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

Strong convergence of the vorticity and conservation
of the energy for the α\alpha-Euler equations

Stefano Abbate Gran Sasso Science Institute (GSSI), Viale Francesco Crispi 7, I-67100, L’Aquila, Italy Gianluca Crippa Departement Mathematik und Informatik, Universität Basel, Spiegelgasse 1, CH-4051, Basel, Switzerland Stefano Spirito DISIM - Dipartimento di Ingegneria e Scienze dell’Informazione e Matematica, Università degli Studi de L’Aquila, Via Vetoio, I-67100, L’Aquila, Italy
Abstract

In this paper, we study the convergence of solutions of the α\alpha-Euler equations to solutions of the Euler equations on the 22-dimensional torus. In particular, given an initial vorticity ω0\omega_{0} in LxpL^{p}_{x} for p(1,)p\in(1,\infty), we prove strong convergence in LtLxpL^{\infty}_{t}L^{p}_{x} of the vorticities qαq^{\alpha}, solutions of the α\alpha-Euler equations, towards a Lagrangian and energy-conserving solution of the Euler equations. Furthermore, if we consider solutions with bounded initial vorticity, we prove a quantitative rate of convergence of qαq^{\alpha} to ω\omega in LpL^{p}, for p(1,)p\in(1,\infty).

1 Introduction

2023 Mathematics Subject Classification. Primary: 76B03, 35Q35. Secondary: 35Q31, 76F65.   Key words and phrases. 2D Euler equations, 2D alpha-Euler equations, Lagrangian solutions, conservation of the energy.

In this paper, we consider the incompressible α\alpha-Euler equations on the two-dimensional torus, which, given α>0\alpha>0, read as

{tvα+uαvα+jvjαujα=p,on(0,T)×𝕋2vα:=uααΔuα,on(0,T)×𝕋2divuα=divvα=0,on(0,T)×𝕋2uα(0,)=u0α,on𝕋2\displaystyle\begin{cases}\partial_{t}v^{\alpha}+u^{\alpha}\cdot\nabla v^{\alpha}+\sum_{j}v^{\alpha}_{j}\nabla u^{\alpha}_{j}=-\nabla p,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ v^{\alpha}:=u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha},&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \operatorname{div}u^{\alpha}=\operatorname{div}v^{\alpha}=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ u^{\alpha}(0,\cdot)=u^{\alpha}_{0},&\quad\text{on}\quad\mathbb{T}^{2}\end{cases} (1)

and our primary objective is the rigorous study of the limit as α0\alpha\to 0 of solutions of (1). Formally, if we substitute α=0\alpha=0 in (1), we obtain that vα=uαv^{\alpha}=u^{\alpha}, and then, employing the identity

jujαujα=|uα|22\sum_{j}u^{\alpha}_{j}\nabla u^{\alpha}_{j}=\frac{\nabla|u^{\alpha}|^{2}}{2}

and defining πα:=p+|uα|2/2,\pi^{\alpha}:=p+|u^{\alpha}|^{2}/2, we obtain the two-dimensional incompressible Euler equations:

{tu+(u)u=π,on(0,T)×𝕋2divu=0,on(0,T)×𝕋2u(0,)=u0,on𝕋2.\begin{cases}\partial_{t}u+(u\cdot\nabla)u=-\nabla\pi,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \operatorname{div}u=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ u(0,\cdot)=u_{0},&\quad\text{on}\quad\mathbb{T}^{2}.\end{cases} (2)

The α\alpha-Euler equations are part of a larger class of approximation schemes called Large Eddies Simulations (LES), which have been first introduced in [smago] by Smagorinsky and later generalized by Leonard [leonard]. LES approximations are relevant for numerical simulations of fluids in turbulent regime. Indeed, due to the high number of scales needed in turbulent dynamics, Direct Numerical Simulation (DNS) of fluid equations are computationally very demanding. The idea of LES models is based on the fact that the whole range of flow scales may not be necessary in order to have an accurate approximation. Therefore, a filter cutting the small scales is applied to the velocity. Since the operation of filtering does not commute with the nonlinear convective term, one needs to estimate the commutator between the filter and the convective term and this procedure gives rise to an approximation. We refer to [bookberselli] and [gop] for more details on the derivation and the motivations of LES models. In particular, we refer to [fht, mars] for the derivation of the α\alpha-Euler equations.

Concerning the system (1), the filter is given by the so-called Helmholtz filter, which is exactly the second equation in (1), namely

uα:=(IαΔ)1vα.u^{\alpha}:=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}v^{\alpha}. (3)

The action of the filter (3) can be written in Fourier variables as

uα^(k)=vα^(k)1+α|k|2,k2.\widehat{u^{\alpha}}(k)=\frac{\widehat{v^{\alpha}}(k)}{1+\alpha|k|^{2}},\qquad\forall k\in\mathbb{Z}^{2}. (4)

The denominator of the right hand side of (4) diverges as the frequency grows, cutting the high frequencies as a consequence.

The analysis of the limit as α0\alpha\to 0 for solutions of the α\alpha-Euler equations has been widely studied in literature. In particular, we mention the important result in [lopestitizang], where the authors analyse the convergence in L2L^{2} for the velocity fields in two dimensions for smooth solutions of the (1) equations on a bounded domain with Dirichlet boundary conditions towards smooth solutions of the Euler equations. The importance relies on the fact that in the limit of α0\alpha\to 0 no boundary layers are created for the velocity. The result in [lopestitizang] has been extended in [bi] where the convergence of the velocity for the problem posed on bounded domains with Dirichlet boundary conditions has been proved for less regular solutions. We also recall the paper [bill], where the convergence as α0\alpha\to 0 is studied for initial vorticity in the space of positive Radon measures. Finally, we refer to [olishk, holm] where the relationship between the α\alpha-Euler equations and the vortex blob method (another commonly used numerical approximation of the Euler equations) has been investigated.

An important feature of the system (1) is that the vorticity structure of the 2D2D Euler equations is preserved in the approximation. Indeed, thanks to the presence of the term jvjαujα\sum_{j}v^{\alpha}_{j}\nabla u^{\alpha}_{j} in the momentum equations of (1), if we consider qα=curlvαq^{\alpha}=\operatorname{curl}v^{\alpha}, we obtain the following vorticity formulation of (1)

{tqα+uαqα=0,on(0,T)×𝕋2q(0,)=q0α,on𝕋2divuα=divvα=0,on(0,T)×𝕋2curlvα=qα,on(0,T)×𝕋2.\displaystyle\begin{cases}\partial_{t}q^{\alpha}+u^{\alpha}\cdot\nabla q^{\alpha}=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ q(0,\cdot)=q_{0}^{\alpha},&\quad\text{on}\quad\mathbb{T}^{2}\\ \operatorname{div}u^{\alpha}=\operatorname{div}v^{\alpha}=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \operatorname{curl}v^{\alpha}=q^{\alpha},&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}.\end{cases} (5)

The vorticity formulation (5) is particularly important for the purpose of this paper, since we are primarily interested in the analysis of the convergence of the vorticity qαq^{\alpha} towards the vorticity of the 2D2D Euler equations in strong norms. This is reminiscent of the analogous results for the vanishing viscosity limit. In two-dimensional turbulence, the transport-structure for the vorticity and the appearance of an inverse cascade (from small to large scales) entail that vanishing viscosity solutions enjoy better properties compared to general weak solutions. In this paper, we show that the same principle holds for solutions that are the limit of the α\alpha-Euler equations.

In our first result, we show the strong convergence in LpL^{p} of the vorticity to a solution of the Euler equations which conserves the energy and is Lagrangian, namely the vorticity is transported by the flow of the associated velocity, see Definition 2.3. We refer to Section 2 for the relevant notations and definitions. In particular, kk is the Biot-Savart kernel on the torus (cf. (18)).

Theorem 1.1.

Let T>0T>0 arbitrary and finite, p(1,)p\in(1,\infty) and ω0Lp(𝕋2)\omega_{0}\in L^{p}(\mathbb{T}^{2}) with 𝕋2ω0=0\int_{\mathbb{T}^{2}}\omega_{0}=0. Let {q0α}\{q_{0}^{\alpha}\} be a sequence of functions uniformly bounded with respect to α\alpha in Lp(𝕋2)L^{p}(\mathbb{T}^{2}) with 𝕋2q0α=0\int_{\mathbb{T}^{2}}q_{0}^{\alpha}=0 such that

q0αω0strongly inLp(𝕋2).q_{0}^{\alpha}\rightarrow\omega_{0}\quad\text{strongly in}\enskip L^{p}(\mathbb{T}^{2}).

Let (uα,qα)(u^{\alpha},q^{\alpha}) be the solution of the α\alpha-Euler equations with initial datum q0αq_{0}^{\alpha}. Then, up to subsequences, there holds,

uαustrongly in𝒞([0,T];L2(𝕋2))u^{\alpha}\rightarrow u\quad\text{strongly in}\enskip\mathcal{C}([0,T];L^{2}(\mathbb{T}^{2}))

and

qαωstrongly in𝒞([0,T];Lp(𝕋2)),q^{\alpha}\rightarrow\omega\quad\text{strongly in}\enskip\mathcal{C}([0,T];L^{p}(\mathbb{T}^{2})),

and (u,ω)(u,\omega) is a Lagrangian solution of the Euler equations. Moreover, for any δ>0\delta>0, there exists K(δ,ω0)K(\delta,\omega_{0}) such that for α\alpha small enough it holds

supt[0,T]qα(t)ω(t)Lpδ+K(δ,ω0)|log(uαuLt1Lx1)|+q0αω0Lp.\sup_{t\in[0,T]}\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq\delta+\frac{K(\delta,\omega_{0})}{\left|\log\left(\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}\right)\right|}+\|q^{\alpha}_{0}-\omega_{0}\|_{L^{p}}. (6)

Finally, the solution uu conserves the kinetic energy, namely

u(t)L2=u0L2,t(0,T),whereu0=kω0.\|u(t)\|_{L^{2}}=\|u_{0}\|_{L^{2}},\quad\forall t\in(0,T),\quad\text{where}\quad u_{0}=k\ast\omega_{0}. (7)

The novelty of our approach consists in the application of techniques related to the Lagrangian perspective, introduced in the non-smooth settings in [odeest] and employed for a vanishing viscosity scheme in [ciampa]. We first prove the convergence of the velocity adapting the proof of [bi] to this setting. Then, by using Lagrangian techniques, we show strong convergence of the vorticity providing a certain quantification of the convergence. Nonetheless, the rate of convergence is not fully quantitative since it depends (logarithmically) on the rate of convergence of the velocities.

Weak solutions of the Euler equations with vorticity in LpL^{p} with p3/2p\geq 3/2 conserve the kinetic energy, but this is not known for p<3/2p<3/2 (cf. [cheski]). In Theorem 1.1, we also prove that limit solutions are energy conserving, for every p(1,]p\in(1,\infty], as previously done for the vanishing viscosity in [cheski] and for the vortex blob in [ciampa2].

The second main result of this paper concerns the study of the rate of convergence in the case of solutions belonging to the Yudovich class. In particular, the next theorem shows that if we consider bounded initial vorticity, we can obtain a rate of convergence independent on uαuLt1Lx1\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}.

Theorem 1.2.

Let ω0L(𝕋2)\omega_{0}\in L^{\infty}(\mathbb{T}^{2}) with 𝕋2ω0=0\int_{\mathbb{T}^{2}}\omega_{0}=0. Let q0αq^{\alpha}_{0} uniformly bounded with respect to α\alpha in L(𝕋2)L^{\infty}(\mathbb{T}^{2}) with 𝕋2q0α=0\int_{\mathbb{T}^{2}}q_{0}^{\alpha}=0 such that q0αω0q^{\alpha}_{0}\rightarrow\omega_{0} in Lp(𝕋2)L^{p}(\mathbb{T}^{2}), for every p<p<\infty. Let u0α:=(IαΔ)1kq0αu^{\alpha}_{0}:=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}k\ast q^{\alpha}_{0} and u0:=kω0u_{0}:=k\ast\omega_{0}, then

γ0α:=u0αu0L2+αΔu0αL2α00and letα¯>0be s.t.γ0α12,αα¯.\gamma_{0}^{\alpha}:=\|u^{\alpha}_{0}-u_{0}\|_{L^{2}}+\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}}\xrightarrow{\alpha\to 0}0\quad\text{and let}\quad\overline{\alpha}>0\quad\text{be s.t.}\quad\gamma_{0}^{\alpha}\leq\frac{1}{2},\quad\forall\alpha\leq\overline{\alpha}. (8)

Let (uα,qα)(u^{\alpha},q^{\alpha}) be the solution to the α\alpha-Euler equations with initial datum q0αq^{\alpha}_{0} and let (u,ω)(u,\omega) be its limit, which is the unique Yudovich solution to the Euler equations. Then, there exist two constants C1C_{1} and C2C_{2} (depending on M:=ω0LM:=\|\omega_{0}\|_{L^{\infty}}) such that, if α(0,α¯]\alpha\in(0,\overline{\alpha}] and T>0T>0 satisfy

αexp{2(22exp(C2T))}γ0α(C1T)2,\alpha\leq\frac{\exp\{2(2-2\exp(C_{2}T))\}-\gamma_{0}^{\alpha}}{(C_{1}T)^{2}}, (9)

it holds

u(t)uα(t)L2exp{22exp(C2t)}(C1αT+γ0α)exp(C2t)+Cα:=K(α,t),tT.\|u(t)-u^{\alpha}(t)\|_{L^{2}}\leq\exp\{2-2\exp(-C_{2}t)\}(C_{1}\sqrt{\alpha}T+\gamma^{\alpha}_{0})^{\exp(-C_{2}t)}+C\sqrt{\alpha}:=K(\alpha,t),\quad\forall t\leq T.

Moreover, there exists a value α0=α0(T,M,ω0)\alpha_{0}=\alpha_{0}(T,M,\omega_{0}) and a continuous function ψω0,p,M:++\psi_{\omega_{0},p,M}:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} vanishing at zero, such that for every αα0\alpha\leq\alpha_{0} there holds

supt[0,T]qα(t)ω(t)LpCM11pmax{ψω0,p,M(K(α,t)),K(α,t)exp(CT)2p},\sup_{t\in[0,T]}\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq CM^{1-\frac{1}{p}}\max\left\{\psi_{\omega_{0},p,M}(K(\alpha,t)),K(\alpha,t)^{\frac{\exp{(-CT)}}{2p}}\right\}, (10)

where CC depends on M=ω0LM=\|\omega_{0}\|_{L^{\infty}}.

One can check that thanks to (8), it is always possible to find α\alpha small enough such that (9) holds for a fixed time TT and vice versa a positive time TT such that (9) holds for α(0,α¯]\alpha\in(0,\overline{\alpha}]. Finally, assuming additional regularity on the initial datum, it is possible provide an explicit expression for the function ψω0,p,M\psi_{\omega_{0},p,M}.

Corollary 1.3.

Under the same assumptions of Theorem 1.2, let s(0,1)s\in(0,1), if ω0\omega_{0} belongs to the Besov space Bp,s(𝕋2)B^{s}_{p,\infty}(\mathbb{T}^{2}), then the function ψω0,p,M()\psi_{\omega_{0},p,M}(\cdot) is controlled, yielding

supt(0,T)qα(t)ω(t)LpCM11pmax{K(α,t)s,K(α,t)exp(CT)2p},\sup_{t\in(0,T)}\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq CM^{1-\frac{1}{p}}\max\left\{K(\alpha,t)^{s},K(\alpha,t)^{\frac{\exp{(-CT)}}{2p}}\right\}, (11)

where CC depends on MM.


Theorem 1.2 is the analogous of the results obtained in [const, ciampa, seiswied] for the vanishing viscosity limit and Corollary 1.3 corresponds to [const], Corollary 2, for the vanishing viscosity limit. In particular, the interest in considering initial vorticity in Bp,sB^{s}_{p,\infty} relies on the fact that some classes of vortex patch are in those spaces. Indeed, in [cowu2] it has been proved that if χΩ\chi_{\Omega} is the characteristic function of Ω2\Omega\subset\mathbb{R}^{2} whose boundary Ω\partial\Omega has box-counting dimension dimF(Ω)<2\dim_{F}(\Omega)<2, then

χΩBp,2dimF(Ω)p(𝕋2),p[1,).\chi_{\Omega}\in B^{\frac{2-\dim_{F}(\Omega)}{p}}_{p,\infty}(\mathbb{T}^{2}),\quad\forall p\in[1,\infty).

Finally, we mention that in the case of the α\alpha-Euler equations, when considering a vortex patch of boundary 𝒞1,γ\mathcal{C}^{1,\gamma}, in [lt] the authors obtained a rate of convergence of order α34\alpha^{\frac{3}{4}} according to our notation. The proof of [lt] is built upon the observations that the C1,γC^{1,\gamma} regularity of the boundary of a vortex patch is preserved in time and the vortex patches under this assumption belong to the space L(0,T;B2,12(𝕋2))L^{\infty}(0,T;{B}^{\frac{1}{2}}_{2,\infty}(\mathbb{T}^{2})).

Organization of the paper

In Section 2, we introduce the notations used throughout the paper and we recall some standard results on the Euler equations and the known results on the well-posedness of the α\alpha-Euler equations. Section 3 is devoted to the proof of Theorem 1.1. In particular, the proof is split between Proposition 3.2 where we prove strong convergence of the velocity and Proposition 3.7 where we prove strong convergence of the vorticity. Finally, in Section 4 we prove Theorem 1.2 and Corollary 1.3 where we estimate the rate of convergence under the additional hypothesis of bounded initial datum.

Acknowledgments

This research has been partially supported by the SNF Project 212573 FLUTURA – Fluids, Turbulence, Advection. The first author acknowledges the hospitality of the University of Basel, where a large part of this work was carried out. The work of the third author is partially supported by INdAM-GNAMPA and by the project PRIN 2020 ”Nonlinear evolution PDEs, fluid dynamics and transport equations: theoretical foundations and applications”.

2 Notations and preliminary results

2.1   Notations

Throughout this work, we always consider as a domain the two-dimensional flat torus 𝕋2=2/2\mathbb{T}^{2}=\mathbb{R}^{2}/\mathbb{Z}^{2} and we denote the Lebesgue measure on it by 2\mathcal{L}^{2}. The distance on the torus is defined as

𝐝(x,y)=min{|xyk|:k2|k|<2}\operatorname{\mathbf{d}}(x,y)=\min\{|x-y-k|:k\in\mathbb{Z}^{2}\enskip|\enskip|k|<2\}

and it is immediate to check that 𝐝(x,y)|xy|\operatorname{\mathbf{d}}(x,y)\leq|x-y|.
We use the standard definitions of functional spaces Lp=Lp(𝕋2)L^{p}=L^{p}(\mathbb{T}^{2}), Hs=Hs(𝕋2)H^{s}=H^{s}(\mathbb{T}^{2}) and Ws,p=Ws,p(𝕋2)W^{s,p}=W^{s,p}(\mathbb{T}^{2}), in which we avoid writing explicitly the dependence on 𝕋2\mathbb{T}^{2} for the norms. To simplify the notation, we use LtpXxL^{p}_{t}X_{x} for the Bochner spaces Lp(0,T;X(𝕋2))L^{p}(0,T;X(\mathbb{T}^{2})). The space 𝒞c([0,T)×𝕋2)\mathcal{C}^{\infty}_{c}([0,T)\times\mathbb{T}^{2}) denotes the space of smooth functions with compact support in time and periodic in space.
Finally, we introduce an ad-hoc norm for the solution of the α\alpha-Euler equations, namely the α\alpha-norm defined by

α2:=L22+α()L22.\|\cdot\|_{\alpha}^{2}:=\|\cdot\|_{L^{2}}^{2}+\alpha\|\nabla(\cdot)\|_{L^{2}}^{2}. (12)

In this exposition, we use multiple times a generic constant CC. This constant does not depend on α\alpha unless the dependence is specified and even if the constant CC appears more than one time in the same computation, its value may change from one line to the next.

2.2   The α\alpha-Euler equations

For the sake of completeness, we derive the vorticity formulation of (1). Let us define

qα:=curlvαandωα:=curluα.q^{\alpha}:=\operatorname{curl}v^{\alpha}\quad\text{and}\quad\omega^{\alpha}:=\operatorname{curl}u^{\alpha}.

We recall that in dimension two the curl is a scalar and we employ the standard identities jujαujα=|uα|2/2\sum_{j}u^{\alpha}_{j}\nabla u^{\alpha}_{j}=\nabla|u^{\alpha}|^{2}/2 and curl(uαuα)=uαωα\operatorname{curl}(u^{\alpha}\cdot\nabla u^{\alpha})=u^{\alpha}\cdot\nabla\omega^{\alpha}. Taking the curl\operatorname{curl} of (1), we get

tqα+uαωααcurl(uα(Δuα))αcurl(jΔujαujα)=0.\partial_{t}q^{\alpha}+u^{\alpha}\cdot\nabla\omega^{\alpha}-\alpha\operatorname{curl}(u^{\alpha}\cdot\nabla(\operatorname{\Delta}u^{\alpha}))-\alpha\operatorname{curl}\left(\sum_{j}\operatorname{\Delta}u^{\alpha}_{j}\nabla u^{\alpha}_{j}\right)=0. (13)

We compute the two curls and obtain

curl(uα(Δuα))=2u1α1(Δu1α)+u1α21(Δu1α)+2u2α2(Δu1α)+u2α22(Δu1α)1u1α1(Δu2α)u1α12(Δu2α)1u2α2(Δu2α)u2α12(Δu2α)\operatorname{curl}(u^{\alpha}\cdot\nabla(\operatorname{\Delta}u^{\alpha}))=\partial_{2}u_{1}^{\alpha}\partial_{1}(\operatorname{\Delta}u_{1}^{\alpha})+u_{1}^{\alpha}\partial_{2}\partial_{1}(\operatorname{\Delta}u_{1}^{\alpha})+\partial_{2}u_{2}^{\alpha}\partial_{2}(\operatorname{\Delta}u_{1}^{\alpha})+u_{2}^{\alpha}\partial_{2}^{2}(\operatorname{\Delta}u_{1}^{\alpha})\\ -\partial_{1}u_{1}^{\alpha}\partial_{1}(\operatorname{\Delta}u_{2}^{\alpha})-u_{1}^{\alpha}\partial_{1}^{2}(\operatorname{\Delta}u_{2}^{\alpha})-\partial_{1}u_{2}^{\alpha}\partial_{2}(\operatorname{\Delta}u_{2}^{\alpha})-u_{2}^{\alpha}\partial_{1}\partial_{2}(\operatorname{\Delta}u_{2}^{\alpha}) (14)

and

curl(jΔujαujα)=2(Δu1α)1u1α+Δu1α21u1α+Δu2α21u2α+Δu2α21u2α1(Δu1α)2u1αΔu1α12u1α1(Δu2α)2u2αΔu2α12u2α.\operatorname{curl}\left(\sum_{j}\operatorname{\Delta}u^{\alpha}_{j}\nabla u^{\alpha}_{j}\right)=\partial_{2}(\operatorname{\Delta}u_{1}^{\alpha})\partial_{1}u_{1}^{\alpha}+\operatorname{\Delta}u_{1}^{\alpha}\partial_{2}\partial_{1}u_{1}^{\alpha}+\operatorname{\Delta}u_{2}^{\alpha}\partial_{2}\partial_{1}u_{2}^{\alpha}+\operatorname{\Delta}u_{2}^{\alpha}\partial_{2}\partial_{1}u_{2}^{\alpha}\\ -\partial_{1}(\operatorname{\Delta}u_{1}^{\alpha})\partial_{2}u_{1}^{\alpha}-\operatorname{\Delta}u_{1}^{\alpha}\partial_{1}\partial_{2}u_{1}^{\alpha}-\partial_{1}(\operatorname{\Delta}u_{2}^{\alpha})\partial_{2}u_{2}^{\alpha}-\operatorname{\Delta}u_{2}^{\alpha}\partial_{1}\partial_{2}u_{2}^{\alpha}. (15)

Therefore, by (14)-(15), simplifying the opposite terms, identity (13) becomes

tqα+uαωαα(uα)Δωαdivuαcurl(Δuα)=0.\partial_{t}q^{\alpha}+u^{\alpha}\cdot\nabla\omega^{\alpha}-\alpha(u^{\alpha}\cdot\nabla)\operatorname{\Delta}\omega^{\alpha}-\operatorname{div}u^{\alpha}\operatorname{curl}(\operatorname{\Delta}u^{\alpha})=0.

Employing the incompressibility constraint, we obtain that qα=curlvαq^{\alpha}=\operatorname{curl}v^{\alpha} satisfies

{tqα+uαqα=0,on(0,T)×𝕋2q(0,)=q0α,on𝕋2,\displaystyle\begin{cases}\partial_{t}q^{\alpha}+u^{\alpha}\cdot\nabla q^{\alpha}=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ q(0,\cdot)=q_{0}^{\alpha},&\quad\text{on}\quad\mathbb{T}^{2},\end{cases} (16)

where vαv^{\alpha} is related to the vorticity qαq^{\alpha} as

{divvα=0,on(0,T)×𝕋2curlvα=qα,on(0,T)×𝕋2.\displaystyle\begin{cases}\operatorname{div}v^{\alpha}=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \operatorname{curl}v^{\alpha}=q^{\alpha},&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}.\end{cases} (17)

The system (17) yields the existence of a stream function ψα:[0,T)×𝕋2\psi^{\alpha}:[0,T)\times\mathbb{T}^{2}\rightarrow\mathbb{R} such that vα=ψαv^{\alpha}=\nabla^{\perp}\psi^{\alpha} and qα=Δψαq^{\alpha}=-\operatorname{\Delta}\psi^{\alpha} on [0,T)×𝕋2[0,T)\times\mathbb{T}^{2}. The solution of this Poisson equation is given in terms of the Green function on the torus, under the condition 𝕋2qα𝑑x=0\int_{\mathbb{T}^{2}}q^{\alpha}dx=0, which is preserved in time by the equation at least formally. The Green function on the torus reads

G𝕋2(x,y)=k2,k0log|xy2πk|2πG_{\mathbb{T}^{2}}(x,y)=\sum_{k\in\mathbb{Z}^{2},k\neq 0}-\frac{\log|x-y-2\pi k|}{2\pi}

and the corresponding Biot-Savart kernel, which can be used to represent the solution, reads

vα=𝕋2G𝕋2(x,y)qα(x)𝑑x=kqα.v^{\alpha}=\int_{\mathbb{T}^{2}}\nabla^{\perp}G_{\mathbb{T}^{2}}(x,y)q^{\alpha}(x)dx=k\ast q^{\alpha}. (18)

This relation implies that the Calderon-Zygmund estimates hold, see [marchioro], namely

vαLpCpqαLp,p(1,),\|\nabla v^{\alpha}\|_{L^{p}}\leq C_{p}\|q^{\alpha}\|_{L^{p}},\quad\forall p\in(1,\infty), (19)

where CppC_{p}\sim p for p>2p>2.
We state the well-posedeness of the α\alpha-Euler equations whose proof can be adapted from [bi, bill] to the two-dimensional torus.

Theorem 2.1 (Well-posedeness of α\alpha-Euler equations).

Let q0αLp(𝕋2)q^{\alpha}_{0}\in L^{p}(\mathbb{T}^{2}) with p(1,)p\in(1,\infty) and 𝕋2q0α=0\int_{\mathbb{T}^{2}}q^{\alpha}_{0}=0. Then u0α=(IαΔ)1kq0αW3,p(𝕋2)u^{\alpha}_{0}=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}k\ast q^{\alpha}_{0}\in W^{3,p}(\mathbb{T}^{2}) and there exists a unique solution uαL(0,T;W3,p(𝕋2))u^{\alpha}\in L^{\infty}(0,T;W^{3,p}(\mathbb{T}^{2})) of (1). Moreover, the solution uαu^{\alpha} conserves the α\alpha-norm, namely

uα(t)α=u0αα,0tT.\|u^{\alpha}(t)\|_{\alpha}=\|u^{\alpha}_{0}\|_{\alpha},\qquad\forall 0\leq t\leq T. (20)

The conservation of the α\alpha-norm can be shown formally by considering (1) and testing it against uαu^{\alpha}, integrating over the torus, which yields

𝕋2t(uααΔuα)uαdx+𝕋2uα(uααΔuα)uα𝑑x+𝕋2j(uααΔuα)jujαuαdx=𝕋2puαdx.\int_{\mathbb{T}^{2}}\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\cdot u^{\alpha}dx+\int_{\mathbb{T}^{2}}u^{\alpha}\cdot\nabla(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\cdot u^{\alpha}dx\\ +\int_{\mathbb{T}^{2}}\sum_{j}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})_{j}\nabla u^{\alpha}_{j}\cdot u^{\alpha}dx=-\int_{\mathbb{T}^{2}}\nabla p\cdot u^{\alpha}dx.

Exploiting divuα=0\operatorname{div}u^{\alpha}=0 and integrating by part, removing the trivially zero terms, we obtain

ddtuαα22=𝕋2j,i=12i(Δujα)uiαujαdx𝕋2j,i=12ujαj(Δuiα)uiαdx=0\operatorname{\frac{d}{dt}}\frac{\|u^{\alpha}\|_{\alpha}^{2}}{2}=\int_{\mathbb{T}^{2}}\sum_{j,i=1}^{2}\partial_{i}(\operatorname{\Delta}u^{\alpha}_{j})u^{\alpha}_{i}u^{\alpha}_{j}dx-\int_{\mathbb{T}^{2}}\sum_{j,i=1}^{2}u^{\alpha}_{j}\partial_{j}(\operatorname{\Delta}u^{\alpha}_{i})u^{\alpha}_{i}dx=0

where the two terms simplified by swapping the indices in the sum.

2.3   The two-dimensional Euler equations

Let T>0T>0 be arbitrary and finite, the two-dimensional Euler equations on the torus are

{tu+uu=p,on(0,T)×𝕋2divu=0,on(0,T)×𝕋2u(0,)=u0,on𝕋2,\displaystyle\begin{cases}\partial_{t}u+u\cdot\nabla u=-\nabla p,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \operatorname{div}u=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ u(0,\cdot)=u_{0},&\quad\text{on}\quad\mathbb{T}^{2},\end{cases} (21)

which in vorticity formulation reads

{tω+uω=0,on(0,T)×𝕋2ω(0,)=ω0,on𝕋2,\displaystyle\begin{cases}\partial_{t}\omega+u\cdot\nabla\omega=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \omega(0,\cdot)=\omega_{0},&\quad\text{on}\quad\mathbb{T}^{2},\end{cases} (22)

where

{divu=0,on(0,T)×𝕋2u=kω,on(0,T)×𝕋2.\displaystyle\begin{cases}\operatorname{div}u=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ u=k\ast\omega,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}.\end{cases} (23)

We introduce the Lagrangian description of (22). Let X:[0,T)×[0,T)×𝕋2𝕋2X:[0,T)\times[0,T)\times\mathbb{T}^{2}\rightarrow\mathbb{T}^{2} be such that

{X˙t,s(x)=u(s,Xt,s(x)),s[0,T],x𝕋2Xt,t(x)=x,x𝕋2,\displaystyle\begin{cases}\dot{X}_{t,s}(x)=u(s,X_{t,s}(x)),\quad\forall s\in[0,T],\quad\forall x\in\mathbb{T}^{2}\\ X_{t,t}(x)=x,\quad\forall x\in\mathbb{T}^{2},\end{cases} (24)

for any given t[0,T]t\in[0,T]. By the theory of characteristics, if uu is smooth, we know that the unique solution of the two-dimensional Euler equations with initial datum ω0\omega_{0} satisfies

u(t,x):=(kω)(t,x),andω(t,x):=ω0(Xt,0(x)).u(t,x):=(k\ast\omega)(t,x),\quad\text{and}\quad\omega(t,x):=\omega_{0}(X_{t,0}(x)). (25)

In order to extend the definition to the non-smooth case, we need to introduce the following definition.

Definition 2.2 (Regular Lagrangian flow).

A map XL((0,T)×(0,T)×𝕋2)X\in L^{\infty}((0,T)\times(0,T)\times\mathbb{T}^{2}) is called a Lagrangian flow for the vector field uL1(0,T;L1(𝕋2))u\in L^{1}(0,T;L^{1}(\mathbb{T}^{2})) if

  • the map sXt,s(x)s\mapsto X_{t,s}(x) is an absolutely continuous solution of (24) for almost every x𝕋2x\in\mathbb{T}^{2} and any t[0,T)t\in[0,T);

  • the map xXt,s(x)x\mapsto X_{t,s}(x) is measure preserving with respect to the Lebesgue measure on the torus for any s,t[0,T)s,t\in[0,T).

The definition of Lagrangian solution in the non-smooth setting is the following.

Definition 2.3 (Lagrangian solution to the Euler equations).

Let p(1,)p\in(1,\infty) and ω0Lp(𝕋2)\omega_{0}\in L^{p}(\mathbb{T}^{2}). The couple (u,ω)LtWx1,p×LtLxp(u,\omega)\in L^{\infty}_{t}W^{1,p}_{x}\times L^{\infty}_{t}L^{p}_{x} is called Lagrangian solution to the two-dimensional Euler equations if there exists a regular Lagrangian flow in the sense of Definition 2.2 and the couple (u,ω)(u,\omega) satisfies (25) for almost every (t,x)(0,T)×𝕋2(t,x)\in(0,T)\times\mathbb{T}^{2}.

3 Quantitative strong convergence of the vorticity

In this section, we present the proof of Theorem 1.1 which is split between Proposition 3.2 and Proposition 3.7. We recall and adapt some lemmas introduced in [bi]. In the first proposition, we prove the convergence in velocity and we show it implies conservation of energy. In the second proposition, we introduce the proof of convergence in vorticity through Lagrangian techniques.

3.1   Preliminaries

In this paragraph, we show some bounds which are needed to prove the convergences in Theorem 1.1. We know that the solution uαu^{\alpha} is regular enough so that the LpL^{p} norms of qαq^{\alpha} are preserved in time, thanks to Theorem 2.1. Hence, using standard elliptic estimate and the Calderon-Zygmund inequality (19), we get

uαW1,pCvαW1,pCqαLp=Cq0αLp,p(1,).\|u^{\alpha}\|_{W^{1,p}}\leq C\|v^{\alpha}\|_{W^{1,p}}\leq C\|q^{\alpha}\|_{L^{p}}=C\|q^{\alpha}_{0}\|_{L^{p}},\quad\forall p\in(1,\infty). (26)

Moreover, the elliptic equation vα=uααΔuαv^{\alpha}=u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha} yields the additional regularity

uαW3,pCα1vαW1,pCα1qαLp=Cα1q0αLp,p(1,).\|u^{\alpha}\|_{W^{3,p}}\leq C\alpha^{-1}\|v^{\alpha}\|_{W^{1,p}}\leq C\alpha^{-1}\|q^{\alpha}\|_{L^{p}}=C\alpha^{-1}\|q^{\alpha}_{0}\|_{L^{p}},\quad\forall p\in(1,\infty). (27)

This elliptic estimate can be adapted on the torus from [gilbargtru], Theorem 8.10. The inequality (27) is not uniform with respect to α\alpha; nevertheless, it is used in Lemma 3.1 to produce an improvement of (26) to control the L2L^{2} norm of the gradient and the Laplacian of uαu^{\alpha}, even for p<2p<2. The proof is an adaptation of Proposition 3.1 in [bi]. This bound refines the proof for the convergence in velocities with respect to [lt]. Moreover, the energy conservation for the limit solution is proven as a direct consequence of this estimate.

Lemma 3.1.

Let qαLp(𝕋2)q^{\alpha}\in L^{p}(\mathbb{T}^{2}) and let uα:=(IαΔ)1kqαu^{\alpha}:=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}k\ast q^{\alpha}. Then, for every t[0,T)t\in[0,T), it holds

uα(t)L2\displaystyle\|\nabla u^{\alpha}(t)\|_{L^{2}} Cα121pqα(t)Lp,p(1,2],\displaystyle\leq C\alpha^{\frac{1}{2}-\frac{1}{p}}\|q^{\alpha}(t)\|_{L^{p}},\qquad\forall p\in(1,2],
Δuα(t)L2\displaystyle\|\operatorname{\Delta}u^{\alpha}(t)\|_{L^{2}} Cα1pqα(t)Lp,p(1,2],\displaystyle\leq C\alpha^{-\frac{1}{p}}\|q^{\alpha}(t)\|_{L^{p}},\qquad\forall p\in(1,2],
Δuα(t)L2\displaystyle\|\operatorname{\Delta}u^{\alpha}(t)\|_{L^{2}} Cα12qα(t)Lp,p2.\displaystyle\leq C\alpha^{-\frac{1}{2}}\|q^{\alpha}(t)\|_{L^{p}},\qquad\forall p\geq 2.
Proof.

For p(1,2]p\in(1,2], we interpolate (26)-(27) using Gagliardo-Nirenberg inequality. In particular to bound uα\nabla u^{\alpha}, we consider the following case

uαL2CD3uαLp1p12uαLp321p,p(1,2].\|\nabla u^{\alpha}\|_{L^{2}}\leq C\|\operatorname{D}^{3}u^{\alpha}\|_{L^{p}}^{\frac{1}{p}-\frac{1}{2}}\|\nabla u^{\alpha}\|_{L^{p}}^{\frac{3}{2}-\frac{1}{p}},\qquad\forall p\in(1,2]. (28)

Owing to (26)-(27), we obtain

uαL2Cα121pqαLp,p(1,2].\|\nabla u^{\alpha}\|_{L^{2}}\leq C\alpha^{\frac{1}{2}-\frac{1}{p}}\|q^{\alpha}\|_{L^{p}},\qquad\forall p\in(1,2].

We proceed in analogous way to control ΔuαL2\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}} and we infer

ΔuαL2CD3uαLp1puαLp11pCα1pqαLp,p(1,2].\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}}\leq C\|\operatorname{D}^{3}u^{\alpha}\|_{L^{p}}^{\frac{1}{p}}\|\nabla u^{\alpha}\|_{L^{p}}^{1-\frac{1}{p}}\leq C\alpha^{-\frac{1}{p}}\|q^{\alpha}\|_{L^{p}},\qquad\forall p\in(1,2].

Let us define vα=kqαv^{\alpha}=k\ast q^{\alpha}, thus

vα=uααΔuα.v^{\alpha}=u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha}.

We test this identity against Δuα-\operatorname{\Delta}u^{\alpha} to infer

uαL22+αΔuαL22=(vα,uα)L212uαL22+αΔuαL22vαL22,\|\nabla u^{\alpha}\|^{2}_{L^{2}}+\alpha\|\operatorname{\Delta}u^{\alpha}\|^{2}_{L^{2}}=(\nabla v^{\alpha},\nabla u^{\alpha})_{L^{2}}\implies\frac{1}{2}\|\nabla u^{\alpha}\|^{2}_{L^{2}}+\alpha\|\operatorname{\Delta}u^{\alpha}\|^{2}_{L^{2}}\leq\|\nabla v^{\alpha}\|^{2}_{L^{2}}, (29)

by Young inequality. Lastly, employing (26), we deduce

αΔuαL2CqαLp,p2,\sqrt{\alpha}\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}}\leq C\|q^{\alpha}\|_{L^{p}},\qquad\forall p\geq 2,

which concludes the proof. ∎

3.2   Strong convergence of the velocity and energy conservation

At this point, we have all the tools to show the strong convergence of the velocity.

Proposition 3.2.

Let T>0T>0 arbitrary and finite, let q0αq^{\alpha}_{0} and ω0\omega_{0} under the assumptions of Theorem 1.1 and let the couple (uα,qα)(u^{\alpha},q^{\alpha}) be the corresponding global solution to the α\alpha-Euler equations. Then, there exists a couple (u,ω)(u,\omega) such that, up to subsequences, it holds

uαustrongly in𝒞([0,T];L2(𝕋2)),u^{\alpha}\rightarrow u\quad\text{strongly in}\enskip\mathcal{C}([0,T];L^{2}(\mathbb{T}^{2})), (30)
qαωweakly- inL(0,T;Lp(𝕋2))q^{\alpha}\rightharpoonup^{*}\omega\quad\text{weakly-$*$ in}\enskip L^{\infty}(0,T;L^{p}(\mathbb{T}^{2})) (31)

and uu is a distributional solution of the Euler equations with initial datum u0=kω0u_{0}=k\ast\omega_{0}. Finally, uu conserves the kinetic energy, namely

u(t)L2=u0L2,t(0,T).\|u(t)\|_{L^{2}}=\|u_{0}\|_{L^{2}},\quad\forall t\in(0,T).
Proof.

Step 1: Weak convergences. We know that the sequence qαq^{\alpha} is bounded uniformly in L(0,T;Lp(𝕋2))L^{\infty}(0,T;L^{p}(\mathbb{T}^{2})) by Theorem 2.1. By a standard compactness argument, we have that up to (non relabelled) subsequences

qαωweakly- inL(0,T;Lp(𝕋2)).q^{\alpha}\rightharpoonup^{*}\omega\quad\text{weakly-$*$ in}\enskip L^{\infty}(0,T;L^{p}(\mathbb{T}^{2})). (32)

By (26), the corresponding velocities uαu^{\alpha} are bounded uniformly in α\alpha with respect to the LtWx1,pL^{\infty}_{t}W^{1,p}_{x}-norm, hence, up to a (sub)subsequence, we have

uαuweakly- inL(0,T;W1,p(𝕋2)),u^{\alpha}\rightharpoonup^{*}u\quad\text{weakly-$*$ in}\enskip L^{\infty}(0,T;W^{1,p}(\mathbb{T}^{2})), (33)

which implies

uαuweakly- inL(0,T;L2(𝕋2)).u^{\alpha}\rightharpoonup^{*}u\quad\text{weakly-$*$ in}\enskip L^{\infty}(0,T;L^{2}(\mathbb{T}^{2})). (34)

Step 2: Strong convergence of the velocity. We rewrite (1) as

t(uααΔuα)=div(uαuα)+αi,jji(ujαiuα)αi,jj(iujαiuα)+αi,ji(iujαujα)πα,\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})=-\operatorname{div}(u^{\alpha}\otimes u^{\alpha})\\ +\alpha\sum_{i,j}\partial_{j}\partial_{i}(u^{\alpha}_{j}\partial_{i}u^{\alpha})-\alpha\sum_{i,j}\partial_{j}(\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha})+\alpha\sum_{i,j}\partial_{i}(\partial_{i}u_{j}^{\alpha}\nabla u_{j}^{\alpha})-\nabla\pi^{\alpha}, (35)

thanks to standard tensor identities and πα=p+|uα|2/2\pi^{\alpha}=p+|u^{\alpha}|^{2}/2.
Now, we want to control the time derivative of the velocity and use Aubin-Lions Lemma, in order to show the strong convergence of uαu^{\alpha}. Let φ𝒞(𝕋2)\varphi\in\mathcal{C}^{\infty}(\mathbb{T}^{2}), then there exists φα:=(1αΔ)1φ\varphi^{\alpha}:=(1-\alpha\operatorname{\Delta})^{-1}\varphi and φα𝒞(𝕋2)\varphi^{\alpha}\in\mathcal{C}^{\infty}(\mathbb{T}^{2}). Multiplying (35) by φα\varphi^{\alpha} and integrating over 𝕋2\mathbb{T}^{2}, after some integrations by parts, we infer

𝕋2t(uααΔuα)φαdx=𝕋2(uαuα):φαdx+αj,i𝕋2ujαiuαjiφαdx+αj,i𝕋2iujαiuαjφαdxαj,i𝕋2iujαujαiφαdx𝕋2παφαdx=T1+T2+T3+T4𝕋2παφαdx.\int_{\mathbb{T}^{2}}\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\cdot\varphi^{\alpha}dx=\int_{\mathbb{T}^{2}}(u^{\alpha}\otimes u^{\alpha}):\nabla\varphi^{\alpha}dx+\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}u^{\alpha}_{j}\partial_{i}u^{\alpha}\cdot\partial_{j}\partial_{i}\varphi^{\alpha}dx\\ +\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}\cdot\partial_{j}\varphi^{\alpha}dx-\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}\partial_{i}u^{\alpha}_{j}\nabla u^{\alpha}_{j}\cdot\partial_{i}\varphi^{\alpha}dx-\int_{\mathbb{T}^{2}}\nabla\pi^{\alpha}\cdot\varphi^{\alpha}dx\\ =T_{1}+T_{2}+T_{3}+T_{4}-\int_{\mathbb{T}^{2}}\nabla\pi^{\alpha}\cdot\varphi^{\alpha}dx. (36)

On the left hand side of (36), we get

𝕋2t(uααΔuα)φαdx=𝕋2tuα(IαΔ)φαdx=𝕋2tuαφdx.\int_{\mathbb{T}^{2}}\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\cdot\varphi^{\alpha}dx=\int_{\mathbb{T}^{2}}\partial_{t}u^{\alpha}\cdot(\operatorname{I}-\alpha\operatorname{\Delta})\varphi^{\alpha}dx=\int_{\mathbb{T}^{2}}\partial_{t}u^{\alpha}\cdot\varphi dx. (37)

Hence, we need to bound every term on the right hand side of (36) to control {tuα}\{\partial_{t}u^{\alpha}\}. The control on the first term in the right hand side of (36) is straightforward

|T1|:=|𝕋2(uαuα):φαdx|CuαL22φαLCu0αα2φαH3.|T_{1}|:=\left|\int_{\mathbb{T}^{2}}(u^{\alpha}\otimes u^{\alpha}):\nabla\varphi^{\alpha}dx\right|\leq C\|u^{\alpha}\|^{2}_{L^{2}}\|\nabla\varphi^{\alpha}\|_{L^{\infty}}\leq C\|u^{\alpha}_{0}\|^{2}_{\alpha}\|\varphi^{\alpha}\|_{H^{3}}. (38)

For the other terms, we employ the bound on the α\alpha-norm given by (20). In particular, for the second term of the right hand side of (36), we obtain

|T2|:=|αj,i𝕋2ujαiuαjiφαdx|CαuαL2uαL2φαW2,Cα12u0αα2φαH4.|T_{2}|:=\left|\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}u^{\alpha}_{j}\partial_{i}u^{\alpha}\cdot\partial_{j}\partial_{i}\varphi^{\alpha}dx\right|\leq C\alpha\|u^{\alpha}\|_{L^{2}}\|\nabla u^{\alpha}\|_{L^{2}}\|\varphi^{\alpha}\|_{W^{2,\infty}}\leq C\alpha^{\frac{1}{2}}\|u^{\alpha}_{0}\|^{2}_{\alpha}\|\varphi^{\alpha}\|_{H^{4}}. (39)

Finally, the third and fourth term are respectively bounded in the following way

|T3|\displaystyle|T_{3}| :=|αj,i𝕋2iujαiuαjφαdx|CαuαL22φαLCu0αα2φαH3,\displaystyle:=\left|\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}\cdot\partial_{j}\varphi^{\alpha}dx\right|\leq C\alpha\|\nabla u^{\alpha}\|_{L^{2}}^{2}\|\nabla\varphi^{\alpha}\|_{L^{\infty}}\leq C\ \|u^{\alpha}_{0}\|^{2}_{\alpha}\|\varphi^{\alpha}\|_{H^{3}}, (40)
|T4|\displaystyle|T_{4}| :=|αj,i𝕋2iujαujαiφαdx|CαuαL22φαLCu0αα2φαH3.\displaystyle:=\left|\alpha\sum_{j,i}\int_{\mathbb{T}^{2}}\partial_{i}u^{\alpha}_{j}\nabla u^{\alpha}_{j}\cdot\partial_{i}\varphi^{\alpha}dx\right|\leq C\alpha\|\nabla u^{\alpha}\|_{L^{2}}^{2}\|\nabla\varphi^{\alpha}\|_{L^{\infty}}\leq C\ \|u^{\alpha}_{0}\|^{2}_{\alpha}\|\varphi^{\alpha}\|_{H^{3}}. (41)

Lastly, we need to estimate the pressure term. Let us consider (35) and take its divergence. Owing to the incompressibility constraint, we get

ij(uiαujα)αl(ji(ujαiulα)j(iujαiulα)+i(iujαlujα))=Δπα.\partial_{i}\partial_{j}(u^{\alpha}_{i}u^{\alpha}_{j})-\alpha\partial_{l}\left(\partial_{j}\partial_{i}(u^{\alpha}_{j}\partial_{i}u^{\alpha}_{l})-\partial_{j}(\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}_{l})+\partial_{i}(\partial_{i}u_{j}^{\alpha}\partial_{l}u_{j}^{\alpha})\right)=-\operatorname{\Delta}\pi^{\alpha}.

Here, we consider the pressure as the sum of two contributions πα=π1α+π2α\pi^{\alpha}=\pi^{\alpha}_{1}+\pi^{\alpha}_{2} such that

Δπ1α=ij(uiαujα),Δπ2α=αl(ji(ujαiulα)j(iujαiulα)+i(iujαlujα)).\operatorname{\Delta}\pi_{1}^{\alpha}=-\partial_{i}\partial_{j}(u^{\alpha}_{i}u^{\alpha}_{j}),\quad\operatorname{\Delta}\pi_{2}^{\alpha}=\alpha\partial_{l}\left(\partial_{j}\partial_{i}(u^{\alpha}_{j}\partial_{i}u^{\alpha}_{l})-\partial_{j}(\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}_{l})+\partial_{i}(\partial_{i}u_{j}^{\alpha}\partial_{l}u_{j}^{\alpha})\right).

For the term π1α\pi^{\alpha}_{1}, we notice that it can be bounded analogously to (38), which is

|𝕋2π1αφαdx|=|𝕋2j(uiαujα)φjαdx||T1|CuαL22φαH3.\left|\int_{\mathbb{T}^{2}}\nabla\pi^{\alpha}_{1}\cdot\varphi^{\alpha}dx\right|=\left|\int_{\mathbb{T}^{2}}\partial_{j}(u_{i}^{\alpha}u_{j}^{\alpha})\varphi^{\alpha}_{j}dx\right|\leq|T_{1}|\leq C\|u^{\alpha}\|^{2}_{L^{2}}\|\varphi^{\alpha}\|_{H^{3}}. (42)

The control on the term π2α\pi^{\alpha}_{2} is exactly equivalent to the ones in (39)-(40)-(41), indeed

|𝕋2π2αφαdx||T2|+|T3|+|T4|C(uαα2)φαH4.\left|\int_{\mathbb{T}^{2}}\nabla\pi^{\alpha}_{2}\cdot\varphi^{\alpha}dx\right|\leq|T_{2}|+|T_{3}|+|T_{4}|\leq C(\|u^{\alpha}\|^{2}_{\alpha})\|\varphi^{\alpha}\|_{H^{4}}. (43)

Now, the estimates for the non-linear terms (38)-(39)-(40)-(41) and the ones for the pressure (42)-(43) complete the control of the right hand side of (36). Indeed, employing the elliptic estimate φαH4CφH4\|\varphi^{\alpha}\|_{H^{4}}\leq C\|\varphi\|_{H^{4}}, we infer

tuα,φCφαH4(𝕋2)CφH4(𝕋2).\langle\partial_{t}u^{\alpha},\varphi\rangle\leq C\|\varphi^{\alpha}\|_{H^{4}(\mathbb{T}^{2})}\leq C\|\varphi\|_{H^{4}(\mathbb{T}^{2})}. (44)

Thus, tuα\partial_{t}u^{\alpha} is uniformly bounded with respect to α\alpha in LtHx4L^{\infty}_{t}H^{-4}_{x}. The immersion of L2(𝕋2)L^{2}(\mathbb{T}^{2}) is continuous in H4(𝕋2)H^{-4}(\mathbb{T}^{2}) and by (33) the velocity converges weakly-* in L(0,T;W1,p(𝕋2))L^{\infty}(0,T;W^{1,p}(\mathbb{T}^{2})) with W1,p(𝕋2)W^{1,p}(\mathbb{T}^{2}) compactly embedded in L2(𝕋2)L^{2}(\mathbb{T}^{2}). Hence, we use Aubin-Lions lemma to infer that up to a new (sub)subsequence

uαustrongly in𝒞([0,T];L2(𝕋2)).u^{\alpha}\rightarrow u\quad\text{strongly in}\enskip\mathcal{C}([0,T];L^{2}(\mathbb{T}^{2})). (45)

Step 3: Equation for the velocity. We want to show that the limit uu is a solution to the Euler equation with initial datum u0u_{0}. First, we recover strong convergence of u0αu^{\alpha}_{0} to u0u_{0} in L2(𝕋2)L^{2}(\mathbb{T}^{2}). We recall that the initial data are defined as

v0α:=kq0α,u0α:=(IαΔ)1v0α,u0:=kω0.v^{\alpha}_{0}:=k\ast q^{\alpha}_{0},\quad u^{\alpha}_{0}:=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}v^{\alpha}_{0},\quad u_{0}:=k\ast\omega_{0}.

Owing to (19) and q0αω0q^{\alpha}_{0}\to\omega_{0} in Lp(𝕋2)L^{p}(\mathbb{T}^{2}) by hypothesis, we have that

v0αL2u0L2.\|v^{\alpha}_{0}\|_{L^{2}}\to\|u_{0}\|_{L^{2}}. (46)

Then, we consider Lemma 3.1 and if p2p\geq 2, we deduce

αΔu0αL2=α(αΔu0αL2)α00.\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}}=\sqrt{\alpha}(\sqrt{\alpha}\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}})\xrightarrow{\alpha\to 0}0. (47)

Whereas, if p(1,2)p\in(1,2), we obtain

αΔu0αL2=α11p(α1pΔu0αL2)α00.\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}}=\alpha^{1-\frac{1}{p}}(\alpha^{\frac{1}{p}}\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}})\xrightarrow{\alpha\to 0}0. (48)

By (47)-(48), we infer from (46) that

u0αv0αL2=αΔu0αL20u0αL2u0L2.\|u^{\alpha}_{0}-v^{\alpha}_{0}\|_{L^{2}}=\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}}\to 0\implies\|u^{\alpha}_{0}\|_{L^{2}}\to\|u_{0}\|_{L^{2}}. (49)

Now, we are left to show that the limit uu is a distributional solution to the velocity formulation of the Euler equations. Thus, we need to pass to the limit into (35). The term t(uααΔuα)tu\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\rightarrow\partial_{t}u in the sense of distribution thanks to (34). Indeed, for any φ𝒞c((0,T)×𝕋2)\varphi\in\mathcal{C}^{\infty}_{c}((0,T)\times\mathbb{T}^{2}), it holds

t(uααΔuα),φ=0T𝕋2uα(tφ+αΔtφ)𝑑x𝑑s0T𝕋2u(tφ)𝑑x𝑑s=tu,φ.\langle\partial_{t}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha}),\varphi\rangle=\int_{0}^{T}\int_{\mathbb{T}^{2}}u^{\alpha}\cdot(-\partial_{t}\varphi+\alpha\operatorname{\Delta}\partial_{t}\varphi)dxds\\ \to\int_{0}^{T}\int_{\mathbb{T}^{2}}u\cdot(-\partial_{t}\varphi)dxds=\langle\partial_{t}u,\varphi\rangle.

Let p(1,2)p\in(1,2), considering the right hand side of (35), thanks to Lemma 3.1, we obtain

αujαiuαL1\displaystyle\|\alpha u^{\alpha}_{j}\partial_{i}u^{\alpha}\|_{L^{1}} CαuαL2uαL2Cα321pu0ααq0αLp0,\displaystyle\leq C\alpha\|u^{\alpha}\|_{L^{2}}\|\nabla u^{\alpha}\|_{L^{2}}\leq C\alpha^{\frac{3}{2}-\frac{1}{p}}\|u^{\alpha}_{0}\|_{\alpha}\|q^{\alpha}_{0}\|_{L^{p}}\rightarrow 0, (50)
αiujαiuαL1\displaystyle\|\alpha\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}\|_{L^{1}} CαuαL22Cα22pq0αLp20,\displaystyle\leq C\alpha\|\nabla u^{\alpha}\|^{2}_{L^{2}}\leq C\alpha^{2-\frac{2}{p}}\|q^{\alpha}_{0}\|^{2}_{L^{p}}\rightarrow 0,
αiujαujαL1\displaystyle\|\alpha\partial_{i}u_{j}^{\alpha}\nabla u_{j}^{\alpha}\|_{L^{1}} CαuαL22Cα22pq0αLp20.\displaystyle\leq C\alpha\|\nabla u^{\alpha}\|^{2}_{L^{2}}\leq C\alpha^{2-\frac{2}{p}}\|q^{\alpha}_{0}\|^{2}_{L^{p}}\rightarrow 0.

Analogously for p[2,)p\in[2,\infty), we use (26) to deduce

αujαiuαL1\displaystyle\|\alpha u^{\alpha}_{j}\partial_{i}u^{\alpha}\|_{L^{1}} CαuαL2uαL2Cαu0ααq0αLp0,\displaystyle\leq C\alpha\|u^{\alpha}\|_{L^{2}}\|\nabla u^{\alpha}\|_{L^{2}}\leq C\alpha\|u^{\alpha}_{0}\|_{\alpha}\|q^{\alpha}_{0}\|_{L^{p}}\rightarrow 0, (51)
αiujαiuαL1\displaystyle\|\alpha\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha}\|_{L^{1}} CαuαL22Cαq0αLp20,\displaystyle\leq C\alpha\|\nabla u^{\alpha}\|^{2}_{L^{2}}\leq C\alpha\|q^{\alpha}_{0}\|^{2}_{L^{p}}\rightarrow 0,
αiujαujαL1\displaystyle\|\alpha\partial_{i}u_{j}^{\alpha}\nabla u_{j}^{\alpha}\|_{L^{1}} CαuαL22Cαq0αLp20.\displaystyle\leq C\alpha\|\nabla u^{\alpha}\|^{2}_{L^{2}}\leq C\alpha\|q^{\alpha}_{0}\|^{2}_{L^{p}}\rightarrow 0.

After integration by parts, for any p(1,)p\in(1,\infty) inequalities (51)-(50) imply the following convergences in the sense of distribution

αi,jji(ujαiuα)0,αi,jj(iujαiuα)0andαi,ji(iujαujα)0.\alpha\sum_{i,j}\partial_{j}\partial_{i}(u^{\alpha}_{j}\partial_{i}u^{\alpha})\rightarrow 0,\quad\alpha\sum_{i,j}\partial_{j}(\partial_{i}u^{\alpha}_{j}\partial_{i}u^{\alpha})\rightarrow 0\quad\text{and}\quad\alpha\sum_{i,j}\partial_{i}(\partial_{i}u_{j}^{\alpha}\nabla u_{j}^{\alpha})\rightarrow 0.

Hence, in the right hand side of (35), we are left to pass to the limit for div(uαuα)\operatorname{div}(u^{\alpha}\otimes u^{\alpha}). However, this is implied by strong convergence of the velocity in (45).
Step 4: Energy conservation. By Theorem 2.1, we know that the α\alpha-norm of the solution is conserved, namely

uα(t)L22+αuα(t)L22=u0αL22+αu0αL22\|u^{\alpha}(t)\|^{2}_{L^{2}}+\alpha\|\nabla u^{\alpha}(t)\|^{2}_{L^{2}}=\|u^{\alpha}_{0}\|^{2}_{L^{2}}+\alpha\|\nabla u^{\alpha}_{0}\|^{2}_{L^{2}} (52)

and we want to pass to the limit as α0\alpha\to 0. Considering Lemma 3.1, we obtain

αuαL22=α22p(α1p12uαL2)2α22pC20,asα0,p(1,2).\alpha\|\nabla u^{\alpha}\|_{L^{2}}^{2}=\alpha^{2-\frac{2}{p}}\left(\alpha^{\frac{1}{p}-\frac{1}{2}}\|\nabla u^{\alpha}\|_{L^{2}}\right)^{2}\leq\alpha^{2-\frac{2}{p}}C^{2}\rightarrow 0,\quad\text{as}\enskip\alpha\rightarrow 0,\quad\forall p\in(1,2). (53)

Moreover, by (26) we get

αuαL22αqαLp20,asα0,p[2,].\alpha\|\nabla u^{\alpha}\|_{L^{2}}^{2}\leq\alpha\|q^{\alpha}\|_{L^{p}}^{2}\rightarrow 0,\quad\text{as}\enskip\alpha\rightarrow 0,\quad\forall p\in[2,\infty]. (54)

We proceed in analogous way to control u0α\nabla u^{\alpha}_{0} and we infer

αu0αL220,asα0,p(1,].\alpha\|\nabla u^{\alpha}_{0}\|_{L^{2}}^{2}\rightarrow 0,\quad\text{as}\enskip\alpha\to 0,\quad\forall p\in(1,\infty]. (55)

Employing (53)-(54) and (55), we pass to the limit into (52) to obtain the thesis

u(t)L22=u0L22,p(1,],\|u(t)\|^{2}_{L^{2}}=\|u_{0}\|^{2}_{L^{2}},\quad\forall p\in(1,\infty], (56)

where we have used the strong convergence of the velocity in L2(𝕋2)L^{2}(\mathbb{T}^{2}) expressed by (45)-(49). ∎

3.3   Strong convergence of the vorticities

We proceed to prove that the limit solution of the α\alpha-Euler equations is a Lagrangian solution of the Euler equations. This proof is analogous to the one in [crippaspirito] for the vanishing viscosity scheme. Let us introduce the transport equation as

{tρ+bρ=0,on(0,T)×𝕋2ρ(0,)=ρ0,on𝕋2,\displaystyle\begin{cases}\partial_{t}\rho+b\cdot\nabla\rho=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \rho(0,\cdot)=\rho_{0},&\quad\text{on}\quad\mathbb{T}^{2},\end{cases} (57)

with divb=0\operatorname{div}b=0. We define in the following way a renormalized solution.

Definition 3.3.

A measurable function ρ\rho is a renormalized solution of (57), if it solves in the sense of distribution

{tβ(ρ)+bβ(ρ)=0,on(0,T)×𝕋2β(ρ)(0,)=β(ρ0),on𝕋2,\displaystyle\begin{cases}\partial_{t}\beta(\rho)+b\cdot\nabla\beta(\rho)=0,&\quad\text{on}\quad(0,T)\times\mathbb{T}^{2}\\ \beta(\rho)(0,\cdot)=\beta(\rho_{0}),&\quad\text{on}\quad\mathbb{T}^{2},\end{cases}

for any β𝒞1()L()\beta\in\mathcal{C}^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}).

We first recall the following lemma, given by Theorem II.6, [dipernalions].

Lemma 3.4.

Let bb be a vector field such that

b(t,x)L1(0,T;W1,p(𝕋2)),divb=0b(t,x)\in L^{1}(0,T;W^{1,p}(\mathbb{T}^{2})),\quad\operatorname{div}b=0

and ρL(0,T;Lp(𝕋2))\rho\in L^{\infty}(0,T;L^{p}(\mathbb{T}^{2})) be a renormalized solution of the transport equation according to Definition 3.3. Let ξL(0,T;Lq(𝕋2))\xi\in L^{\infty}(0,T;L^{q}(\mathbb{T}^{2})), where q=p1pq=\frac{p-1}{p}, be a renormalized solution of the following backward transport problem

{tξdiv(bξ)=χ,in[0,T)×𝕋2ξ(T,)=ξT,in𝕋2,\displaystyle\begin{cases}-\partial_{t}\xi-\operatorname{div}(b\xi)=\chi,&\quad\text{in}\quad[0,T)\times\mathbb{T}^{2}\\ \xi(T,\cdot)=\xi_{T},&\quad\text{in}\quad\mathbb{T}^{2},\end{cases}

where χL1(0,T;Lq(𝕋2))\chi\in L^{1}(0,T;L^{q}(\mathbb{T}^{2})) and ξTLq(𝕋2)\xi_{T}\in L^{q}(\mathbb{T}^{2}). Then, it holds

0T𝕋2χρ𝑑x𝑑s=𝕋2ξ(x,0)ρ0(x)𝑑x𝕋2ξT(x)ρ(x,T)𝑑x.\int_{0}^{T}\int_{\mathbb{T}^{2}}\chi\rho dxds=\int_{\mathbb{T}^{2}}\xi(x,0)\rho_{0}(x)dx-\int_{\mathbb{T}^{2}}\xi_{T}(x)\rho(x,T)dx. (58)

With the introduction of this setting, we are able to prove that the limit ω\omega is a Lagrangian solution of the Euler equations.

Proposition 3.5.

Let (u,ω)(u,\omega) be the limit of (uα,qα)(u^{\alpha},q^{\alpha}) according to Proposition 3.2. Then, (u,ω)(u,\omega) is a Lagrangian solution of the Euler equations according to Definition 2.3.

Proof.

Step 1: Consistent limit. We begin the proof by showing that u=kωu=k\ast\omega. The starting point is the equation

uααΔuα=kqα,a.a.t(0,T).u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha}=k\ast q^{\alpha},\quad\text{a.a.}\enskip t\in(0,T).

Given a scalar test function φ𝒞c((0,T)×𝕋2)\varphi\in\mathcal{C}^{\infty}_{c}((0,T)\times\mathbb{T}^{2}), we have that

0T𝕋2(uααΔuα)φ𝑑x𝑑s=0T𝕋2(kqα)φ𝑑x𝑑s.\int_{0}^{T}\int_{\mathbb{T}^{2}}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\varphi dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}(k\ast q^{\alpha})\varphi dxds.

Considering the left hand side, we obtain

limα00T𝕋2(uααΔuα)φ𝑑x𝑑s==limα00T𝕋2uαφ𝑑x𝑑slimα0α0T𝕋2uαΔφdxds=0T𝕋2uφ𝑑x𝑑s,\lim_{\alpha\rightarrow 0}\int_{0}^{T}\int_{\mathbb{T}^{2}}(u^{\alpha}-\alpha\operatorname{\Delta}u^{\alpha})\varphi dxds=\\ =\lim_{\alpha\rightarrow 0}\int_{0}^{T}\int_{\mathbb{T}^{2}}u^{\alpha}\varphi dxds-\lim_{\alpha\rightarrow 0}\alpha\int_{0}^{T}\int_{\mathbb{T}^{2}}u^{\alpha}\operatorname{\Delta}\varphi dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}u\varphi dxds,

where we exploited the smoothness of φ\varphi and (45). Instead, for the right hand side we have

0T𝕋2(kqα)φ𝑑x𝑑s=0T𝕋2(𝕋2k𝕋2(xy)qα(x,s))φ(x,s)𝑑x𝑑s==0T𝕋2qα(x,s)(𝕋2k𝕋2(xy)φ(x,s))𝑑x𝑑s=0T𝕋2qα(kφ)𝑑x𝑑s.\int_{0}^{T}\int_{\mathbb{T}^{2}}(k\ast q^{\alpha})\varphi dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}\left(\int_{\mathbb{T}^{2}}k_{\mathbb{T}^{2}}(x-y)q^{\alpha}(x,s)\right)\varphi(x,s)dxds=\\ =\int_{0}^{T}\int_{\mathbb{T}^{2}}q^{\alpha}(x,s)\left(\int_{\mathbb{T}^{2}}k_{\mathbb{T}^{2}}(x-y)\varphi(x,s)\right)dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}q^{\alpha}(k\ast\varphi)dxds.

Here, being φ𝒞c((0,T)×𝕋2)\varphi\in\mathcal{C}^{\infty}_{c}((0,T)\times\mathbb{T}^{2}), by Young inequality we have

kφLqkL1φLqC.\|k\ast\varphi\|_{L^{q}}\leq\|k\|_{L^{1}}\|\varphi\|_{L^{q}}\leq C. (59)

Employing (32), taking the limit as α0+\alpha\rightarrow 0^{+} and switching back the convolution, we infer

0T𝕋2uφ𝑑x𝑑s=0T𝕋2ω(kφ)𝑑x𝑑s=0T𝕋2(kω)φ𝑑x𝑑s,\int_{0}^{T}\int_{\mathbb{T}^{2}}u\varphi dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}\omega(k\ast\varphi)dxds=\int_{0}^{T}\int_{\mathbb{T}^{2}}(k\ast\omega)\varphi dxds,

for any φ𝒞c((0,T)×𝕋2)\varphi\in\mathcal{C}^{\infty}_{c}((0,T)\times\mathbb{T}^{2}). This implies u=kωu=k\ast\omega for almost every (t,x)(0,T)×𝕋2(t,x)\in(0,T)\times\mathbb{T}^{2} as wanted.
Step 2: Lagrangian solution. We show that (u,ω)(u,\omega) is a Lagrangian solution. We only need to prove it for p(1,2)p\in(1,2), because for p2p\geq 2 it follows directly from uniqueness of the solution of the transport equation (see [dipernalions], Theorem II.3). Let us consider (16), let χ𝒞((0,T)×𝕋2)\chi\in\mathcal{C}^{\infty}((0,T)\times\mathbb{T}^{2}) and let us define the backward transport problem for uαu^{\alpha} of the form

{tξαdiv(uαξα)=χ,in(0,T)×𝕋2ξα(T,)=0,in𝕋2\displaystyle\begin{cases}-\partial_{t}\xi^{\alpha}-\operatorname{div}(u^{\alpha}\xi^{\alpha})=\chi,&\quad\text{in}\quad(0,T)\times\mathbb{T}^{2}\\ \xi^{\alpha}(T,\cdot)=0,&\quad\text{in}\quad\mathbb{T}^{2}\end{cases} (60)

and the limit backward problem

{tξdiv(uξ)=χ,in(0,T)×𝕋2ξ(T,)=0,in𝕋2.\displaystyle\begin{cases}-\partial_{t}\xi-\operatorname{div}(u\xi)=\chi,&\quad\text{in}\quad(0,T)\times\mathbb{T}^{2}\\ \xi(T,\cdot)=0,&\quad\text{in}\quad\mathbb{T}^{2}.\end{cases} (61)

Thanks to the stability theorem in DiPerna-Lions, [dipernalions] Theorem II.4, it holds that ξαξ\xi^{\alpha}\rightarrow\xi in 𝒞([0,T];Lq(𝕋2))\mathcal{C}([0,T];L^{q}(\mathbb{T}^{2})), for every q[1,]q\in[1,\infty]. Since uαu^{\alpha} is smooth, qαq^{\alpha} satisfies

𝕋20Tχqα𝑑x𝑑s=𝕋2ξα(x,0)q0α(x)𝑑x,\int_{\mathbb{T}^{2}}\int_{0}^{T}\chi q^{\alpha}dxds=\int_{\mathbb{T}^{2}}\xi^{\alpha}(x,0)q_{0}^{\alpha}(x)dx, (62)

where ξα\xi_{\alpha} solves (60). We recall that qαωq^{\alpha}\rightharpoonup^{*}\omega in L(0,T;Lp(𝕋2))L^{\infty}(0,T;L^{p}(\mathbb{T}^{2})) and q0αω0q^{\alpha}_{0}\rightarrow\omega_{0} in Lp(𝕋2)L^{p}(\mathbb{T}^{2}), therefore passing to the limit in (62), we obtain

𝕋20Tχω𝑑x𝑑s=𝕋2ξ(x,0)ω0(x)𝑑x,\int_{\mathbb{T}^{2}}\int_{0}^{T}\chi\omega dxds=\int_{\mathbb{T}^{2}}\xi(x,0)\omega_{0}(x)dx, (63)

where ξ\xi is the unique solution in 𝒞([0,T];Lq(𝕋2))\mathcal{C}([0,T];L^{q}(\mathbb{T}^{2})) of (61). Using Lemma 3.4 on the limit backward problem, we infer

𝕋20TχωL𝑑x𝑑s=𝕋2ξ(x,0)ω0(x)𝑑x,\int_{\mathbb{T}^{2}}\int_{0}^{T}\chi\omega_{L}dxds=\int_{\mathbb{T}^{2}}\xi(x,0)\omega_{0}(x)dx, (64)

where ωL\omega_{L} is the unique renormalized solution, thus Lagrangian of the transport equation (57) with velocity field uu and initial datum u0u_{0} (cf. [dipernalions] Theorem II.3). Subtracting (63) and (64) we get

𝕋20Tχ(ωLω)𝑑x𝑑s=0,χ𝒞((0,T)×𝕋2),\int_{\mathbb{T}^{2}}\int_{0}^{T}\chi(\omega_{L}-\omega)dxds=0,\quad\forall\chi\in\mathcal{C}^{\infty}((0,T)\times\mathbb{T}^{2}),

which implies that ω=ωL\omega=\omega_{L}. ∎

We notice that for the α\alpha-Euler equations, we can introduce the flow map XαX^{\alpha} using the classical theory of characteristic (cf. (24)-(25)), since the velocity field uαu^{\alpha} Lipschitz due to the embedding W3,pW1,W^{3,p}\hookrightarrow W^{1,\infty}. Knowing that the limit solution is Lagrangian for any p(1,)p\in(1,\infty), we want to study the convergence of the flows as it has been done in [odeest].

Lemma 3.6.

Let T>0T>0 arbitrary and finite and let (u0α,q0α)(u^{\alpha}_{0},q^{\alpha}_{0}) be under the assumptions of Theorem 1.1 and let (uα,qα)(u^{\alpha},q^{\alpha}) be the sequence of unique solutions of α\alpha-Euler equations according to Theorem 2.1. Let (u,ω)(u,\omega) be the limit of (uα,qα)(u^{\alpha},q^{\alpha}) obtained in Proposition 3.2 and let Xt,sαX^{\alpha}_{t,s} and Xt,sX_{t,s} be the corresponding Lagrangian flows according to (24) and Definition 2.2. Then, it holds

𝕋2𝐝(Xt,sα(x),Xt,s(x))𝑑xC|log(uαuLt1Lx1)|,\int_{\mathbb{T}^{2}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\leq\frac{C}{\left|\log\left(\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}\right)\right|}, (65)

where the constant CC depends on uLt1Lxp\|\nabla u\|_{L^{1}_{t}L^{p}_{x}} and on TT.

Proof.

Let δ=uαuLt1Lx1\delta=\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}. Let us define the quantity

gδ(s):=𝕋2log(|Xt,sα(x)Xt,s(x)|δ+1)𝑑x.g_{\delta}(s):=\int_{\mathbb{T}^{2}}\log\left(\frac{|X^{\alpha}_{t,s}(x)-X_{t,s}(x)|}{\delta}+1\right)dx. (66)

Let us consider xlog(1+xδ)x\mapsto\log\left(1+\frac{x}{\delta}\right), increasing in [0,)[0,\infty). We use Chebyshev inequality to infer

2({x𝕋2|𝐝(Xt,sα(x),Xt,s(x))>ε})1log(εδ+1)𝕋2log(|Xt,sα(x)Xt,s(x)|δ+1)𝑑x,\mathcal{L}^{2}\left(\left\{x\in\mathbb{T}^{2}|\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))>\varepsilon\right\}\right)\leq\frac{1}{\log\left(\frac{\varepsilon}{\delta}+1\right)}\int_{\mathbb{T}^{2}}\log\left(\frac{|X^{\alpha}_{t,s}(x)-X_{t,s}(x)|}{\delta}+1\right)dx, (67)

for every ε>0\varepsilon>0. We have used that on the torus it holds 𝐝(Xt,sα(x),Xt,s(x))|Xt,sα(x)Xt,s(x)|\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))\leq|X^{\alpha}_{t,s}(x)-X_{t,s}(x)|. We split the integral over the torus in two complementary sets as

𝕋2𝐝(Xt,sα(x),Xt,s(x))𝑑x{x𝕋2|𝐝(Xt,sα(x),Xt,s(x))ε}𝐝(Xt,sα(x),Xt,s(x))𝑑x+{x𝕋2|𝐝(Xt,sα(x),Xt,s(x))>ε}𝐝(Xt,sα(x),Xt,s(x))𝑑x.\int_{\mathbb{T}^{2}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\leq\int_{\left\{x\in\mathbb{T}^{2}|\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))\leq\varepsilon\right\}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\\ +\int_{\left\{x\in\mathbb{T}^{2}|\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))>\varepsilon\right\}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx.

We use (67) to infer

𝕋2𝐝(Xt,sα(x),Xt,s(x))𝑑xε+1log(εδ+1)𝕋2log(|Xt,sα(x)Xt,s(x)|δ+1)𝑑x.\int_{\mathbb{T}^{2}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\leq\varepsilon+\frac{1}{\log\left(\frac{\varepsilon}{\delta}+1\right)}\int_{\mathbb{T}^{2}}\log\left(\frac{|X^{\alpha}_{t,s}(x)-X_{t,s}(x)|}{\delta}+1\right)dx. (68)

The inequality (68) holds true for any ε>0\varepsilon>0, thus we can choose ε=δ\varepsilon=\sqrt{\delta}. We recall that δ\delta goes to zero as α0\alpha\rightarrow 0, by definition. Therefore, we can take a value α\alpha for which δ<1\delta<1, which yields

1log(1δ+1)1|logδ|=2|logδ|.\frac{1}{\log\left(\frac{1}{\sqrt{\delta}}+1\right)}\leq\frac{1}{\left|\log\sqrt{\delta}\right|}=\frac{2}{\left|\log\delta\right|}.

Substituting into (68) with (66), we infer

𝕋2𝐝(Xt,sα(x),Xt,s(x))𝑑xδ+2|logδ|gδ(s).\int_{\mathbb{T}^{2}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\leq\sqrt{\delta}+\frac{2}{\left|\log\delta\right|}g_{\delta}(s).

By Definition 2.2 and (66), we know gδ(t)=0g_{\delta}(t)=0 and

gδ(s)=tsgδ(τ)𝑑τts𝕋2|X˙t,τα(x)X˙t,τ(x)||Xt,τα(x)Xt,τ(x)|+δ𝑑x𝑑τts𝕋2|uα(τ,Xt,τα)u(τ,Xt,τα)||Xt,τα(x)Xt,τ(x)|+δ𝑑x𝑑τ+ts𝕋2|u(τ,Xt,τα)u(τ,Xt,τ)||Xt,τα(x)Xt,τ(x)|+δ𝑑x𝑑τ,g_{\delta}(s)=\int_{t}^{s}g_{\delta}^{\prime}(\tau)d\tau\leq\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|\dot{X}^{\alpha}_{t,\tau}(x)-\dot{X}_{t,\tau}(x)|}{|X^{\alpha}_{t,\tau}(x)-X_{t,\tau}(x)|+\delta}dxd\tau\\ \leq\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X^{\alpha}_{t,\tau})|}{|X^{\alpha}_{t,\tau}(x)-X_{t,\tau}(x)|+\delta}dxd\tau+\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|u(\tau,X^{\alpha}_{t,\tau})-u(\tau,X_{t,\tau})|}{|X^{\alpha}_{t,\tau}(x)-X_{t,\tau}(x)|+\delta}dxd\tau, (69)

where we have summed ±uα(τ,Xt,τα)\pm u^{\alpha}(\tau,X^{\alpha}_{t,\tau}) in the numerator and we have used the triangular inequality. The first term in the right hand side of (69) is controlled as

ts𝕋2|uα(τ,Xt,τα)u(τ,Xt,τα)||Xt,τα(x)Xt,τ(x)|+δ𝑑x𝑑τts𝕋2|uα(τ,Xt,τα)u(τ,Xt,τα)|δ𝑑x𝑑τuuαLt1Lx1δC,\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X^{\alpha}_{t,\tau})|}{|X^{\alpha}_{t,\tau}(x)-X_{t,\tau}(x)|+\delta}dxd\tau\\ \leq\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X^{\alpha}_{t,\tau})|}{\delta}dxd\tau\leq\frac{\|u-u^{\alpha}\|_{L^{1}_{t}L^{1}_{x}}}{\delta}\leq C, (70)

thanks to (69). We are left to bound the second term in the right hand side of (69). Let \mathcal{M} be the maximal function operator defined on L1L^{1} functions as

f(x):=supr>012(Br)Br(x)|f(ξ)|𝑑ξ,x𝕋2.\mathcal{M}f(x):=\sup_{r>0}\frac{1}{\mathcal{L}^{2}(B_{r})}\int_{B_{r}(x)}|f(\xi)|d\xi,\quad\forall x\in\mathbb{T}^{2}.

We recall that for any fLpf\in L^{p} with p(1,]p\in(1,\infty], it holds

fLpCfLp.\|\mathcal{M}f\|_{L^{p}}\leq C\|f\|_{L^{p}}.

Moreover, for any fW1,1f\in W^{1,1}, there exists a set 𝒩𝕋2\mathcal{N}\subset\mathbb{T}^{2} such that 2(𝒩)=0\mathcal{L}^{2}(\mathcal{N})=0 and

|f(x)f(y)|C𝐝(x,y)(Df(x)+Df(y)),x,y𝕋𝒩.|f(x)-f(y)|\leq C\operatorname{\mathbf{d}}(x,y)(\mathcal{M}\operatorname{D}f(x)+\mathcal{M}\operatorname{D}f(y)),\quad\forall x,y\in\mathbb{T}\setminus\mathcal{N}. (71)

This allows us to deduce

ts𝕋2|u(τ,Xt,τα)u(τ,Xt,τ)||Xt,τα(x)Xt,τ(x)|+δ𝑑x𝑑τCts𝕋2|u(τ,Xt,τα)|𝑑x𝑑τ++Cts𝕋2|u(τ,Xt,τ)|𝑑x𝑑τCuLt1Lx1CuLt1Lxp,\int_{t}^{s}\int_{\mathbb{T}^{2}}\frac{|u(\tau,X^{\alpha}_{t,\tau})-u(\tau,X_{t,\tau})|}{|X^{\alpha}_{t,\tau}(x)-X_{t,\tau}(x)|+\delta}dxd\tau\leq C\int_{t}^{s}\int_{\mathbb{T}^{2}}|\mathcal{M}\nabla u(\tau,X^{\alpha}_{t,\tau})|dxd\tau+\\ +C\int_{t}^{s}\int_{\mathbb{T}^{2}}|\mathcal{M}\nabla u(\tau,X_{t,\tau})|dxd\tau\leq C\|\nabla u\|_{L^{1}_{t}L^{1}_{x}}\leq C\|\nabla u\|_{L^{1}_{t}L^{p}_{x}}, (72)

thanks to the measure preserving property of the flows Xt,sαX^{\alpha}_{t,s} and Xt,sX_{t,s}. Substituting (70) and (72) into (69), we obtain

𝕋2𝐝(Xt,sα(x),Xt,s(x))𝑑xδ+2|logδ|CC|logδ|,\int_{\mathbb{T}^{2}}\operatorname{\mathbf{d}}(X^{\alpha}_{t,s}(x),X_{t,s}(x))dx\leq\sqrt{\delta}+\frac{2}{\left|\log\delta\right|}C\leq\frac{C}{\left|\log\delta\right|},

exploiting that δ1log(δ)\sqrt{\delta}\leq\frac{1}{\log(\delta)}, for δ<1\delta<1. Recalling that δ=uαuLt1Lx1\delta=\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}, we have the thesis. ∎

Lemma 3.6 allows to prove a result on the strong convergence of the vorticities.

Proposition 3.7.

Let T>0T>0 arbitrary and finite and let (u0α,q0α)(u^{\alpha}_{0},q^{\alpha}_{0}) be under the assumptions of Theorem 1.1. Let the couple (uα,qα)(u^{\alpha},q^{\alpha}) be the corresponding global solution to the α\alpha-Euler equations. Let (u,ω)(u,\omega) be the limit of (uα,qα)(u^{\alpha},q^{\alpha}) obtained in Proposition 3.2, then for any δ>0\delta>0, there exists K(δ,ω0)K(\delta,\omega_{0}) such that for α\alpha small enough it holds

supt[0,T)qα(t)ω(t)Lpδ+K(δ,ω0)|log(uαuLt1Lx1)|+q0αω0Lp,\sup_{t\in[0,T)}\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq\delta+\frac{K(\delta,\omega_{0})}{\left|\log\left(\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}\right)\right|}+\|q^{\alpha}_{0}-\omega_{0}\|_{L^{p}}, (73)

which implies

qαωstrongly in𝒞([0,T];Lp(𝕋2)).q^{\alpha}\rightarrow\omega\quad\text{strongly in}\enskip\mathcal{C}([0,T];L^{p}(\mathbb{T}^{2})). (74)
Proof.

We recall that Lipschitz function are dense in Lp(𝕋2)L^{p}(\mathbb{T}^{2}), for every p<p<\infty. Thus, we take a sequence of Lipschitz function {ω0j}j\{\omega_{0}^{j}\}_{j\in\mathbb{N}}, such that ω0jω\omega_{0}^{j}\rightarrow\omega as jj\rightarrow\infty in Lp(𝕋2)L^{p}(\mathbb{T}^{2}). We notice

qα(t)ω(t)Lp=(𝕋2|q0α(Xt,0α)ω0(Xt,0)|p𝑑x)1p(𝕋2|ω0j(Xt,0α)ω0(Xt,0α)|p𝑑x)1p+(𝕋2|ω0j(Xt,0)ω0(Xt,0)|p𝑑x)1p+(𝕋2|ω0j(Xt,0α)ω0j(Xt,0)|p𝑑x)1p+(𝕋2|q0α(Xt,0α)ω0(Xt,0α)|p𝑑x)1p.\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}=\left(\int_{\mathbb{T}^{2}}|q_{0}^{\alpha}(X^{\alpha}_{t,0})-\omega_{0}(X_{t,0})|^{p}dx\right)^{\frac{1}{p}}\leq\\ \left(\int_{\mathbb{T}^{2}}|\omega_{0}^{j}(X^{\alpha}_{t,0})-\omega_{0}(X^{\alpha}_{t,0})|^{p}dx\right)^{\frac{1}{p}}+\left(\int_{\mathbb{T}^{2}}|\omega_{0}^{j}(X_{t,0})-\omega_{0}(X_{t,0})|^{p}dx\right)^{\frac{1}{p}}+\\ \left(\int_{\mathbb{T}^{2}}|\omega_{0}^{j}(X^{\alpha}_{t,0})-\omega_{0}^{j}(X_{t,0})|^{p}dx\right)^{\frac{1}{p}}+\left(\int_{\mathbb{T}^{2}}|q_{0}^{\alpha}(X^{\alpha}_{t,0})-\omega_{0}(X^{\alpha}_{t,0})|^{p}dx\right)^{\frac{1}{p}}.

Knowing that ω0j\omega_{0}^{j} are Lipschitz and using the estimate on the flows (65), we deduce

qα(t)ω(t)Lp2ω0j(t)ω0(t)Lp+K(δ,ω0)|log(uαuLt1Lx1)|+q0α(t)ω0(t)Lp.\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq 2\|\omega_{0}^{j}(t)-\omega_{0}(t)\|_{L^{p}}+\frac{K(\delta,\omega_{0})}{\left|\log\left(\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}\right)\right|}+\|q^{\alpha}_{0}(t)-\omega_{0}(t)\|_{L^{p}}. (75)

If we send first α0\alpha\rightarrow 0 and then jj\rightarrow\infty, by employing the strong convergence of uu in LtLx2L^{\infty}_{t}L^{2}_{x}, thus in Lt1Lx1L^{1}_{t}L^{1}_{x}, we get that qαωq^{\alpha}\rightarrow\omega strongly in 𝒞tLxp\mathcal{C}_{t}L^{p}_{x}. Given δ>0\delta>0, we can take the sequence {ω0j}j\{\omega_{0}^{j}\}_{j\in\mathbb{N}} such that δ2ω0j(t)ω0(t)Lp\delta\geq 2\|\omega_{0}^{j}(t)-\omega_{0}(t)\|_{L^{p}}, for any jj\in\mathbb{N}, and infer (73) from (75).
The convergence (74) follows from uαuLt1Lx10\|u^{\alpha}-u\|_{L^{1}_{t}L^{1}_{x}}\rightarrow 0 as α0\alpha\to 0. ∎

This result completes the proof of Theorem 1.1.

4 Rate of convergence for bounded vorticity

In this section, we analyse the rate of convergence for the initial vorticity in L(𝕋2)L^{\infty}(\mathbb{T}^{2}). The main tool we employ is a standard generalization of Gronwall inequality, known as Osgood’s lemma. We recall here the statement as it has been proven in [cheminosgood].

Lemma 4.1 (Osgood Lemma).

Let ρ\rho be a positive Borelian function, γ\gamma a locally integrable positive function, μ\mu a continuous increasing function and (x)=x1drμ(r)\mathcal{M}(x)=\int_{x}^{1}\frac{dr}{\mu(r)}. Let us assume that, for a strictly positive number η\eta, the function ρ\rho satisfies

ρ(t)η+t0tγ(s)μ(ρ(s))𝑑s.\rho(t)\leq\eta+\int_{t_{0}}^{t}\gamma(s)\mu(\rho(s))ds.

Then, we have

(η)(ρ(t))t0tγ(s)𝑑s.\mathcal{M}(\eta)-\mathcal{M}(\rho(t))\leq\int_{t_{0}}^{t}\gamma(s)ds.

4.1   Rate of convergence for the velocities

As a first step to prove Theorem 1.2, we use a technique introduced by Chemin in [chemin] to control the convergence of the velocities, employing Lemma 4.1.

Proposition 4.2.

Let q0αq^{\alpha}_{0} and ω0\omega_{0} be under the assumptions of Theorem 1.2. Let u0α:=(IαΔ)1kq0αu^{\alpha}_{0}:=(\operatorname{I}-\alpha\operatorname{\Delta})^{-1}k\ast q^{\alpha}_{0} and u0:=kω0u_{0}:=k\ast\omega_{0}, then

γ0α:=u0αu0L2+αΔu0αL2α00and letα¯>0be s.t.γ0α12,αα¯.\gamma_{0}^{\alpha}:=\|u^{\alpha}_{0}-u_{0}\|_{L^{2}}+\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|_{L^{2}}\xrightarrow{\alpha\to 0}0\quad\text{and let}\quad\overline{\alpha}>0\quad\text{be s.t.}\quad\gamma_{0}^{\alpha}\leq\frac{1}{2},\quad\forall\alpha\leq\overline{\alpha}.

Let (uα,qα)(u^{\alpha},q^{\alpha}) be the solution to the α\alpha-Euler equations with initial datum q0αq^{\alpha}_{0} and let (u,ω)(u,\omega) be its limit, which is the unique Yudovich solution to the Euler equations. Then, there exist two constants C1C_{1} and C2C_{2} (depending on M:=ω0LM:=\|\omega_{0}\|_{L^{\infty}}) such that, if α(0,α¯]\alpha\in(0,\overline{\alpha}] and T>0T>0 satisfy

αexp{2(22exp(C2T))}γ0α(C1T)2,\alpha\leq\frac{\exp\{2(2-2\exp(C_{2}T))\}-\gamma_{0}^{\alpha}}{(C_{1}T)^{2}},

it holds

u(t)uα(t)L2exp{22exp(C2t)}(C1αT+γ0α)exp(C2t)+Cα:=K(α,t),tT.\|u(t)-u^{\alpha}(t)\|_{L^{2}}\leq\exp\{2-2\exp(-C_{2}t)\}(C_{1}\sqrt{\alpha}T+\gamma^{\alpha}_{0})^{\exp(-C_{2}t)}+C\sqrt{\alpha}:=K(\alpha,t),\quad\forall t\leq T.
Proof.

We observe that considering the assumptions of Theorem 1.2, we have that Theorem 1.1 holds true for every p<p<\infty. Therefore, we know that qαq^{\alpha} converges to ω\omega in 𝒞([0,T];Lp(𝕋2))\mathcal{C}([0,T];L^{p}(\mathbb{T}^{2})), for every p<p<\infty and it is uniformly bounded in L((0,T)×𝕋2)L^{\infty}((0,T)\times\mathbb{T}^{2}). Hence, we deduce that ωL((0,T)×𝕋2)\omega\in L^{\infty}((0,T)\times\mathbb{T}^{2}), which implies that uu is the unique Yudovich solution to the Euler equations (cf. [yudo]). Now, set

yα:=vαuandzα(t)=yα(t,)L2.y^{\alpha}:=v^{\alpha}-u\quad\text{and}\quad z^{\alpha}(t)=\|y^{\alpha}(t,\cdot)\|_{L^{2}}. (76)

Let us take the difference between the Euler equations and the α\alpha-Euler equations and test it against yαy^{\alpha}. Summing the terms ±𝕋2vαuyαdx\pm\int_{\mathbb{T}^{2}}v^{\alpha}\cdot\nabla u\cdot y^{\alpha}dx, we obtain

12ddtyα2=𝕋2(vα)uyα𝑑x𝕋2(uα)vαyα𝑑x𝕋2j=1,2vjαiujαyiαdx𝕋2(vα)uyα𝑑x+𝕋2(u)uyα𝑑x.\frac{1}{2}\operatorname{\frac{d}{dt}}\|y^{\alpha}\|^{2}=\int_{\mathbb{T}^{2}}(v^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx-\int_{\mathbb{T}^{2}}(u^{\alpha}\cdot\nabla)v^{\alpha}\cdot y^{\alpha}dx-\int_{\mathbb{T}^{2}}\sum_{j=1,2}v_{j}^{\alpha}\partial_{i}u_{j}^{\alpha}y_{i}^{\alpha}dx\\ -\int_{\mathbb{T}^{2}}(v^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx+\int_{\mathbb{T}^{2}}(u\cdot\nabla)u\cdot y^{\alpha}dx.

Using the definitions of vαv^{\alpha} and yαy^{\alpha}, after some computations, we infer

12ddtyα2=α𝕋2(Δuα)uyα𝑑x𝕋2(uα)yαyα𝑑x𝕋2j=1,2ujαiujαyiαdx+α𝕋2j=1,2Δujαiujαyiαdx𝕋2(yα)uyα𝑑x.\frac{1}{2}\operatorname{\frac{d}{dt}}\|y^{\alpha}\|^{2}=-\alpha\int_{\mathbb{T}^{2}}(\operatorname{\Delta}u^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx-\int_{\mathbb{T}^{2}}(u^{\alpha}\cdot\nabla)y^{\alpha}\cdot y^{\alpha}dx\\ -\int_{\mathbb{T}^{2}}\sum_{j=1,2}u_{j}^{\alpha}\partial_{i}u_{j}^{\alpha}y_{i}^{\alpha}dx+\alpha\int_{\mathbb{T}^{2}}\sum_{j=1,2}\operatorname{\Delta}u_{j}^{\alpha}\partial_{i}u_{j}^{\alpha}y_{i}^{\alpha}dx-\int_{\mathbb{T}^{2}}(y^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx.

Owing to the incompressibility constrains, we cancel out some terms and we obtain

12ddtyα2=α𝕋2(Δuα)uyα𝑑x+α𝕋2j=1,2Δujαiujαyiαdx𝕋2(yα)uyα𝑑x.\frac{1}{2}\operatorname{\frac{d}{dt}}\|y^{\alpha}\|^{2}=-\alpha\int_{\mathbb{T}^{2}}(\operatorname{\Delta}u^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx+\alpha\int_{\mathbb{T}^{2}}\sum_{j=1,2}\operatorname{\Delta}u_{j}^{\alpha}\partial_{i}u_{j}^{\alpha}y_{i}^{\alpha}dx-\int_{\mathbb{T}^{2}}(y^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx. (77)

We first control the last term in the right hand side of (77) with a technique analogous to the one used in [chemin]. Using Hölder inequality and (19), we get

𝕋2(yα)uyα𝑑xuLp(𝕋2|yα|2pp1𝑑x)p1pCp(𝕋2|yα|2pp1𝑑x)p1p.\int_{\mathbb{T}^{2}}(y^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx\leq\|\nabla u\|_{L^{p}}\left(\int_{\mathbb{T}^{2}}|y^{\alpha}|^{\frac{2p}{p-1}}dx\right)^{\frac{p-1}{p}}\leq Cp\left(\int_{\mathbb{T}^{2}}|y^{\alpha}|^{\frac{2p}{p-1}}dx\right)^{\frac{p-1}{p}}. (78)

We observe that yα=k(ωqα)y^{\alpha}=-k\ast(\omega-q^{\alpha}). By Calderon-Zygmund inequality, for q>2q>2, we have

yαLCyαW1,qCωqαLqCω0LqC.\|y^{\alpha}\|_{L^{\infty}}\leq C\|y^{\alpha}\|_{W^{1,q}}\leq C\|\omega-q^{\alpha}\|_{L^{q}}\leq C\|\omega_{0}\|_{L^{q}}\leq C. (79)

The right hand side of (78) is controlled as

(𝕋2|yα|2pp1𝑑x)p1p=(𝕋2|yα|2|yα|pp1𝑑x)p1pyαL22p1pyαL2pCyαL22p1p.\left(\int_{\mathbb{T}^{2}}|y^{\alpha}|^{\frac{2p}{p-1}}dx\right)^{\frac{p-1}{p}}=\left(\int_{\mathbb{T}^{2}}|y^{\alpha}|^{2}|y^{\alpha}|^{\frac{p}{p-1}}dx\right)^{\frac{p-1}{p}}\leq\|y^{\alpha}\|_{L^{2}}^{2\frac{p-1}{p}}\|y^{\alpha}\|_{L^{\infty}}^{\frac{2}{p}}\leq C\|y^{\alpha}\|_{L^{2}}^{2\frac{p-1}{p}}. (80)

We are left to control the first two terms in the right hand side of (77). By Lemma 3.1, we deduce αΔuαL2Cq0αL2\sqrt{\alpha}\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}}\leq C\|q^{\alpha}_{0}\|_{L^{2}}. Exploiting this fact, we control the first two terms in the right hand side of (77) with some applications of Hölder inequality and we deduce

α|𝕋2(Δuα)uyα𝑑x|+α|𝕋2j=1,2Δujαiujαyiαdx|α(αΔuαL2)(uL2+uαL2)yαLCα,\alpha\left|\int_{\mathbb{T}^{2}}(\operatorname{\Delta}u^{\alpha}\cdot\nabla)u\cdot y^{\alpha}dx\right|+\alpha\left|\int_{\mathbb{T}^{2}}\sum_{j=1,2}\operatorname{\Delta}u_{j}^{\alpha}\partial_{i}u_{j}^{\alpha}y_{i}^{\alpha}dx\right|\\ \leq\sqrt{\alpha}(\sqrt{\alpha}\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}})(\|\nabla u\|_{L^{2}}+\|\nabla u^{\alpha}\|_{L^{2}})\|y^{\alpha}\|_{L^{\infty}}\leq C\sqrt{\alpha}, (81)

where we have used (79). Finally, we substitute back into (77) the estimates (78)-(81) and (80) and by definition (76), we obtain the ordinary differential inequality

zα(t)Cp(zα(t))(11p)+Cα,p2.z^{\alpha}(t)^{\prime}\leq Cp(z^{\alpha}(t))^{\left(1-\frac{1}{p}\right)}+C\sqrt{\alpha},\quad\forall p\geq 2. (82)

Thanks to (47)-(48)-(49), we deduce (8), namely

γ0α:=u0αu0L2+αΔu0α0,asα0.\gamma_{0}^{\alpha}:=\|u^{\alpha}_{0}-u_{0}\|_{L^{2}}+\alpha\|\operatorname{\Delta}u^{\alpha}_{0}\|\rightarrow 0,\quad\text{as}\quad\alpha\to 0.

Thus, let α¯\overline{\alpha} be such that γ0α1/2\gamma_{0}^{\alpha}\leq 1/2, for every αα¯\alpha\leq\overline{\alpha}. Hereafter until the end of the proof, we consider α(0,α¯]\alpha\in(0,\overline{\alpha}] and an application of triangular inequality yields

zα(0)γ0α12.z^{\alpha}(0)\leq\gamma_{0}^{\alpha}\leq\frac{1}{2}. (83)

We want to apply a substitution for which we need zα(t)(0,1]z^{\alpha}(t)\in(0,1]. We know that zα(0)1/2z^{\alpha}(0)\leq 1/2 by (83) and we define a “modified” zαz^{\alpha} as

zδα(t):=zα(t)+δwhereδ(0,12]such thatzδα(0)<1andzδα(t)>0,t0.z^{\alpha}_{\delta}(t):=z^{\alpha}(t)+\delta\quad\text{where}\enskip\delta\in\left(0,\frac{1}{2}\right]\quad\text{such that}\quad z^{\alpha}_{\delta}(0)<1\quad\text{and}\quad z^{\alpha}_{\delta}(t)>0,\enskip\forall t\geq 0.

From this definition, it is easy to check that (82) still holds for zδαz^{\alpha}_{\delta}, namely

zδα(t)Cp(zδα(t))(11p)+Cα,p2.z^{\alpha}_{\delta}(t)^{\prime}\leq Cp(z^{\alpha}_{\delta}(t))^{\left(1-\frac{1}{p}\right)}+C\sqrt{\alpha},\quad\forall p\geq 2. (84)

We notice that (84) holds for any p2p\geq 2, therefore we consider p=2log(zδα(t))p=2-\log(z^{\alpha}_{\delta}(t)), which is well defined thanks to zδα(t)(0,1]z^{\alpha}_{\delta}(t)\in(0,1]. We infer with some simple computations

(zδα)C(2log(zδα))zδα+Cα,(z^{\alpha}_{\delta})^{\prime}\leq C(2-\log(z^{\alpha}_{\delta}))z^{\alpha}_{\delta}+C\sqrt{\alpha},

thanks to (zδα)1log(zδα)=e1(z^{\alpha}_{\delta})^{-\frac{1}{\log(z^{\alpha}_{\delta})}}=e^{-1}. Integrating over time, we obtain

zδα(t)zδα(0)C1αT+0tC2(2log(zδα(s)))zδα(s)𝑑s,t(0,T].z^{\alpha}_{\delta}(t)-z^{\alpha}_{\delta}(0)\leq C_{1}\sqrt{\alpha}T+\int_{0}^{t}C_{2}(2-\log(z^{\alpha}_{\delta}(s)))z^{\alpha}_{\delta}(s)ds,\qquad\forall t\in(0,T].

We apply Lemma 4.1 with μ(x):=x(2log(x))\mu(x):=x(2-\log(x)), (x):=log(2log(x))log2\mathcal{M}(x):=\log(2-\log(x))-\log 2, η:=C1αT+zδα(0)\eta:=C_{1}\sqrt{\alpha}T+z^{\alpha}_{\delta}(0) and γ(s):=C2\gamma(s):=C_{2}, yielding

log(2log(zδα(t))+log(2log(C1α+zδα(0)))C2t,-\log(2-\log(z^{\alpha}_{\delta}(t))+\log(2-\log(C_{1}\sqrt{\alpha}+z^{\alpha}_{\delta}(0)))\leq C_{2}t,

which implies

zδα(t)exp{22exp(C2t)}(C1αT+zδα(0))exp(C2t).z^{\alpha}_{\delta}(t)\leq\exp\{2-2\exp(-C_{2}t)\}(C_{1}\sqrt{\alpha}T+z^{\alpha}_{\delta}(0))^{\exp(-C_{2}t)}. (85)

We recall that we required to start with an initial datum such that zδα(0)<1z^{\alpha}_{\delta}(0)<1 and the inequality (85) holds as long as zδα(t)1z^{\alpha}_{\delta}(t)\leq 1. This statement holds true as long as we consider

αexp{2(22exp(C2T))}γ0α(C1T)2.\alpha\leq\frac{\exp\{2(2-2\exp(C_{2}T))\}-\gamma_{0}^{\alpha}}{(C_{1}T)^{2}}.

We can pass to the limit as δ0\delta\to 0 and deduce that (85) still holds if we take zα(t)z^{\alpha}(t) in place of zδα(t)z^{\alpha}_{\delta}(t). We complete the proof of the theorem through an application of triangular inequality, yielding

uuαL2uvαL2+vαuαL2zα+αΔuαL2exp{22exp(C2t)}(C1αT+γ0α)exp(C2t)+Cα,\|u-u^{\alpha}\|_{L^{2}}\leq\|u-v^{\alpha}\|_{L^{2}}+\|v^{\alpha}-u^{\alpha}\|_{L^{2}}\leq z^{\alpha}+\alpha\|\operatorname{\Delta}u^{\alpha}\|_{L^{2}}\\ \leq\exp\{2-2\exp(-C_{2}t)\}(C_{1}\sqrt{\alpha}T+\gamma_{0}^{\alpha})^{\exp(-C_{2}t)}+C\sqrt{\alpha},

where we have used (85) and Lemma 3.1. ∎

4.2   Rate of convergence for the vorticities

In this paragraph, we conclude the proof of Theorem 1.2. We use the “Chemin-type” estimate given by Proposition 4.2 to control the rate of convergence of the flows. This allows us to bound the rate of convergence of the vorticities using the definition of Lagrangian solution.

Proposition 4.3.

Let ω0\omega_{0} and q0αq^{\alpha}_{0} be under the assumptions of Theorem 1.2, let (uα,qα)(u^{\alpha},q^{\alpha}) be the solution to the α\alpha-Euler equations with initial datum q0αq^{\alpha}_{0} and let (u,ω)(u,\omega) be its limit, which is the unique Yudovich solution to the Euler equations.
Then, there exists a value α0=α0(T,M,ω0)\alpha_{0}=\alpha_{0}(T,M,\omega_{0}) and a continuous function ψω0,p,M:++\psi_{\omega_{0},p,M}:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} vanishing at zero, such that for every αα0\alpha\leq\alpha_{0} holds

supt(0,T)qα(t)ω(t)LpCM11pmax{ψω0,p,M(K(α,t)),K(α,t)exp(CT)2p},\sup_{t\in(0,T)}\|q^{\alpha}(t)-\omega(t)\|_{L^{p}}\leq CM^{1-\frac{1}{p}}\max\left\{\psi_{\omega_{0},p,M}(K(\alpha,t)),K(\alpha,t)^{\frac{\exp{(-CT)}}{2p}}\right\},

where CC depends on MM.

Proof.

Step 1: Quantitative estimate for the flows. Let Xt,sαX^{\alpha}_{t,s} and Xt,sX_{t,s} be the two Lagrangian flows relative to uαu^{\alpha} and uu. Subtracting the two respective definitions of the flows, multiplying them by Xt,sαXt,sX^{\alpha}_{t,s}-X_{t,s} and integrating over time, we obtain

|Xt,sαXt,s|22=ts(uα(τ,Xt,τα)u(τ,Xt,τ))(Xt,ταXt,τ)𝑑τ.\frac{|X^{\alpha}_{t,s}-X_{t,s}|^{2}}{2}=\int_{t}^{s}\left(u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X_{t,\tau})\right)\left(X^{\alpha}_{t,\tau}-X_{t,\tau}\right)d\tau. (86)

We proceed by summing to the right hand side ±u(τ,Xt,τα)\pm u(\tau,X^{\alpha}_{t,\tau}) and we infer

ts|(uα(τ,Xt,τα)u(τ,Xt,τ))(Xt,ταXt,τ)|𝑑τts|(uα(τ,Xt,τα)u(τ,Xt,τα))(Xt,ταXt,τ)|𝑑τ+ts|(u(τ,Xt,τα)u(τ,Xt,τα))(Xt,ταXt,τ)|𝑑τ.\int_{t}^{s}\left|\left(u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X_{t,\tau})\right)\left(X^{\alpha}_{t,\tau}-X_{t,\tau}\right)\right|d\tau\leq\\ \int_{t}^{s}\left|\left(u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X^{\alpha}_{t,\tau})\right)\left(X^{\alpha}_{t,\tau}-X_{t,\tau}\right)\right|d\tau+\int_{t}^{s}\left|\left(u(\tau,X^{\alpha}_{t,\tau})-u(\tau,X^{\alpha}_{t,\tau})\right)\left(X^{\alpha}_{t,\tau}-X_{t,\tau}\right)\right|d\tau.

Integrating over the torus, we bound the first term owing to the measure preserving property of Xt,ταX^{\alpha}_{t,\tau} and using Young inequality. The second term is bounded thanks to (71), hence we infer

𝕋2ts|(uα(τ,Xt,τα)u(τ,Xt,τ))(Xt,ταXt,τ)|𝑑τ𝑑x12uαuLtLx2|ts|+𝕋2ts|Xt,ταXt,τ|22𝑑τ𝑑x+𝕋2ts|Xt,ταXt,τ|22[u(τ,)(Xt,τα)+u(τ,)(Xt,τ))]dτdx.\int_{\mathbb{T}^{2}}\int_{t}^{s}\left|\left(u^{\alpha}(\tau,X^{\alpha}_{t,\tau})-u(\tau,X_{t,\tau})\right)\left(X^{\alpha}_{t,\tau}-X_{t,\tau}\right)\right|d\tau dx\\ \leq\frac{1}{2}\|u^{\alpha}-u\|_{L^{\infty}_{t}L^{2}_{x}}|t-s|+\int_{\mathbb{T}^{2}}\int_{t}^{s}\frac{|X^{\alpha}_{t,\tau}-X_{t,\tau}|^{2}}{2}d\tau dx\\ +\int_{\mathbb{T}^{2}}\int_{t}^{s}\frac{|X^{\alpha}_{t,\tau}-X_{t,\tau}|^{2}}{2}\left[\mathcal{M}\nabla u(\tau,\cdot)(X^{\alpha}_{t,\tau})+\mathcal{M}\nabla u(\tau,\cdot)(X_{t,\tau}))\right]d\tau dx. (87)

By substituting (86) into (87) and employing Hölder inequality on the last term, we get

𝕋2|Xt,sαXt,s|22𝑑x12uαuLtLx2|ts|+𝕋2ts|Xt,ταXt,τ|22𝑑τ𝑑x+uLpts𝕋2(|Xt,ταXt,τ|22)p1p𝑑s𝑑x,\int_{\mathbb{T}^{2}}\frac{|X^{\alpha}_{t,s}-X_{t,s}|^{2}}{2}dx\leq\frac{1}{2}\|u^{\alpha}-u\|_{L^{\infty}_{t}L^{2}_{x}}|t-s|+\int_{\mathbb{T}^{2}}\int_{t}^{s}\frac{|X^{\alpha}_{t,\tau}-X_{t,\tau}|^{2}}{2}d\tau dx\\ +\|\nabla u\|_{L^{p}}\int_{t}^{s}\int_{\mathbb{T}^{2}}\left(\frac{|X^{\alpha}_{t,\tau}-X_{t,\tau}|^{2}}{2}\right)^{\frac{p-1}{p}}dsdx, (88)

where we have used that Xt,ταX^{\alpha}_{t,\tau} and Xt,τX_{t,\tau} take values in 𝕋2\mathbb{T}^{2}.
Let y(t,s)=𝕋2(|Xt,sαXt,s|22)𝑑xy(t,s)=\int_{\mathbb{T}^{2}}\left(\frac{|X^{\alpha}_{t,s}-X_{t,s}|^{2}}{2}\right)dx, then the equation (88) can be written as

{y(t,s)K(α,t)+st(y(t,τ)+Cpy(t,τ)p1p)𝑑τy(t,t)=0.\displaystyle\begin{cases}y(t,s)&\leq K(\alpha,t)+\int^{t}_{s}\left(y(t,\tau)+Cpy(t,\tau)^{\frac{p-1}{p}}\right)d\tau\\ y(t,t)&=0.\end{cases}

Noticing the similarity with (82), we can proceed in analogous way because y(0,1)y\in(0,1). We define p=2log(y(t,τ))p=2-\log(y(t,\tau)) and applying Lemma 4.1, we obtain

log(2log(y(t,s))+log(2log(K(α,t))C(ts),-\log(2-\log(y(t,s))+\log(2-\log(K(\alpha,t))\leq C(t-s),

which implies

𝕋2|Xt,sαXt,s|2𝑑x2exp{22ec(ts)}K(α,t)eCT.\int_{\mathbb{T}^{2}}|X^{\alpha}_{t,s}-X_{t,s}|^{2}dx\leq 2\exp\left\{2-2e^{-c(t-s)}\right\}K(\alpha,t)e^{-CT}. (89)

Step 2: Quantitative estimate for the vorticities. Since ω0L(𝕋2)\omega_{0}\in L^{\infty}(\mathbb{T}^{2}) and therefore in L1(𝕋2)L^{1}(\mathbb{T}^{2}), there exists a modulus of continuity in L1L^{1}, that is a continuous function ψω0()\psi_{\omega_{0}}(\cdot) vanishing at zero, such that

ω0(+h)ω0()L1(𝕋2)ψω0(|h|),for anyh𝕋2.\|\omega_{0}(\cdot+h)-\omega_{0}(\cdot)\|_{L^{1}(\mathbb{T}^{2})}\leq\psi_{\omega_{0}}(|h|),\quad\text{for any}\quad h\in\mathbb{T}^{2}.

Therefore for any ε>0\varepsilon>0, we get

qα(t)ω(t)L1(𝕋2)\displaystyle\|q^{\alpha}(t)-\omega(t)\|_{L^{1}(\mathbb{T}^{2})} =q0α(Xt,0α)ω0(Xt,0)L1(𝕋2)\displaystyle=\|q^{\alpha}_{0}(X^{\alpha}_{t,0})-\omega_{0}(X_{t,0})\|_{L^{1}(\mathbb{T}^{2})}
𝐝(Xt,0α(x),Xt,0(x))ε|ω0(Xt,0α(x))ω0(Xt,0(x))|𝑑x\displaystyle\leq\int_{\operatorname{\mathbf{d}}\left(X^{\alpha}_{t,0}(x),X_{t,0}(x)\right)\leq\varepsilon}\left|\omega_{0}(X^{\alpha}_{t,0}(x))-\omega_{0}(X_{t,0}(x))\right|dx
+𝐝(Xt,0α(x),Xt,0(x))>ε|ω0(Xt,0α(x))ω0(Xt,0(x))|𝑑x\displaystyle+\int_{\operatorname{\mathbf{d}}\left(X^{\alpha}_{t,0}(x),X_{t,0}(x)\right)>\varepsilon}\left|\omega_{0}(X^{\alpha}_{t,0}(x))-\omega_{0}(X_{t,0}(x))\right|dx
+q0α(Xt,0α)ω0(Xt,0α)L1\displaystyle+\|q^{\alpha}_{0}(X^{\alpha}_{t,0})-\omega_{0}(X^{\alpha}_{t,0})\|_{L^{1}}
ψω0(ε)+2ω0L(𝕋2)ε2𝕋2|Xt,sαXt,s|2𝑑x+q0α(Xt,0α)ω0(Xt,0α)L1.\displaystyle\leq\psi_{\omega_{0}}(\varepsilon)+\frac{2\|\omega_{0}\|_{L^{\infty}(\mathbb{T}^{2})}}{\varepsilon^{2}}\int_{\mathbb{T}^{2}}|X^{\alpha}_{t,s}-X_{t,s}|^{2}dx+\|q^{\alpha}_{0}(X^{\alpha}_{t,0})-\omega_{0}(X^{\alpha}_{t,0})\|_{L^{1}}.

Exploiting the fact that Xt,sαX^{\alpha}_{t,s} is measure preserving and using inequality (89), we infer

qα(t)ω(t)L1(𝕋2)ψω0(ε)+2ω0L(𝕋2)ε2{22ec(ts)}K(α,t)eCT+q0αω0L1.\|q^{\alpha}(t)-\omega(t)\|_{L^{1}(\mathbb{T}^{2})}\leq\psi_{\omega_{0}}(\varepsilon)+\frac{2\|\omega_{0}\|_{L^{\infty}(\mathbb{T}^{2})}}{\varepsilon^{2}}\left\{2-2e^{-c(t-s)}\right\}K(\alpha,t)e^{-CT}+\|q^{\alpha}_{0}-\omega_{0}\|_{L^{1}}. (90)

Finally, we take ε=K(α,t)eCT4\varepsilon=K(\alpha,t)e^{-\frac{CT}{4}} and interpolate LpL^{p} between L1L^{1} and LL^{\infty} to obtain the thesis. ∎

Combining Proposition 4.2 with Proposition 4.3 completes the proof of Theorem 1.2. Lastly, the proof of Corollary 1.3 follows directly from the definition of Besov spaces.

Proof of Corollary 1.3.

We prove a control on the modulus of continuity under the additional hypothesis ω0Bp,s(𝕋2)\omega_{0}\in B^{s}_{p,\infty}(\mathbb{T}^{2}). A well known characterization of the Besov spaces Bp,s(𝕋2)B^{s}_{p,\infty}(\mathbb{T}^{2}), for s(0,1)s\in(0,1), is given by

Bp,s(𝕋2)={f(x)Lp(𝕋2)s.t.suph𝕋2{0}f(+h)f()Lp|h|s<}.B^{s}_{p,\infty}(\mathbb{T}^{2})=\left\{f(x)\in L^{p}(\mathbb{T}^{2})\enskip\text{s.t.}\enskip\sup_{h\in\mathbb{T}^{2}\setminus\{0\}}\frac{\|f(\cdot+h)-f(\cdot)\|_{L^{p}}}{|h|^{s}}<\infty\right\}. (91)

Hence, the characterization (91) implies

ω0(+h)ω0()LpC|h|s,\|\omega_{0}(\cdot+h)-\omega_{0}(\cdot)\|_{L^{p}}\leq C|h|^{s},

thus, we can take ψω0(|h|)|h|s\psi_{\omega_{0}}(|h|)\simeq|h|^{s} into (10) and therefore (11) follows. ∎

References