{k}\right\|_{\chi_{1}}+\left\|\partial_{t}b^{k}\right\|_{\chi_{1}}\leq C(T_{1},A), (3.66)

where χ1\chi_{1} is the space dual to L3(0,T1;W1,30(B1))L^{3}(0,T_{1};W^{1,3}_{0}(B_{1})). Hence there exists a sequence (v1,k,b1,k)(v^{1,k},b^{1,k}) (where the corresponding ϵ\epsilon are denoted by ϵ1,k\epsilon_{1,k}) that converges to a solution (v1,b1)(v_{1},b_{1}) of (MHD) on B1×(0,T1)B_{1}\times(0,T_{1}) in the following sense:

(v1,k,b1,k)(v1,b1) in L(0,T1;(L2×L2)(B1)),(v1,k,b1,k)(v1,b1) in L2(0,T1;(H1×H1)(B1)),(v1,k,b1,k)(v1,b1) in L3(0,T1;(L3×L3)(B1)),(𝒥ϵ1,kv1,k,𝒥ϵ1,kb1,k)(v1,b1) in L3(0,T1;(L3×L3)(B1)).\begin{split}(v^{1,k},b^{1,k})\overset{*}{\rightharpoonup}(v_{1},b_{1})&\ \text{ in }L^{\infty}(0,T_{1};(L^{2}\times L^{2})(B_{1})),\\ (v^{1,k},b^{1,k})\rightharpoonup(v_{1},b_{1})&\ \text{ in }L^{2}(0,T_{1};(H^{1}\times H^{1})(B_{1})),\\ (v^{1,k},b^{1,k})\rightarrow(v_{1},b_{1})&\ \text{ in }L^{3}(0,T_{1};(L^{3}\times L^{3})(B_{1})),\\ (\mathcal{J}_{\epsilon_{1,k}}v^{1,k},\mathcal{J}_{\epsilon_{1,k}}b^{1,k})\rightarrow(v_{1},b_{1})&\ \text{ in }L^{3}(0,T_{1};(L^{3}\times L^{3})(B_{1})).\end{split}

By Lemma 3.17 all v1,kv^{1,k} are also uniformly bounded on Bn×[0,Tn]B_{n}\times[0,T_{n}] for nn\in\mathbb{N}, n2n\geq 2 and, recursively, we can extract subsequences {(vn,k,bn,k)}k\{(v^{n,k},b^{n,k})\}_{k\in\mathbb{N}} from {(vn1,k,bn1,k)}k\{(v^{n-1,k},b^{n-1,k})\}_{k\in\mathbb{N}} which converge to a solution (vn,bn)(v_{n},b_{n}) of (MHD) on Bn×(0,Tn)B_{n}\times(0,T_{n}) as kk\to\infty in the following sense:

(vn,k,bn,k)(vn,bn) in L(0,Tn;(L2×L2)(Bn)),(vn,k,bn,k)(vn,bn) in L2(0,Tn;(H1×H1)(Bn)),(vn,k,bn,k)(vn,bn) in L3(0,Tn;(L3×L3)(Bn)),(𝒥ϵn,kvn,k,𝒥ϵn,kbn,k)(vn,bn) in L3(0,Tn;(L3×L3)(Bn)).\begin{split}(v^{n,k},b^{n,k})\overset{*}{\rightharpoonup}(v_{n},b_{n})&\ \text{ in }L^{\infty}(0,T_{n};(L^{2}\times L^{2})(B_{n})),\\ (v^{n,k},b^{n,k})\rightharpoonup(v_{n},b_{n})&\ \text{ in }L^{2}(0,T_{n};(H^{1}\times H^{1})(B_{n})),\\ (v^{n,k},b^{n,k})\rightarrow(v_{n},b_{n})&\ \text{ in }L^{3}(0,T_{n};(L^{3}\times L^{3})(B_{n})),\\ (\mathcal{J}_{\epsilon_{n,k}}v^{n,k},\mathcal{J}_{\epsilon_{n,k}}b^{n,k})\rightarrow(v_{n},b_{n})&\ \text{ in }L^{3}(0,T_{n};(L^{3}\times L^{3})(B_{n})).\end{split}

Let (v~n,b~n)(\tilde{v}_{n},\tilde{b}_{n}) be the extension by 0 of (vn,bn)(v_{n},b_{n}) to 3×(0,)\mathbb{R}^{3}\times(0,\infty). Note that, at each step, (v~n,b~n)(\tilde{v}_{n},\tilde{b}_{n}) agrees with (v~n1,b~n1)(\tilde{v}_{n-1},\tilde{b}_{n-1}) on Bn1×(0,Tn1)B_{n-1}\times(0,T_{n-1}). Let (v,b)=limn(v~n,b~n)(v,b)=\lim_{n\to\infty}(\tilde{v}_{n},\tilde{b}_{n}). Then (v,b)=(vn,bn)(v,b)=(v_{n},b_{n}) on Bn×(0,Tn)B_{n}\times(0,T_{n}) for every nn\in\mathbb{N}.

Let (vk,bk)=(vk,k,bk,k)(v^{k},b^{k})=(v^{k,k},b^{k,k}) on Bk×(0,Tk)B_{k}\times(0,T_{k}) and equal 0 elsewhere. Let ϵk\epsilon_{k} denote the corresponding regularization parameter. Then, for every fixed nn and as kk\to\infty,

(vk,bk)(v,b) in L(0,Tn;(L2×L2)(Bn)),(vk,bk)(v,b) in L2(0,Tn;(H1×H1)(Bn)),(vk,bk)(v,b) in L3(0,Tn;(L3×L3)(Bn)),(𝒥ϵkvk,𝒥ϵkbk)(v,b) in L3(0,Tn;(L3×L3)(Bn)).\begin{split}(v^{k},b^{k})\overset{*}{\rightharpoonup}(v,b)&\ \text{ in }L^{\infty}(0,T_{n};(L^{2}\times L^{2})(B_{n})),\\ (v^{k},b^{k})\rightharpoonup(v,b)&\ \text{ in }L^{2}(0,T_{n};(H^{1}\times H^{1})(B_{n})),\\ (v^{k},b^{k})\rightarrow(v,b)&\ \text{ in }L^{3}(0,T_{n};(L^{3}\times L^{3})(B_{n})),\\ (\mathcal{J}_{\epsilon_{k}}v^{k},\mathcal{J}_{\epsilon_{k}}b^{k})\rightarrow(v,b)&\ \text{ in }L^{3}(0,T_{n};(L^{3}\times L^{3})(B_{n})).\end{split} (3.67)

Based on the uniform bounds for the approximates, we have that (v,b)(v,b) satisfies (1.30).

To resolve the pressure, let

πk(x,t)=13[𝒥ϵk(vk)vk(x,t)Φϵk(x)𝒥ϵk(bk)bk(x,t)Φϵk(x)]+p.v.B2Kij(xy)[𝒥ϵk(vki)vkj(y,t)𝒥ϵk(bki)bkj(y,t)]Φϵk(y)dy+p.v.B2c(Kij(xy)Kij(y))[𝒥ϵk(vki)vkj(y,t)𝒥ϵk(bki)bkj(y,t)]Φϵk(y)dy,\begin{split}\pi^{k}(x,t)&=-\frac{1}{3}\left[\mathcal{J}_{\epsilon_{k}}(v^{k})\cdot v^{k}(x,t)\Phi_{\epsilon_{k}}(x)-\mathcal{J}_{\epsilon_{k}}(b^{k})\cdot b^{k}(x,t)\Phi_{\epsilon_{k}}(x)\right]\\ &\quad+\text{p.v.}\int_{B_{2}}K_{ij}(x-y)\left[\mathcal{J}_{\epsilon_{k}}(v^{k}_{i})v^{k}_{j}(y,t)-\mathcal{J}_{\epsilon_{k}}(b^{k}_{i})b^{k}_{j}(y,t)\right]\Phi_{\epsilon_{k}}(y)\,dy\\ &\quad+\text{p.v.}\int_{B_{2}^{c}}(K_{ij}(x-y)-K_{ij}(-y))\left[\mathcal{J}_{\epsilon_{k}}(v^{k}_{i})v^{k}_{j}(y,t)-\mathcal{J}_{\epsilon_{k}}(b^{k}_{i})b^{k}_{j}(y,t)\right]\Phi_{\epsilon_{k}}(y)\,dy,\end{split}

which, together with (vk,bk)=(vϵk,bϵk)(v^{k},b^{k})=(v^{\epsilon_{k}},b^{\epsilon_{k}}), solves (3.31) with ϵ=ϵk\epsilon=\epsilon_{k} in the distributional sense.

From the convergence properties of (vk,bk)(v^{k},b^{k}), it follows that πkπ\pi^{k}\to\pi in L3/2(0,Tn;L3/2(Bn))L^{3/2}(0,T_{n};L^{3/2}(B_{n})) for all nn where π(x,t)=limnπ¯n(x,t)\pi(x,t)=\lim_{n\to\infty}\bar{\pi}^{n}(x,t) in which π¯n(x,t)\bar{\pi}^{n}(x,t) is defined for |x|<2n|x|<2^{n} by

π¯n(x,t)=13(|v(x,t)|2|b(x,t)|2)+p.v.B2Kij(xy)(vivjbibj)(y,t)dy+π¯n3+π¯n4,\bar{\pi}^{n}(x,t)=-\frac{1}{3}\left(|v(x,t)|^{2}-|b(x,t)|^{2}\right)+\text{p.v.}\int_{B_{2}}K_{ij}(x-y)\left(v_{i}v_{j}-b_{i}b_{j}\right)(y,t)\,dy+\bar{\pi}^{n}_{3}+\bar{\pi}^{n}_{4},

with

π¯n3(x,t)=p.v.B2n+1B2(Kij(xy)Kij(y))(vivjbibj)(y,t)dy,π¯n4(x,t)=B2n+1c(Kij(xy)Kij(y))(vivjbibj)(y,t)dy.\begin{split}\bar{\pi}^{n}_{3}(x,t)&=\text{p.v.}\int_{B_{2^{n+1}}\setminus B_{2}}\left(K_{ij}(x-y)-K_{ij}(-y)\right)\left(v_{i}v_{j}-b_{i}b_{j}\right)(y,t)\,dy,\\ \bar{\pi}^{n}_{4}(x,t)&=\int_{B_{2^{n+1}}^{c}}\left(K_{ij}(x-y)-K_{ij}(-y)\right)\left(v_{i}v_{j}-b_{i}b_{j}\right)(y,t)\,dy.\end{split}

We have π¯n3,π¯n4L3/2((0,T)×B2n)\bar{\pi}^{n}_{3},\bar{\pi}^{n}_{4}\in L^{3/2}((0,T)\times B_{2^{n}}) and

π¯n3+π¯n4=π¯n+13+π¯n+14 in L3/2((0,T)×B2n).\bar{\pi}^{n}_{3}+\bar{\pi}^{n}_{4}=\bar{\pi}^{n+1}_{3}+\bar{\pi}^{n+1}_{4}\ \text{ in }L^{3/2}((0,T)\times B_{2^{n}}).

Thus π¯n\bar{\pi}^{n} is independent of nn for n>log2|x|n>\log_{2}|x|.

We now establish the above local pressure expression for all scales. Note that the formula is valid for πk\pi^{k} at all scales, that is, for any T>0T>0, fixed R>0R>0 and x03x_{0}\in\mathbb{R}^{3}, we have the following equality in L3/2(B2R(x0)×(0,T))L^{3/2}(B_{2R}(x_{0})\times(0,T)),

π^x0,Rk(x,t):=πk(x,t)cx0,Rk(t)=Δ1divdiv[((𝒥kvkvk𝒥kbkbk)Φk)χ4R(xx0)]3(K(xy)K(x0y))((𝒥kvkvk𝒥kbkbk)Φk)(y,t)(1χ4R(yx0))dy,\begin{split}\hat{\pi}_{x_{0},R}^{k}(x,t)&:=\pi^{k}(x,t)-c_{x_{0},R}^{k}(t)\\ &=-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[\left((\mathcal{J}_{k}v^{k}\otimes v^{k}-\mathcal{J}_{k}b^{k}\otimes b^{k})\Phi_{k}\right)\chi_{4R}(x-x_{0})\right]\\ &\quad-\int_{\mathbb{R}^{3}}\left(K(x-y)-K(x_{0}-y)\right)\left((\mathcal{J}_{k}v^{k}\otimes v^{k}-\mathcal{J}_{k}b^{k}\otimes b^{k})\Phi_{k}\right)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy,\end{split}

where 𝒥k=𝒥ϵk\mathcal{J}_{k}=\mathcal{J}_{\epsilon_{k}} and Φk=Φϵk\Phi_{k}=\Phi_{\epsilon_{k}}. Similarly, let

π^x0,R(x,t)=Δ1divdiv[(vvbb)χ4R(xx0)]3(K(xy)K(x0y))(vvbb)(y,t)(1χ4R(yx0))dy.\begin{split}\hat{\pi}_{x_{0},R}(x,t)&=-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[(v\otimes v-b\otimes b)\chi_{4R}(x-x_{0})\right]\\ &\quad-\int_{\mathbb{R}^{3}}\left(K(x-y)-K(x_{0}-y)\right)(v\otimes v-b\otimes b)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy.\end{split}

Fix T>0T>0, x03x_{0}\in\mathbb{R}^{3} and R>0R>0. Choose nn large enough that B8R(x0)×(0,T)Qn=Bn×(0,Tn)B_{8R}(x_{0})\times(0,T)\subset Q_{n}=B_{n}\times(0,T_{n}). We claim that π^x0,Rk(x,t)\hat{\pi}_{x_{0},R}^{k}(x,t) converges to p^x0,R(x,t)\hat{p}_{x_{0},R}(x,t) in L3/2(B2R(x0)×(0,T))L^{3/2}(B_{2R}(x_{0})\times(0,T)). If this is the case, by taking the limit of the weak form of (3.31), we can show that (v,b,π^x0,R)(v,b,\hat{\pi}_{x_{0},R}) also satisfies (MHD) in B2R(x0)×(0,T)B_{2R}(x_{0})\times(0,T). Hence ππ^x0,R=0\nabla\pi-\nabla\hat{\pi}_{x_{0},R}=0, and we may define

cx0,R(t)=π(x,t)π^x0,R(x,t)c_{x_{0},R}(t)=\pi(x,t)-\hat{\pi}_{x_{0},R}(x,t)

which is hence a function of tt in L3/2(0,T)L^{3/2}(0,T) that is independent of xx. This gives the desired local pressure expansion in B2R(x0)×(0,T)B_{2R}(x_{0})\times(0,T).

To verify the claim we work term by term. Note that the estimate in [23, (3.26)] shows that

vivj(𝒥kvki)vkjΦkL3/2(BM×[0,Tn]),bibj(𝒥kbki)bkjΦkL3/2(BM×[0,Tn])0,\left\|v_{i}v_{j}-(\mathcal{J}_{k}v^{k}_{i})v^{k}_{j}\Phi_{k}\right\|_{L^{3/2}(B_{M}\times[0,T_{n}])},\qquad\left\|b_{i}b_{j}-(\mathcal{J}_{k}b^{k}_{i})b^{k}_{j}\Phi_{k}\right\|_{L^{3/2}(B_{M}\times[0,T_{n}])}\to 0,

as kk\to\infty for every M>0M>0 and nn\in\mathbb{N}. This implies

Δ1divdiv[((𝒥kvkvk𝒥kbkbk)Φk)χ4R(xx0)]Δ1divdiv[(vvbb)χ4R(xx0)]\begin{split}-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits&\left[\left((\mathcal{J}_{k}v^{k}\otimes v^{k}-\mathcal{J}_{k}b^{k}\otimes b^{k})\Phi_{k}\right)\chi_{4R}(x-x_{0})\right]\\ &\qquad\rightarrow-\Delta^{-1}\mathop{\rm div}\nolimits\mathop{\rm div}\nolimits\left[(v\otimes v-b\otimes b)\chi_{4R}(x-x_{0})\right]\end{split}

in L3/2(B2R(x0)×(0,Tn))L^{3/2}(B_{2R}(x_{0})\times(0,T_{n})), and

|x|<M(K(xy)K(x0y))((𝒥kvkvk𝒥kbkbk)Φk)(y,t)(1χ4R(yx0))dy|x|<M(K(xy)K(x0y))(vvbb)(y,t)(1χ4R(yx0))dy\begin{split}-\int_{|x|<M}&\left(K(x-y)-K(x_{0}-y)\right)\left((\mathcal{J}_{k}v^{k}\otimes v^{k}-\mathcal{J}_{k}b^{k}\otimes b^{k})\Phi_{k}\right)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy\\ &\rightarrow-\int_{|x|<M}\left(K(x-y)-K(x_{0}-y)\right)(v\otimes v-b\otimes b)(y,t)\left(1-\chi_{4R}(y-x_{0})\right)dy\end{split}

in L3/2(B2R(x0)×(0,Tn))L^{3/2}(B_{2R}(x_{0})\times(0,T_{n})) for every M>8RM>8R. For the far-field part, still assuming M>8RM>8R, we have

|x|M(K(xy)K(x0y))((𝒥kvkvk𝒥kbkbk)Φk(vvbb))(y,t)dyL3/2(B2R(x0)×(0,Tn))C(R,n,(v0,b0)L2uloc×L2uloc)M1C(R,n,(v0,b0)E2q×E2q)M1,\begin{split}&\left\|\int_{|x|\geq M}\left(K(x-y)-K(x_{0}-y)\right)\left((\mathcal{J}_{k}v^{k}\otimes v^{k}-\mathcal{J}_{k}b^{k}\otimes b^{k})\Phi_{k}-(v\otimes v-b\otimes b)\right)(y,t)dy\right\|_{L^{3/2}(B_{2R}(x_{0})\times(0,T_{n}))}\\ &\qquad\leq C(R,n,\left\|(v_{0},b_{0})\right\|_{L^{2}_{\mathrm{uloc}}\times L^{2}_{\mathrm{uloc}}})M^{-1}\leq C(R,n,\left\|(v_{0},b_{0})\right\|_{E^{2}_{q}\times E^{2}_{q}})M^{-1},\end{split}

where we’ve used the embedding E2qL2ulocE^{2}_{q}\subset L^{2}_{\mathrm{uloc}}. This can be made arbitrarily small by taking MM large and noting RR and nn are fixed. Consequently, and since the other parts of the pressure converge, we conclude that

π^x0,Rk(x,t)π^x0,R(x,t) in L3/2(B2R(x0)×(0,Tn)),\hat{\pi}_{x_{0},R}^{k}(x,t)\rightarrow\hat{\pi}_{x_{0},R}(x,t)\ \text{ in }L^{3/2}(B_{2R}(x_{0})\times(0,T_{n})), (3.68)

which leads to the desired local pressure expansion. Since nn was arbitrary, this gives the pressure formula for arbitrarily large times.

At this point we have established items 1.-3. from the definition of local energy solutions. We now check remaining items.

Fix T0T_{0} and choose nn so that TnT0T_{n}\geq T_{0}. Then (3.67) holds for all nn with TnT_{n} replaced by T0T_{0}. Furthermore, the estimates (3.63)–(3.66) and (3.68) are valid in Bn×[0,T0]B_{n}\times[0,T_{0}] up to a re-definition of AA. Moreover, we have

tvχn+tbχnC(n,T0,A),\left\|\partial_{t}v\right\|_{\chi_{n}}+\left\|\partial_{t}b\right\|_{\chi_{n}}\leq C(n,T_{0},A), (3.69)

and

esssup0tT0B1(k)(|v|2+|b|2)dx+0T0B1(k)(|v|2+|b|2)dxdtq/2(k3)2A.\bigg{\|}\mathop{\rm ess\,sup}_{0\leq t\leq T_{0}}\int_{B_{1}(k)}\left(|v|^{2}+|b|^{2}\right)dx+\int_{0}^{T_{0}}\int_{B_{1}(k)}\left(|\nabla v|^{2}+|\nabla b|^{2}\right)dx\,dt\bigg{\|}_{\ell^{q/2}(k\in\mathbb{Z}^{3})}\leq 2A. (3.70)

It follows from (3.69) and (3.70) that for every nn,

tBnv(x,t)w(x)dxtBnb(x,t)w(x)dxt\mapsto\int_{B_{n}}v(x,t)\cdot w(x)\,dx\qquad t\mapsto\int_{B_{n}}b(x,t)\cdot w(x)\,dx (3.71)

are continuous in t[0,T0]t\in[0,T_{0}] for every wL2(B2)w\in L^{2}(B_{2}). Since T0T_{0} was arbitrary, we can extend this to all times. The local energy inequality follows from the local energy equality for (vk,bk)(v^{k},b^{k}) and πk\pi^{k}, and (3.67), (3.68) in Bn×[0,T0]B_{n}\times[0,T_{0}], (3.69), and π^n(x,t)=π(x,t)cn(t)\hat{\pi}_{n}(x,t)=\pi(x,t)-c_{n}(t) for some cnL3/2(0,T0)c_{n}\in L^{3/2}(0,T_{0}). Convergence to initial data in L2locL^{2}_{\mathrm{loc}} follows from (3.71) and the local energy inequality. This confirms that items 4.-6. from the definition of local energy solutions are satisfied and finishes the proof of Theorem 1.10 for 1q<21\leq q<2. ∎

Acknowledgments

I warmly thank Zachary Bradshaw and Tai-Peng Tsai for helpful comments. The research was partially support by the AMS-Simons Travel Grant and the Simons Foundation Math + X Investigator Award #376319 (Michael I. Weinstein). The author gratefully acknowledges the unwavering financial and emotional support of his wife, Anyi Bao.


References

  • [1] Z. Bradshaw, C.-C. Lai, and T.-P. Tsai. Mild solutions and spacetime integral bounds for Stokes and Navier-Stokes flows in Wiener amalgam spaces. Math. Ann., 388(3):3053–3126, 2024.
  • [2] Z. Bradshaw and T.-P. Tsai. Forward discretely self-similar solutions of the Navier-Stokes equations II. Ann. Henri Poincaré, 18(3):1095–1119, 2017.
  • [3] Z. Bradshaw and T.-P. Tsai. Global existence, regularity, and uniqueness of infinite energy solutions to the Navier-Stokes equations. Comm. Partial Differential Equations, 45(9):1168–1201, 2020.
  • [4] Z. Bradshaw and T.-P. Tsai. Local energy solutions to the Navier-Stokes equations in Wiener amalgam spaces. SIAM J. Math. Anal., 53(2):1993–2026, 2021.
  • [5] Z. Bradshaw and T.-P. Tsai. On the local pressure expansion for the Navier–Stokes equations. J. Math. Fluid Mech., 24:1–32, 2022.
  • [6] L. Caffarelli, R. Kohn, and L. Nirenberg. Partial regularity of suitable weak solutions of the Navier-Stokes equations. Comm. Pure Appl. Math., 35(6):771–831, 1982.
  • [7] C. Cao and J. Wu. Two regularity criteria for the 3D MHD equations. J. Differential Equations, 248(9):2263–2274, 2010.
  • [8] D. Chamorro, F. Cortez, J. He, and O. Jarrín. On the local regularity theory for the magnetohydrodynamic equations. Doc. Math., 26:125–148, 2021.
  • [9] Q. Chen, C. Miao, and Z. Zhang. On the regularity criterion of weak solution for the 3D viscous magneto-hydrodynamics equations. Comm. Math. Phys., 284:919–930, 2008.
  • [10] Y. Chen and P. Zhang. The global existence of small solutions to the incompressible viscoelastic fluid system in 2 and 3 space dimensions. Comm. Partial Differential Equations, 31(10-12):1793–1810, 2006.
  • [11] G. Duvaut and J.-L. Lions. Inéquations en thermoélasticité et magnétohydrodynamique. Arch. Rational Mech. Anal., 46:241–279, 1972.
  • [12] P. G. Fernández-Dalgo and O. Jarrín. Discretely self-similar solutions for 3D MHD equations and global weak solutions in weighted L2L^{2} spaces. J. Math. Fluid Mech., 23(1):Paper No. 22, 30, 2021.
  • [13] P. G. Fernández-Dalgo and O. Jarrín. Weak-strong uniqueness in weighted L2 spaces and weak suitable solutions in local Morrey spaces for the MHD equations. J. Differential Equations, 271:864–915, 2021.
  • [14] Y. Giga, K. Inui, and S. Matsui. On the cauchy problem for the Navier-Stokes equations with nondecaying initial data. Quad. Mat., 4:28–68, 1999.
  • [15] C. He and Z. Xin. Partial regularity of suitable weak solutions to the incompressible magnetohydrodynamic equations. J. Funct. Anal., 227(1):113–152, 2005.
  • [16] C. He and Z. Xin. On the self-similar solutions of the magneto-hydro-dynamic equations. Acta Math. Sci. Ser. B (Engl. Ed.), 29(3):583–598, 2009.
  • [17] R. Hynd. Partial regularity of weak solutions of the viscoelastic Navier-Stokes equations with damping. SIAM J. Math. Anal., 45(2):495–517, 2013.
  • [18] K. Kang, B. Lai, C.-C. Lai, and T.-P. Tsai. The Green tensor of the nonstationary Stokes system in the half space. Comm. Math. Phys., 399(2):1291–1372, 2023.
  • [19] K. Kang and J. Lee. Interior regularity criteria for suitable weak solutions of the magnetohydrodynamic equations. J. Differential Equations, 247(8):2310–2330, 2009.
  • [20] K. Kang, H. Miura, and T.-P. Tsai. Short time regularity of Navier–Stokes flows with locally L3 initial data and applications. Int. Math. Res. Not. IMRN, 2021(11):8763–8805, 2021.
  • [21] N. Kikuchi and G. Seregin. Weak solutions to the Cauchy problem for the Navier-Stokes equations satisfying the local energy inequality. In Nonlinear equations and spectral theory, volume 220 of Amer. Math. Soc. Transl. Ser. 2, pages 141–164. Amer. Math. Soc., Providence, RI, 2007.
  • [22] J.-M. Kim. On regularity criteria of weak solutions to the 3D viscoelastic Navier-Stokes equations with damping. Appl. Math. Lett., 69:153–160, 2017.
  • [23] H. Kwon and T.-P. Tsai. Global Navier–Stokes flows for non-decaying initial data with slowly decaying oscillation. Comm. Math. Phys., 375(3):1665–1715, 2020.
  • [24] B. Lai, J. Lin, and C. Wang. Forward self-similar solutions to the viscoelastic Navier-Stokes equation with damping. SIAM J. Math. Anal., 49(1):501–529, 2017.
  • [25] C.-C. Lai. Forward discretely self-similar solutions of the MHD equations and the viscoelastic Navier-Stokes equations with damping. J. Math. Fluid Mech., 21(3):Paper No. 38, 28, 2019.
  • [26] Z. Lei, C. Liu, and Y. Zhou. Global solutions for incompressible viscoelastic fluids. Arch. Ration. Mech. Anal., 188(3):371–398, 2008.
  • [27] P. G. Lemarié-Rieusset. Recent developments in the Navier-Stokes problem, volume 431 of Chapman & Hall/CRC Research Notes in Mathematics. Chapman & Hall/CRC, Boca Raton, FL, 2002.
  • [28] F. Lin. A new proof of the Caffarelli-Kohn-Nirenberg theorem. Comm. Pure Appl. Math., 51(3):241–257, 1998.
  • [29] F.-H. Lin, C. Liu, and P. Zhang. On hydrodynamics of viscoelastic fluids. Comm. Pure Appl. Math., 58(11):1437–1471, 2005.
  • [30] Y. Lin, H. Zhang, and Y. Zhou. Global smooth solutions of MHD equations with large data. J. Differential Equations, 261(1):102–112, 2016.
  • [31] F. Liu, S. Xi, Z. Zeng, and S. Zhu. Global mild solutions to three-dimensional magnetohydrodynamic equations in Morrey spaces. J. Differential Equations, 314:752–807, 2022.
  • [32] M. Liu and X. Lin. Global classical solutions to the elastodynamic equations with damping. J. Inequal. Appl., 2021(1):88, 2021.
  • [33] Y. Maekawa and Y. Terasawa. The Navier-Stokes equations with initial data in uniformly local LpL^{p} spaces. Differential Integral Equations, 19(4):369–400, 2006.
  • [34] A. Mahalov, B. Nicolaenko, and T. Shilkin. L3,L_{3,\infty}-solutions to the MHD equations. J. Math. Sci, 143:2911–2923, 2007.
  • [35] C. Miao, B. Yuan, and B. Zhang. Well-posedness for the incompressible magneto-hydrodynamic system. Math. Methods Appl. Sci., 30(8):961–976, 2007.
  • [36] J. Nečas, M. Růžička, and V. Šverák. On Leray’s self-similar solutions of the Navier-Stokes equations. Acta Math., 176(2):283–294, 1996.
  • [37] C. W. Oseen. Neuere methoden und ergebnisse in der hydrodynamik, volume 1. Akademische Verlagsgesellschaft, 1927.
  • [38] M. Sermange and R. Temam. Some mathematical questions related to the MHD equations. Comm. Pure Appl. Math., 36(5):635–664, 1983.
  • [39] Z. Tan, W. Wu, and J. Zhou. Existence of mild solutions and regularity criteria of weak solutions to the viscoelastic Navier–Stokes equation with damping. Commun. Math. Sci., 18(1):205–226, 2020.
  • [40] T.-P. Tsai. Lectures on Navier-Stokes equations, volume 192 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2018.
  • [41] J. Wu. Regularity results for weak solutions of the 3D MHD equations. Discrete Contin. Dyn. Syst, 10(1/2):543–556, 2004.
  • [42] Y. Yang. Forward self-similar solutions to the mhd equations in the whole space. arXiv preprint arXiv:2404.02601, 2024.
  • [43] G. Yue, Z. Pang, and Y. Wu. On the interior regularity criteria for the viscoelastic fluid system with damping. Adv. Calc. Var., 18(2):323–338, 2025.
  • [44] Y. Zhou. Remarks on regularities for the 3D MHD equations. Discrete Contin. Dyn. Syst, 12(5):881–886, 2005.