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

Sharp conditions for energy balance in two-dimensional incompressible ideal flow with external force

Fabian Jin1 Samuel Lanthaler2 Milton C. Lopes Filho3  and  Helena J. Nussenzveig Lopes3 1 Seminar for applied mathematics
ETH Zurich
Raemistrasse 101, 8092 Zurich, Switzerland
[email protected] 2 Computing and Mathematical Sciences
California Institute of Technology
1200 East California Boulevard, Pasadena, CA–91125, USA
[email protected] 3 Instituto de Matemática
Universidade Federal do Rio de Janeiro
Cidade Universitária – Ilha do Fundão
Caixa Postal 68530
21941-909 Rio de Janeiro, RJ – BRAZIL
[email protected] [email protected]
Abstract.

Smooth solutions of the forced incompressible Euler equations satisfy an energy balance, where the rate-of-change in time of the kinetic energy equals the work done by the force per unit time. Interesting phenomena such as turbulence are closely linked with rough solutions which may exhibit inviscid dissipation, or, in other words, for which energy balance does not hold. This article provides a characterization of energy balance for physically realizable weak solutions of the forced incompressible Euler equations, i.e. solutions which are obtained in the limit of vanishing viscosity. More precisely, we show that, in the two-dimensional periodic setting, strong convergence of the zero-viscosity limit is both necessary and sufficient for energy balance of the limiting solution, under suitable conditions on the external force. As a consequence, we prove energy balance for a general class of solutions with initial vorticity belonging to rearrangement-invariant spaces, and going beyond Onsager’s critical regularity.

1. Introduction

The present work concerns weak solutions of the forced incompressible Euler equations on a two-dimensional periodic domain, focusing on solutions that arise as vanishing viscosity limits. The Euler equations describe the motion of an idealized fluid in the absence of friction and other diffusive effects, and formally satisfy the principle of conservation of energy [MajdaBertozzi]. Without an external force, conservation of energy is the conservation in time of the square of the L2L^{2}-norm of the fluid velocity. When allowing for an external force in the equations of motion, the conservation of energy instead takes the form of an identity in which the rate-of-change of the kinetic energy equals the work of the external force per unit time [LN2022].

While smooth solutions conserve energy, this may not be the case for weak solutions. In fact, weak solutions which dissipate energy are an essential part of Kolmogorov’s K41 theory of homogeneous 3D turbulence. In 1949, L. Onsager argued that only ideal flows with a third of a derivative or less may exhibit inviscid dissipation, a statement later referred to as the Onsager Conjecture, see [Kolmogorov41a, Kolmogorov41b, Frisch1995, ES2006]. There is substantial recent work connected with this issue. Inviscid dissipation is indeed ruled out for weak solutions at higher than 1/31/3 regularity, see [CET1994, CCFS08]. Starting with the pioneering work of DeLellis and Székelyhidi [LS2009], convex integration techniques have been employed to rigorously prove that dissipative Hölder continuous solutions exist in three spatial dimensions [DS2017, Isett2018, BLSV2019], up to the Onsager-critical regularity. Despite this extensive body of work, it is unclear whether such weak solutions can be obtained in the zero-viscosity limit, and hence their physical significance remains unclear.

For solutions obtained in the zero-viscosity limit in two dimensions, Cheskidov, Nussenzveig Lopes, Lopes Filho and Shvydkoy [CLNS2016] prove energy conservation under the assumption that the initial vorticity is pp-th power integrable, for p>1p>1, and in the absence of forcing. This is surprising in view of the fact that for p<3/2p<3/2, such solutions go beyond Onsager-criticality (see [Shvydkoy2010] for a discussion of Onsager-criticality), and their result hints at non-trivial dynamic constraints for solutions obtained in the zero-viscosity limit. Following [CLNS2016], solutions obtained in the zero-viscosity limit will be hereafter referred to as “physically realizable”.

Still in the absence of forcing, the results of [CLNS2016] have subsequently been extended to provide a complete characterization of physically realizable energy-conservative solutions in [LMPP2021a], where it is shown that energy conservation of such solutions is equivalent to the strong convergence of the zero-viscosity limit. In a different direction, Ciampa [ciampa2022energy] derives sufficient conditions for energy conservation when the fluid domain is the full plane. The forced periodic case has been considered in [LN2022], where sufficient conditions for energy balance of physically realizable solutions were derived based on suitable LpL^{p}-integrability assumptions on the vorticity and the curl of the external force.

External forcing is a natural mechanism for the generation of small scales in incompressible flow, and small-scale motion is an inherent feature of turbulence. Hence, finding precise conditions which rule out inviscid dissipation in the presence of an external force is particularly pertinent. In view of the characterization of physically realizable energy conservative solutions in the unforced case [LMPP2021a], one might hope that a similar characterization could also be obtained for energy balanced solutions when considering external forcing. This is the motivation for the present work.

We derive necessary and sufficient conditions for energy balance in the zero-viscosity limit for solutions of the two-dimensional Euler equations with external force. Under suitable assumptions on the force, we show that energy balance in the limit is equivalent to the strong convergence in the zero-viscosity limit. We then discuss how the results obtained here sharpen those of the previous work [LN2022], and allow for an extension of those results to vorticity in general rearrangement-invariant spaces with compact embedding in H1H^{-1}. To prove such an extension, we derive novel a priori bounds for the rearrangement-invariant maximal vorticity function for solutions of the Navier-Stokes equations. To the best of our knowledge such a result is not contained in the literature on the two-dimensional Navier-Stokes equations. Our work also fills a gap in [LMPP2021a, Corollary 2.13], where these bounds were asserted without proof.

We briefly discuss two results related to our work. First, in [BD2022], E. Bruè and C. De Lellis construct a sequence of viscous approximations, converging weakly to a solution of the incompressible Euler equations on the three-dimensional torus 𝕋3\mathbb{T}^{3}, exhibiting both anomalous dissipation along the sequence, see Definition 2.6, and inviscid dissipation. This is an illustration of the sharpness of our result, in particular since the implication “energy balance \Longrightarrow strong convergence” holds in any spatial dimension, see Remark 3.9. Second, in [Cheskidov23], A. Cheskidov, carries out an extensive discussion of the vanishing viscosity limit in the three-dimensional torus, constructing a vanishing viscosity sequence which exhibits both anomalous and inviscid dissipation, and, surprisingly, an example of inviscid dissipation without anomalous dissipation along the viscous approximation. All examples mentioned above require forcing. For details, and additional related work, see [BD2022, Cheskidov23] and references therein.

Overview: In Section 2, we recall several definitions before stating our main result, Theorem 2.8, that strong convergence in the zero-viscosity limit is both necessary and sufficient for energy conservation of physically realizable solutions. The proof of this theorem is detailed in Section 3; necessity is shown in subsection 3.2, sufficiency in subsection 3.3. Applications of our main result to rearrangement-invariant spaces can be found in Section 4; we refer to subsection 4.1 for a discussion of LpL^{p}-bounded vorticity, subsection 4.2 for a priori estimates in more general rearrangement-invariant spaces, and subsection 4.3 for sufficient conditions for energy balance of solutions with vorticity in the Lorentz space L(1,2)L^{(1,2)}, the largest rearrangement-invariant space with continuous embedding in H1H^{-1} (cf. Theorem 4.18). The derivation of our a priori estimates in rearrangement-invariant spaces is based on operator splitting. We include in Appendix A the required results on convergence of these particular operator splitting approximations.

2. Problem setting

We study the incompressible Euler equations with initial data u0u_{0}, external forcing ff and subject to periodic boundary conditions, given by

{tu+uu+p=f,in 𝕋2×(0,T),divu=0,in 𝕋2×[0,T],u=u0,at 𝕋2×{0}.\displaystyle\left\{\begin{array}[]{ll}\partial_{t}u+u\cdot\nabla u+\nabla p=f,&\text{in }\mathbb{T}^{2}\times(0,T),\\ \operatorname{div}u=0,&\text{in }\mathbb{T}^{2}\times[0,T],\\ u=u_{0},&\text{at }\mathbb{T}^{2}\times\{0\}.\end{array}\right. (2.4)

Above 𝕋2\mathbb{T}^{2} denotes the two-dimensional flat torus. Throughout this work, we fix a time-horizon T>0T>0. We also assume, without loss of generality, that the forcing ff is divergence-free as, otherwise, the gradient part can be absorbed in the pressure term. In the following, we are interested in the evolution of the kinetic energy 12u(t)Lx22\frac{1}{2}\|u(t)\|_{L^{2}_{x}}^{2} of weak solutions of the system (2.4).

Definition 2.1.

Fix T>0T>0, let u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}) be a divergence-free vector field. Assume fL1(0,T;L2(𝕋2))f\in L^{1}(0,T;L^{2}(\mathbb{T}^{2})), divf=0\operatorname{div}f=0. A vector field uL(0,T;L2(𝕋2))u\in L^{\infty}(0,T;L^{2}(\mathbb{T}^{2})) is a weak solution of (2.4), if

  1. (1)

    for all divergence-free test vector fields ϕCc(𝕋2×[0,T))\phi\in C_{c}^{\infty}(\mathbb{T}^{2}\times[0,T)), we have

    0T𝕋2{utϕ+uu:ϕ}𝑑x𝑑t+𝕋2u0(x)ϕ(x,0)𝑑x=0T𝕋2fϕ𝑑x𝑑t,\int_{0}^{T}\int_{\mathbb{T}^{2}}\left\{u\cdot\partial_{t}\phi+u\otimes u:\nabla\phi\right\}\,dx\,dt+\int_{\mathbb{T}^{2}}u_{0}(x)\cdot\phi(x,0)\,dx\\ =\int_{0}^{T}\int_{\mathbb{T}^{2}}f\cdot\phi\,dx\,dt, (2.5)
  2. (2)

    for almost every t[0,T]t\in[0,T], divu(,t)=0\operatorname{div}u({\,\cdot\,},t)=0 holds in the sense of distributions.

Existence of weak solutions in the sense of Definition 2.1 can be established under additional assumptions, such as LpL^{p}-bounds, p1p\geq 1, on the initial vorticity ω0=curl(u0)\omega_{0}=\operatorname{curl}(u_{0}) and on the curl of the forcing [LN2022].

Before moving forward let us comment on notation. We will use the subscript ‘c’ in a function space to denote elements of the space whose support is compact; we use, whenever convenient, subscripts ‘t’ and ‘x’ to denote time and spatial dependence respectively, so that Lt2Hx1L^{2}_{t}H^{1}_{x} is shorthand for L2((0,T);H1(𝕋2))L^{2}((0,T);H^{1}(\mathbb{T}^{2})).

Given that the Euler equations describe the motion of an ideal fluid, i.e. one for which friction and other dissipative effects are neglected, it is reasonable to expect that a change in kinetic energy can only be caused by the action of the external forcing ff. Following [LN2022], we recall the following definition.

Definition 2.2.

We say uL((0,T);L2(𝕋2))u\in L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})) is an energy balanced weak solution if uu is a weak solution of the incompressible Euler equations with forcing fL1((0,T);L2(𝕋2))f\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})), divf=0\operatorname{div}f=0, such that

12u(t)Lx22=12u0Lx22+0tf,uLx2𝑑τ,\displaystyle\frac{1}{2}\|u(t)\|^{2}_{L^{2}_{x}}=\frac{1}{2}\|u_{0}\|^{2}_{L^{2}_{x}}+\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau, (2.6)

for almost every t[0,T]t\in[0,T]. Here, f,uLx2(t):=𝕋2f(x,t)u(x,t)𝑑x\langle f,u\rangle_{L^{2}_{x}}(t):=\int_{\mathbb{T}^{2}}f(x,t)\cdot u(x,t)\,dx denotes the Lx2L^{2}_{x}-inner product.

Remark 2.3.

Note that the right-hand-side of (2.6) is continuous with respect to tt. Therefore, redefining uu on a set of measure zero in [0,T][0,T], we will assume hereafter that an energy balanced weak solution uu satisfies (2.6) for all t[0,T]t\in[0,T] and, hence, tu(t)L2t\mapsto\|u(t)\|_{L^{2}} is continuous.

In the absence of additional constraints, weak solutions with initial data u0Lx2u_{0}\in L^{2}_{x} are not unique and may not satisfy energy balance (or even energy conservation, without external forcing), see [szekelyhidi2011weak] for an example with vortex sheet data. It is thus natural to impose additional constraints. Such constraints arise, for example, when considering the Euler equations (2.4) as the zero-viscosity limit of the physically relevant Navier-Stokes equations:

{tuν+uνuν+pν=νΔuν+fν,in 𝕋2×(0,T),div(uν)=0,in 𝕋2×[0,T)uν=u0ν,at 𝕋2×{0}.\displaystyle\left\{\begin{array}[]{ll}\partial_{t}u^{\nu}+u^{\nu}\cdot\nabla u^{\nu}+\nabla p^{\nu}=\nu\Delta u^{\nu}+f^{\nu},&\text{in }\mathbb{T}^{2}\times(0,T),\\ \operatorname{div}(u^{\nu})=0,&\text{in }\mathbb{T}^{2}\times[0,T)\\ u^{\nu}=u^{\nu}_{0},&\text{at }\mathbb{T}^{2}\times\{0\}.\end{array}\right. (2.10)

In contrast to the incompressible Euler equations, it is well-known that the initial value problem (2.10) is well-posed for u0νL2(𝕋2)u^{\nu}_{0}\in L^{2}(\mathbb{T}^{2}), fνL1((0,T);L2(𝕋2))f^{\nu}\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})), and the solution uνu^{\nu} belongs to L((0,T);L2(𝕋2))L2((0,T);H1(𝕋2))L^{\infty}((0,T);L^{2}(\mathbb{T}^{2}))\cap L^{2}((0,T);H^{1}(\mathbb{T}^{2})), e.g. [Lions] and references therein. Furthermore, since we consider only two dimensional flows, solutions of (2.10) satisfy the following energy identity for all t[0,T]t\in[0,T],

12uν(t)Lx22\displaystyle\frac{1}{2}\|u^{\nu}(t)\|^{2}_{L^{2}_{x}} =12u0νLx22ν0tων(τ)Lx22𝑑τ+0tfν,uνLx2𝑑τ.\displaystyle=\frac{1}{2}\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}-\nu\int_{0}^{t}\|\omega^{\nu}(\tau)\|^{2}_{L^{2}_{x}}\,d\tau+\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau. (2.11)

Above, ων=curl(uν)\omega^{\nu}=\operatorname{curl}(u^{\nu}) denotes the vorticity of the flow uνu^{\nu}. Formally, the energy dissipation term ν0tων(τ)Lx22𝑑τ\nu\int_{0}^{t}\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau vanishes when ν=0\nu=0, corresponding to the energy balance relation (2.6).

Remark 2.4.

Observe that, since uνL((0,T);L2(𝕋2))L2((0,T);H1(𝕋2))u^{\nu}\in L^{\infty}((0,T);L^{2}(\mathbb{T}^{2}))\cap L^{2}((0,T);H^{1}(\mathbb{T}^{2})) it follows immediately from (2.11) that tuν(t)Lx2t\mapsto\|u^{\nu}(t)\|_{L^{2}_{x}} is continuous on [0,T][0,T].

Let us recall the definition of physically realizable solutions of (2.4):

Definition 2.5.

A weak solution uL((0,T);L2(𝕋2))u\in L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})) of the incompressible Euler equations with (divergence-free) forcing fL1((0,T);L2(𝕋2))f\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})) is physically realizable, if there exists a sequence ν0\nu\to 0, initial data u0νL2(𝕋2)u_{0}^{\nu}\in L^{2}(\mathbb{T}^{2}) and forces fνL1((0,T);L2(𝕋2))f^{\nu}\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})), divfν=0\operatorname{div}f^{\nu}=0, such that

  • u0νu0u_{0}^{\nu}\to u_{0} strongly in L2(𝕋2)L^{2}(\mathbb{T}^{2}),

  • fνff^{\nu}{\rightharpoonup}f weakly in L1((0,T);L2(𝕋2))L^{1}((0,T);L^{2}(\mathbb{T}^{2})),

and the corresponding solutions uνu^{\nu} of the Navier-Stokes equations (2.10)

  • uνuu^{\nu}{\rightharpoonup}u weakly-\ast in L((0,T);L2(𝕋2))L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})).

In this case, the sequence uνu^{\nu}, ν0\nu\to 0, is referred to as a physical realization of uu.

The class of physically realizable solutions was originally introduced in [CLNS2016] in the absence of forcing, and extended to the forced case in [LN2022]. In both of these papers sufficient conditions for physically realizable solutions to be energy conservative, or energy balanced, were obtained in terms of LpL^{p}-control of the vorticity. For the unforced case, a sharp characterization of energy conservation for physically realizable solutions was achieved in [LMPP2021a]. The goal of the present work is to carry out a programme similar to [LMPP2021a] in the forced case.

Definition 2.6.

Let uu be a physically realizable weak solution and consider uνuu^{\nu}\rightharpoonup u a physical realization. Let ωνcurl(uν)\omega^{\nu}\equiv\operatorname{curl}(u^{\nu}). We say the family {uν}ν\{u^{\nu}\}_{\nu} exhibits anomalous dissipation if

lim infν0+ν0tων(τ)Lx22𝑑τ>0.\liminf_{\nu\to 0^{+}}\nu\int_{0}^{t}\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau>0.
Remark 2.7.

Note that our definition of anomalous dissipation corresponds to what was defined as “dissipation anomaly” in [Cheskidov23]. Furthermore, what was defined as “anomalous dissipation” in [Cheskidov23] is what we refer to as inviscid dissipation.

We are now ready to state our main result.

Theorem 2.8.

Let uu be a physically realizable solution of the incompressible Euler equations (2.4) with divergence-free forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})) and initial data u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}). Let uνu^{\nu}, ν0\nu\to 0, be a physical realization of uu with forcing fνL2((0,T);L2(𝕋2))f^{\nu}\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})), divfν=0\operatorname{div}f^{\nu}=0, and initial data u0νL2(𝕋2)u^{\nu}_{0}\in L^{2}(\mathbb{T}^{2}). Assume that the convergence fνff^{\nu}\to f is strong in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Then the following assertions are equivalent:

  1. (1)

    uu is energy balanced,

  2. (2)

    the convergence uνuu^{\nu}\to u is strong in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})),

  3. (3)

    the convergence uνuu^{\nu}\to u is strong in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})).

Before delving into the proof of Theorem 2.8 we make a few remarks on the assumptions.

Remark 2.9.

Observe that, from the hypotheses of Theorem 2.8, we assume implicitly that the sequence of external forcings fνf^{\nu} is uniformly bounded in Lt2Lx2L^{2}_{t}L^{2}_{x} as ν0\nu\to 0. This assumption is related to vorticity estimates. Indeed, recall the vorticity formulation of the the Navier-Stokes equations (2.10):

tων+uνων=νΔων+gν,\displaystyle\partial_{t}\omega^{\nu}+u^{\nu}\cdot\nabla\omega^{\nu}=\nu\Delta\omega^{\nu}+g^{\nu}, (2.12)

with gν=curl(fν)g^{\nu}=\operatorname{curl}(f^{\nu}). Multiplying (2.12) by ων\omega^{\nu}, and integrating the forcing term by parts gives

ddt12ωνLx22\displaystyle\frac{d}{dt}\frac{1}{2}\|\omega^{\nu}\|_{L^{2}_{x}}^{2} =νωνLx22+𝕋2gνων𝑑x\displaystyle=-\nu\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}+\int_{\mathbb{T}^{2}}g^{\nu}\omega^{\nu}\,dx
=νωνLx22𝕋2fνωνdx.\displaystyle=-\nu\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}-\int_{\mathbb{T}^{2}}f^{\nu}\cdot\nabla^{\perp}\omega^{\nu}\,dx.

We estimate the last term from above by 12νωνLx22+(2ν)1fνLx22\frac{1}{2}\nu\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}+(2\nu)^{-1}\|f^{\nu}\|_{L^{2}_{x}}^{2}, thus obtaining the differential inequality

ddtωνLx22νωνLx22+1νfνLx22.\displaystyle\frac{d}{dt}\|\omega^{\nu}\|_{L^{2}_{x}}^{2}\leq-\nu\|\nabla\omega^{\nu}\|^{2}_{L^{2}_{x}}+\frac{1}{\nu}\|f^{\nu}\|_{L^{2}_{x}}^{2}. (2.13)

Neglecting the non-positive term and upon integration in time, we find

ων(t)L22ων(τ)L22+1ντtfνLx22𝑑s,\displaystyle\|\omega^{\nu}(t)\|^{2}_{L^{2}}\leq\|\omega^{\nu}(\tau)\|^{2}_{L^{2}}+\frac{1}{\nu}\int_{\tau}^{t}\|f^{\nu}\|^{2}_{L^{2}_{x}}\,ds, (2.14)

for τ(0,t]\tau\in(0,t]. We will see that, in order to obtain a bound on ων(t)L22\|\omega^{\nu}(t)\|^{2}_{L^{2}}, we require fνLt2Lx2f^{\nu}\in L^{2}_{t}L^{2}_{x}; see Lemma 3.7.

Remark 2.10.

In Theorem 2.8 it is furthermore assumed that fνf^{\nu} converges strongly to ff in Lt2Lx2L^{2}_{t}L^{2}_{x} as ν0\nu\to 0. This assumption ensures that

0Tfν,uνLx2𝑑t0Tf,uLx2𝑑t,\int_{0}^{T}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,dt\to\int_{0}^{T}\langle f,u\rangle_{L^{2}_{x}}\,dt, (2.15)

thus ruling out the failure of energy balance arising from the forcing term. If the convergence uνuu^{\nu}\to u is strong, then (2.15) holds even under the relaxed condition that fνff^{\nu}{\rightharpoonup}f only weakly in Lt2Lx2L^{2}_{t}L^{2}_{x}. In fact we will see in Proposition 3.12, that, assuming strong convergence of uνu^{\nu} to uu, energy balance of the limit uu follows under this weaker condition on fνf^{\nu}, ff.

In Example 2.11 below we will show that the hypothesis of strong convergence of fνf^{\nu} to ff is actually required for Theorem 2.8. More precisely we exhibit a physically realizable solution uu which is energy balanced, for which the forcings fνf^{\nu} of the physical realization converge weakly in Lt2Lx2L^{2}_{t}L^{2}_{x} to the forcing ff of the limit uu, yet uνu^{\nu} does not converge strongly to uu. Therefore the direction “energy balance \Rightarrow strong convergence” does not hold under the assumption that fνf^{\nu} converges only weakly to ff.

Example 2.11.

We claim that u0u\equiv 0 is a physically realizable solution, for which there exists a physical realization uνuu^{\nu}{\rightharpoonup}u with forcing fν0f^{\nu}{\rightharpoonup}0 in Lt2Lx2L^{2}_{t}L^{2}_{x}, but such that uνu^{\nu} does not converge strongly to uu in Lt2Lx2L^{2}_{t}L^{2}_{x}.

To this end, fix two non-zero functions γ,ϕCc((0,+))\gamma,\phi\in C^{\infty}_{c}((0,+\infty)) supported on [1/2,1][1/2,1]. Let uf0u\equiv f\equiv 0, and consider

uν(x,t)\displaystyle u^{\nu}(x,t) :=x|x|2sin(|x|ν1/3)ϕ(|x|)γ(t),\displaystyle:=\frac{x^{\perp}}{|x|^{2}}\sin\left(\frac{|x|}{\nu^{1/3}}\right)\phi(|x|)\gamma(t), (2.16)
fν\displaystyle f^{\nu} :=tuννΔuν.\displaystyle:=\partial_{t}u^{\nu}-\nu\Delta u^{\nu}. (2.17)

Of course, uu is an energy balanced solution of the incompressible Euler equations, with forcing ff (both vanishing identically). We next verify that uνu^{\nu} is a physical realization of uu: It is straightforward to show that uνu0u^{\nu}\overset{\ast}{\rightharpoonup}u\equiv 0 in LtLx2L^{\infty}_{t}L^{2}_{x}. Furthermore, we have uν(,0)0u^{\nu}({\,\cdot\,},0)\equiv 0, since γ(0)=0\gamma(0)=0 by assumption. In particular, this implies that u0νu0u^{\nu}_{0}\to u_{0} in Lx2L^{2}_{x}. Next, by construction of fν=tuννΔuνf^{\nu}=\partial_{t}u^{\nu}-\nu\Delta u^{\nu}, one readily verifies that

fν=x|x|2sin(|x|ν1/3)ϕ(|x|)γ(t)+O(ν1/3),f^{\nu}=\frac{x^{\perp}}{|x|^{2}}\sin\left(\frac{|x|}{\nu^{1/3}}\right)\phi(|x|)\gamma^{\prime}(t)+O(\nu^{1/3}),

where the O(ν1/3)O(\nu^{1/3})-term is with respect to LtLxL^{\infty}_{t}L^{\infty}_{x}. Thus, fν0f^{\nu}{\rightharpoonup}0 in Lt2Lx2L^{2}_{t}L^{2}_{x} as ν0\nu\to 0 by the oscillatory nature of the sine-factor. Finally, we point out that uνu^{\nu} is a solution of the Navier-Stokes equations (2.10) with forcing fνf^{\nu}: Indeed, any velocity field of the form U(x)=x|x|2sin(|x|ν1/3)ϕ(|x|)U(x)=\frac{x^{\perp}}{|x|^{2}}\sin\left(\frac{|x|}{\nu^{1/3}}\right)\phi(|x|) is a stationary solution of the Euler equations (see e.g. [MajdaBertozzi, Chapter 2.2.1, Example 2.1]), i.e. there exists a pressure PP such that UU+P=0U\cdot\nabla U+\nabla P=0. Thus, uν(x,t)=U(x)γ(t)u^{\nu}(x,t)=U(x)\gamma(t) solves the Navier–Stokes equations

tuν+uνuν+pν=νΔuν+fν,\partial_{t}u^{\nu}+u^{\nu}\cdot\nabla u^{\nu}+\nabla p^{\nu}=\nu\Delta u^{\nu}+f^{\nu},

where pν=P(x)γ(t)2p^{\nu}=P(x)\gamma(t)^{2}, and fν=tuννΔuνf^{\nu}=\partial_{t}u^{\nu}-\nu\Delta u^{\nu}. Lastly, we remark that, even though uνu0u^{\nu}{\rightharpoonup}u\equiv 0, it is readily verified from (2.16), that uνLx22↛0\|u^{\nu}\|_{L^{2}_{x}}^{2}\not\to 0 as ν0\nu\to 0, and hence uνu^{\nu} does not converge strongly to 0u0\equiv u in Lt2Lx2L^{2}_{t}L^{2}_{x}. This establishes the claim.

3. Proof of Theorem 2.8

In the present section, we will provide a detailed proof of Theorem 2.8. After recalling several useful a priori estimates on the Navier-Stokes equations (2.10) in Section 3.1, a proof of the direction “energy balance \Rightarrow strong convergence uνuu^{\nu}\to u” will be given in Proposition 3.8 in Section 3.2. A proof of the converse is given in Section 3.3, see Proposition 3.12. Finally, in Section 3.4, Proposition 3.15, we show that, under the assumptions of Theorem 2.8, the convergence uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x} can be improved to uniform-in-time convergence, i.e. uνuu^{\nu}\to u in C([0,T];Lx2)C([0,T];L^{2}_{x}).

3.1. A priori estimates

We collect several useful a priori estimates for solutions of the Navier-Stokes equations. Although these estimates are well-known, we include precise statements and proofs for completeness.

Lemma 3.1.

Let uνuu^{\nu}\rightharpoonup u be a physically realizable solution of the forced Euler equations with initial data u0νu0u_{0}^{\nu}\to u_{0} in L2L^{2}. Assume that supνfνLt2Lx2M\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\leq M. Then

u(t)Lx22(u0Lx22+tM)etM,t[0,T].\|u(t)\|_{L^{2}_{x}}^{2}\leq\left(\|u_{0}\|_{L^{2}_{x}}^{2}+\sqrt{t}\,M\right)e^{\sqrt{t}\,M},\quad\forall\,t\in[0,T].
Proof.

For ν>0\nu>0, we have

uν(t)Lx22\displaystyle\|u^{\nu}(t)\|_{L^{2}_{x}}^{2} =u0νLx222ν0tωνLx22𝑑τ+20tfν,uν𝑑τ\displaystyle=\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}-2\nu\int_{0}^{t}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau+2\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle\,d\tau
u0νLx22+20tfν(τ)Lx2uν(τ)Lx2𝑑τ\displaystyle\leq\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+2\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\|u^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau

We estimate the forcing term as follows,

20tfν(τ)Lx2uν(τ)Lx2𝑑τ\displaystyle 2\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\|u^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau 0tfν(τ)Lx2𝑑τ+0tfν(τ)Lx2uν(τ)Lx22𝑑τ,\displaystyle\leq\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau+\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\|u^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau,

to obtain

uν(t)Lx22\displaystyle\|u^{\nu}(t)\|_{L^{2}_{x}}^{2} u0νLx22+0tfν(τ)Lx2𝑑τ+0tfν(τ)Lx2uν(τ)Lx22𝑑τ.\displaystyle\leq\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau+\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\|u^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau.

The integral form of Gronwall’s inequality then implies that

uν(t)Lx22(u0νLx22+0tfν(τ)Lx2𝑑τ)e0tfνLx2𝑑τ.\|u^{\nu}(t)\|_{L^{2}_{x}}^{2}\leq\left(\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+\int_{0}^{t}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau\right)e^{\int_{0}^{t}\|f^{\nu}\|_{L^{2}_{x}}\,d\tau}. (3.1)

This provides a quantitative upper bound on uν(t)Lx2\|u^{\nu}(t)\|_{L^{2}_{x}}, provided that fνLt1Lx2f^{\nu}\in L^{1}_{t}L^{2}_{x}. The additional assumption supνfνLt2Lx2M\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\leq M implies an estimate which is uniform in ν\nu:

uν(t)Lx22(u0νLx22+tfνLt2Lx2)etfνLt2Lx2(u0νLx22+tM)etM.\displaystyle\begin{aligned} \|u^{\nu}(t)\|_{L^{2}_{x}}^{2}&\leq\left(\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+\sqrt{t}\,\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\right)e^{\sqrt{t}\,\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}}\\ &\leq\left(\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+\sqrt{t}\,M\right)e^{\sqrt{t}\,M}.\end{aligned} (3.2)

Since uνuu^{\nu}\rightharpoonup u weakly-\ast in LtLx2L^{\infty}_{t}L^{2}_{x}, and u0νu0u_{0}^{\nu}\to u_{0} converges strongly in L2L^{2}, the above estimate implies that

u(t)Lx22(u0Lx22+tM)etM.\|u(t)\|_{L^{2}_{x}}^{2}\leq\left(\|u_{0}\|_{L^{2}_{x}}^{2}+\sqrt{t}\,M\right)e^{\sqrt{t}\,M}.

Remark 3.2.

If fνff^{\nu}\to f strongly in Lt1Lx2L^{1}_{t}L^{2}_{x}, and if the convergence u0νu0u_{0}^{\nu}\to u_{0} is strong in Lx2L^{2}_{x}, then, using weak lower-semicontinuity of the L2L^{2}-norm on the estimate (3.1) derived in the proof of Lemma 3.1, we obtain

u(t)Lx22(u0Lx22+0tf(τ)Lx2𝑑τ)e0tf(τ)Lx2𝑑τ.\|u(t)\|^{2}_{L^{2}_{x}}\leq\left(\|u_{0}\|^{2}_{L^{2}_{x}}+\int_{0}^{t}\|f(\tau)\|_{L^{2}_{x}}\,d\tau\right)e^{\int_{0}^{t}\|f(\tau)\|_{L^{2}_{x}}\,d\tau}.

We next show that, under the Lt2Lx2L^{2}_{t}L^{2}_{x}-bound on the forcing, physically realizable solutions belong to C([0,T];w-L2(𝕋2))C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})).

Lemma 3.3.

Let uLtLx2u\in L^{\infty}_{t}L^{2}_{x} be a physically realizable weak solution of the incompressible Euler equations. Consider a physical realization uνuu^{\nu}{\rightharpoonup}u with external forcing fνff^{\nu}{\rightharpoonup}f, such that supνfνLt2Lx2<\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}<\infty. Let ,L2\langle\cdot,\cdot\rangle_{L^{2}} denote the L2L^{2}-inner product. Then, up to redefinition on a set of times of Lebesgue measure zero, it follows that, for every ϕL2(𝕋2)\phi\in L^{2}(\mathbb{T}^{2}), the function tu(t),ϕL2t\mapsto\langle u(t),\phi\rangle_{L^{2}} is continuous, i.e. uC([0,T];w-L2(𝕋2))u\in C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})).

Remark 3.4.

Given the result of Lemma 3.3, we will assume that any physically realizable solution uνuu^{\nu}{\rightharpoonup}u with supνfνLt2Lx2<\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}<\infty belongs to uC([0,T];w-L2(𝕋2))u\in C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})) without further comment.

Proof of Lemma 3.3.

Let M:=supνfνLt2Lx2M:=\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}. It follows from the strong convergence u0νu0u^{\nu}_{0}\to u_{0} and by Lemma 3.1 that we can bound

uνLx22(u0νLx2+TM)eTMC,\|u^{\nu}\|_{L^{2}_{x}}^{2}\leq\left(\|u^{\nu}_{0}\|_{L^{2}_{x}}+\sqrt{T}M\right)e^{\sqrt{T}M}\leq C,

by a constant CC that is independent of ν\nu. Therefore {uν}\{u^{\nu}\} is a bounded subset of L((0,T);L2(𝕋2))L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})).

Next we will bound tuν\partial_{t}u^{\nu}.

Let ϕ=ϕ(x)\phi=\phi(x) be a smooth test vector field. We write

tuν=uνuνpν+νΔuν+fν,\partial_{t}u^{\nu}=-u^{\nu}\cdot\nabla u^{\nu}-\nabla p^{\nu}+\nu\Delta u^{\nu}+f^{\nu},

we take the inner product with ϕ\phi and integrate on 𝕋2\mathbb{T}^{2} to find

𝕋2tuνϕ\displaystyle\int_{\mathbb{T}^{2}}\partial_{t}u^{\nu}\cdot\phi =𝕋2[div(uνuν)+pν+νΔuν+fν]ϕ.\displaystyle=-\int_{\mathbb{T}^{2}}\left[{\mathrm{div}}\left(u^{\nu}\otimes u^{\nu}\right)+\nabla p^{\nu}+\nu\Delta u^{\nu}+f^{\nu}\right]\cdot\phi.

Without loss of generality we may assume that div(ϕ)=0{\mathrm{div}}(\phi)=0, since we have div(uν)=0{\mathrm{div}}(u^{\nu})=0; thus the pressure term drops out. Transferring derivatives to ϕ\phi, and bounding the terms on the right, we find

|𝕋2tuνϕ|\displaystyle\left|\int_{\mathbb{T}^{2}}\partial_{t}u^{\nu}\cdot\phi\right| uνLx22DϕLx+νuνLx2ΔϕLx2+fνLx2ϕLx2,\displaystyle\leq\|u^{\nu}\|_{L^{2}_{x}}^{2}\|D\phi\|_{L^{\infty}_{x}}+\nu\|u^{\nu}\|_{L^{2}_{x}}\|\Delta\phi\|_{L^{2}_{x}}+\|f^{\nu}\|_{L^{2}_{x}}\|\phi\|_{L^{2}_{x}},

where DϕD\phi denotes the Jacobian matrix of ϕ\phi.

It follows from Sobolev embedding that, for sufficiently large LL, we have

|𝕋2tuνϕ|C(1+fνLx2)ϕHxL,\left|\int_{\mathbb{T}^{2}}\partial_{t}u^{\nu}\cdot\phi\right|\leq C\left(1+\|f^{\nu}\|_{L^{2}_{x}}\right)\|\phi\|_{H^{L}_{x}},

for all ϕHL(𝕋2)\phi\in H^{L}(\mathbb{T}^{2}), with a constant C>0C>0 that is independent of ν\nu. Interpreting the left-hand-side above as a duality pairing between HLH^{-L} and HLH^{L} yields

tuνHxL1+fνLx2.\|\partial_{t}u^{\nu}\|_{H^{-L}_{x}}\leq 1+\|f^{\nu}\|_{L^{2}_{x}}.

Therefore

tuνLt2HxLC(1+M),\|\partial_{t}u^{\nu}\|_{L^{2}_{t}H^{-L}_{x}}\leq C(1+M), (3.3)

so that {tuν}\{\partial_{t}u^{\nu}\} is a bounded subset of L2((0,T);HL(𝕋2))L^{2}((0,T);H^{-L}(\mathbb{T}^{2})).

It now follows immediately from the Aubin-Lions-Simon lemma, see for instance [BoyerFabrie2013, Theorem II.5.16], that {uν}ν\{u^{\nu}\}_{\nu} is a compact subset of C([0,T];w-L2(𝕋2))C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})), because the embedding L2w-L2L^{2}\subset w\text{-}L^{2} is compact. Given that uνuu^{\nu}\rightharpoonup u weak-\ast LtLx2L^{\infty}_{t}L^{2}_{x} it follows from the uniqueness of limits that uC([0,T];w-L2(𝕋2))u\in C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})), as desired.

Lemma 3.5.

Let uνuu^{\nu}\rightharpoonup u be a physically realizable solution of the forced Euler equations with initial data u0νu0u_{0}^{\nu}\to u_{0} strongly in Lx2L^{2}_{x}. Assume that supνfνLt2Lx2M\sup_{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\leq M. Then u=u(t)u=u(t) is right-continuous at t=0t=0, i.e.

limt0+u(t)u0Lx2=0.\lim_{\begin{subarray}{c}t\to 0^{+}\end{subarray}}\|u(t)-u_{0}\|_{L^{2}_{x}}=0.
Proof.

Since uC([0,T];w-Lx2)u\in C([0,T];w\text{-}L^{2}_{x}), and by lower-semicontinuity of the Lx2L^{2}_{x}-norm under weak limits and the upper bound from Lemma 3.1, it follows that:

u0Lx22\displaystyle\|u_{0}\|_{L^{2}_{x}}^{2} lim inft0u(t)Lx22lim supt0u(t)Lx22\displaystyle\leq\liminf_{t\to 0}\|u(t)\|_{L^{2}_{x}}^{2}\leq\limsup_{t\to 0}\|u(t)\|_{L^{2}_{x}}^{2}
lim supt0(u0Lx22+tM)etM=u0Lx22.\displaystyle\leq\limsup_{t\to 0}\left(\|u_{0}\|_{L^{2}_{x}}^{2}+\sqrt{t}\,M\right)e^{\sqrt{t}\,M}=\|u_{0}\|_{L^{2}_{x}}^{2}.

These lower and upper bounds imply that limt0u(t)Lx2=u0Lx2\lim_{t\to 0}\|u(t)\|_{L^{2}_{x}}=\|u_{0}\|_{L^{2}_{x}}. We also have u(t)u0u(t)\rightharpoonup u_{0} as t0t\to 0, from uC([0,T];w-Lx2)u\in C([0,T];w\text{-}L^{2}_{x}). Weak convergence together with convergence of the norms implies strong convergence, which concludes the proof. ∎

Remark 3.6.

It follows from Lemma 3.5, in particular, that

limδ0u(δ)Lx2=u0Lx2.\lim_{\delta\to 0}\|u(\delta)\|_{L^{2}_{x}}=\|u_{0}\|_{L^{2}_{x}}. (3.4)

Furthermore, since uLtLx2u\in L^{\infty}_{t}L^{2}_{x} and fLt2Lx2f\in L^{2}_{t}L^{2}_{x}, we have f,uLx2Lt1\langle f,u\rangle_{L^{2}_{x}}\in L^{1}_{t}. Therefore,

limδ0δtf,uLx2𝑑s=0tf,uLx2𝑑s.\lim_{\delta\to 0}\int_{\delta}^{t}\langle f,u\rangle_{L^{2}_{x}}\,ds=\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,ds. (3.5)

Following the classical terminology of turbulence theory, we refer to the square of the L2L^{2}-norm of vorticity as the enstrophy.

Lemma 3.7.

Let uνu^{\nu} be a solution of the forced Navier-Stokes equations with forcing fνLt2Lx2f^{\nu}\in L^{2}_{t}L^{2}_{x}. Then there exists a constant C=C(u0νLx2,fνLt2Lx2)>0C=C(\|u^{\nu}_{0}\|_{L^{2}_{x}},\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}})>0, such that

ων(t)Lx2Cνt.\|\omega^{\nu}(t)\|_{L^{2}_{x}}\leq\frac{C}{\sqrt{\nu t}}.
Proof.

From the energy balance for solutions of Navier-Stokes, we obtain

ν0TωνLx22𝑑t\displaystyle\nu\int_{0}^{T}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,dt =12u0Lx2212uν(T)Lx220Tfν,uνLx2𝑑t\displaystyle=\frac{1}{2}\|u_{0}\|^{2}_{L^{2}_{x}}-\frac{1}{2}\|u^{\nu}(T)\|_{L^{2}_{x}}^{2}-\int_{0}^{T}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,dt
12u0Lx22+12fνLt2Lx22+T2uνLtLx22.\displaystyle\leq\frac{1}{2}\|u_{0}\|^{2}_{L^{2}_{x}}+\frac{1}{2}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}^{2}+\frac{T}{2}\|u^{\nu}\|_{L^{\infty}_{t}L^{2}_{x}}^{2}.

We use the a priori estimate from Lemma 3.1 to deduce that all terms on the right-hand side are bounded by a positive constant C=C(u0νLx2,fνLt2Lx2)C=C(\|u^{\nu}_{0}\|_{L^{2}_{x}},\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}):

ν0TωνLx22𝑑τC(u0νLx2,fνLt2Lx2).\displaystyle\nu\int_{0}^{T}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau\leq C(\|u^{\nu}_{0}\|_{L^{2}_{x}},\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}). (3.6)

To obtain a pointwise estimate on ων(t)Lx2\|\omega^{\nu}(t)\|_{L^{2}_{x}}, we note that the vorticity equation implies the following upper bound on the enstrophy (cf. (2.14)), for τ[0,t]\tau\in[0,t]:

ων(t)Lx22\displaystyle\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2} ων(τ)Lx22+1ντtfν(s)Lx22𝑑s.\displaystyle\leq\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}+\frac{1}{\nu}\int_{\tau}^{t}\|f^{\nu}(s)\|_{L^{2}_{x}}^{2}\,ds.
ων(τ)Lx22+1νfνLt2Lx22.\displaystyle\leq\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}+\frac{1}{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}^{2}.

In particular, integrating in τ\tau from 0 to tt, the above estimate implies that

νtων(t)Lx22\displaystyle\nu t\,\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2} ν0tωνLx22𝑑τ+TfνLt2Lx22\displaystyle\leq\nu\int_{0}^{t}\|\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau+T\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}^{2}

By (3.6), the first term on the right-hand side is bounded by a constant depending only on the initial data and forcing. In particular, we conclude that there exists a positive constant C=C(T,u0νLx2,fνLt2Lx2)C=C(T,\|u^{\nu}_{0}\|_{L^{2}_{x}},\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}), such that

ων(t)Lx2Cνt.\|\omega^{\nu}(t)\|_{L^{2}_{x}}\leq\frac{C}{\sqrt{\nu t}}.

3.2. Energy balance implies strong convergence

Proposition 3.8.

Let uu be a physically realizable solution of the forced Euler equations (2.4), with initial data u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}) and forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Let uνu^{\nu} be a physical realization of uu, and assume additionally that the forcing fνf^{\nu} converges strongly to ff in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Then it holds that, if uu is energy balanced, then uνuu^{\nu}\to u strongly in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})).

Proof.

As uu is energy balanced it follows that

u(t)Lx22=u0Lx22+20tf,uLx2𝑑τ.\|u(t)\|_{L^{2}_{x}}^{2}=\|u_{0}\|_{L^{2}_{x}}^{2}+2\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau.

Furthermore, in view of the convergence uνuu^{\nu}{\rightharpoonup}u in weak-\ast L((0,T);L2(𝕋2))L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})) we have

uνu weakLt2Lx2,u^{\nu}{\rightharpoonup}u\text{ weak}\,L^{2}_{t}L^{2}_{x},

so that, by weak lower semicontinuity,

uLt2Lx2lim infν0+uνLt2Lx2.\|u\|_{L^{2}_{t}L^{2}_{x}}\leq\liminf_{\nu\to 0^{+}}\|u^{\nu}\|_{L^{2}_{t}L^{2}_{x}}.

Recall the energy inequality for the physical realization, valid for any 0τT0\leq\tau\leq T:

uν(τ)Lx22u0νLx22+20τfν,uνLx2𝑑s.\|u^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\leq\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+2\int_{0}^{\tau}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,ds.

The information above yields

0Tu(τ)Lx22𝑑τ\displaystyle\int_{0}^{T}\|u(\tau)\|_{L^{2}_{x}}^{2}\,d\tau\leq lim infν0+0Tuν(τ)Lx22𝑑τ\displaystyle\liminf_{\nu\to 0^{+}}\int_{0}^{T}\|u^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau (3.7)
\displaystyle\leq lim supν0+0Tuν(τ)Lx22𝑑τ\displaystyle\limsup_{\nu\to 0^{+}}\int_{0}^{T}\|u^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau (3.8)
\displaystyle\leq lim supν0+0T(u0νLx22+20τfν,uνLx2𝑑s)𝑑τ\displaystyle\limsup_{\nu\to 0^{+}}\int_{0}^{T}\left(\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}+2\int_{0}^{\tau}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,ds\right)\,d\tau (3.9)
=\displaystyle= 0T(u0Lx22+20τf,uLx2𝑑s)𝑑τ\displaystyle\int_{0}^{T}\left(\|u_{0}\|_{L^{2}_{x}}^{2}+2\int_{0}^{\tau}\langle f,u\rangle_{L^{2}_{x}}\,ds\right)\,d\tau (3.10)
=\displaystyle= 0Tu(τ)Lx22𝑑τ.\displaystyle\int_{0}^{T}\|u(\tau)\|_{L^{2}_{x}}^{2}\,d\tau. (3.11)

Therefore uνLt2Lx2uLt2Lx2\|u^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\to\|u\|_{L^{2}_{t}L^{2}_{x}} and, using again that convergence of norms and weak convergence implies strong convergence, the proof is concluded. ∎

Remark 3.9.

Notice that Proposition 3.8 is valid in any space dimension. In other words, after appropriately adjusting the definition of an energy balanced physically realizable solution, we may substitute 𝕋2\mathbb{T}^{2} in the statement by 𝕋d\mathbb{T}^{d} for any d2d\geq 2.

3.3. Strong convergence implies energy balance

The proof that strong convergence uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x} of the physical realization implies energy balance of the limit uu will be based on the following inequality for the enstrophy, for δ<t\delta<t, with δ,t[0,T]\delta,t\in[0,T], which follows from (2.14):

ων(t)Lx22ων(δ)Lx22νδtων2𝑑τ+1νδtfνLx22𝑑τ.\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2}\leq\|\omega^{\nu}(\delta)\|_{L^{2}_{x}}^{2}-\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}\|^{2}\,d\tau+\frac{1}{\nu}\int_{\delta}^{t}\|f^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau.

The basic idea, introduced in [CLNS2016], is to reduce the inequality above to a differential inequality for ωνLx22\|\omega^{\nu}\|_{L^{2}_{x}}^{2}, which in turn can be used to show that the energy dissipation term ν0TωνLx22𝑑τ0\nu\int_{0}^{T}\|\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau\to 0 as ν0\nu\to 0. To this end, the crucial ingredient is a good lower bound on ωνLx22\|\nabla\omega^{\nu}\|^{2}_{L^{2}_{x}} in terms of ωνLx22\|\omega^{\nu}\|_{L^{2}_{x}}^{2}, which is obtained in Lemma 3.10 below. Before we state this lemma we must recall the notation for structure functions introduced in [LMPP2021a].

If vLx2v\in L^{2}_{x} then the (L2L^{2}-based) structure function S2(v;r)S_{2}(v;r) is given by

S2(v;r)=(Br(0)|v(x+h)v(x)|2𝑑h)1/2.S_{2}(v;r)=\left(\fint_{B_{r}(0)}|v(x+h)-v(x)|^{2}\,dh\right)^{1/2}.

If, now, vLt2Lx2v\in L^{2}_{t}L^{2}_{x} then the time-integrated structure function S2T(v;r)S^{T}_{2}(v;r) is defined as

S2T(v;r)=(0T[S2(v(t);r)]2𝑑t)1/2.S^{T}_{2}(v;r)=\left(\int_{0}^{T}[S_{2}(v(t);r)]^{2}\,dt\right)^{1/2}.
Lemma 3.10.

Let {uν}ν>0\{u^{\nu}\}_{\nu>0} be a precompact family of divergence-free vector fields in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). There exists a monotonically increasing function σ:[0,)[0,)\sigma:[0,\infty)\to[0,\infty), such that limzσ(z)=\lim_{z\to\infty}\sigma(z)=\infty, and such that for each ν>0\nu>0 and δ,t[0,T]\delta,t\in[0,T], δ<t\delta<t, the vorticity ων=curluν\omega^{\nu}=\operatorname{curl}u^{\nu} satisfies the following inequality

(δtων(τ)Lx22𝑑τ)2σ(δtων(τ)Lx22𝑑τ)δtων(τ)Lx22𝑑τ.\left(\int_{\delta}^{t}\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\sigma\left(\int_{\delta}^{t}\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau\right)\leq\int_{\delta}^{t}\|\nabla\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau.
Proof.

This result is implicitly contained in the proof of [LMPP2021a, Thm. 2.11]. We outline the argument here. The first step is the following “interpolation-type” inequality valid for any ω=curluH1(𝕋2)\omega=\operatorname{curl}u\in H^{1}(\mathbb{T}^{2}), see [LMPP2021a, Lemma 2.6]: There exists an absolute constant C>0C>0, such that

ωLx2CrωLx2+2S2(u;r)r,r>0.\|\omega\|_{L^{2}_{x}}\leq Cr\|\nabla\omega\|_{L^{2}_{x}}+\frac{2S_{2}(u;r)}{r},\quad\forall r>0.

Applying this estimate to ων\omega^{\nu}, squaring terms and integrating in time, it follows that

δtωνLx22𝑑τr2δtωνLx22𝑑τ+[S2T(uν;r)]2r2,r>0,\int_{\delta}^{t}\|\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau\lesssim r^{2}\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau+\frac{[S_{2}^{T}(u^{\nu};r)]^{2}}{r^{2}},\quad\forall\,r>0,

with an implied constant independent of ν,r>0\nu,r>0 and independent of δ,t[0,T]\delta,t\in[0,T]. By [LMPP2021a, Prop. 2.10], the precompactness of {uν}Lt2Lx2\{u^{\nu}\}\subset L^{2}_{t}L^{2}_{x} implies that there exists a (monotonically increasing) modulus of continuity ϕ:[0,)[0,)\phi:[0,\infty)\to[0,\infty), with limr0ϕ(r)=0\lim_{r\to 0}\phi(r)=0, such that

supν>0[S2T(uν;r)]2[ϕ(r)]2.\sup_{\nu>0}[S^{T}_{2}(u^{\nu};r)]^{2}\leq[\phi(r)]^{2}.

Since the left-hand side is uniformly bounded from above by

supνS2T(uν;r)supν2uνLt2Lx2<,\sup_{\nu}S^{T}_{2}(u^{\nu};r)\leq\sup_{\nu}2\|u^{\nu}\|_{L^{2}_{t}L^{2}_{x}}<\infty,

we may assume, without loss of generality, that [ϕ(r)]2β[\phi(r)]^{2}\leq\beta for some β>0\beta>0, for all r0r\geq 0. Optimizing with respect to r>0r>0 so as to balance terms (cf. [LMPP2021a, eq. (2.13)]) results in

(δtωLx22𝑑τ)2C[ϕ(β[δtωνLx22𝑑τ]1/4)]2δtωνLx22𝑑τ,\left(\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\leq C\left[\phi\left(\beta\left[\textstyle\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau\right]^{-1/4}\right)\right]^{2}\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau,

with a constant C>0C>0, independent of ν\nu, rr, δ\delta and tt. To simplify notation, let us define a new modulus of continuity ϕ~:=Cϕ2\widetilde{\phi}:=C\phi^{2}, so that

(δtωLx22𝑑τ)2ϕ~(β[δtωνLx22𝑑τ]1/4)δtωνLx22𝑑τ.\left(\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\leq\widetilde{\phi}\left(\beta\left[\textstyle\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau\right]^{-1/4}\right)\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau. (3.12)

We next claim that the right-hand side of inequality (3.12) is bounded by a monotonically increasing function of z=δtωνLx22𝑑τz=\int_{\delta}^{t}\|\nabla\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau. Indeed, if we denote the right-hand side of (3.12) by f(z)=ϕ~(βz1/4)zf(z)=\widetilde{\phi}(\beta z^{-1/4})z, then f(z)=o(z)f(z)=o(z) as zz\to\infty, since ϕ~\widetilde{\phi} is a modulus of continuity and hence ϕ~(βz1/4)0\widetilde{\phi}(\beta z^{-1/4})\to 0 as zz\to\infty. In [LMPP2021a, Appendix C, Lemma C.1] it is shown that, for such f=f(z)f=f(z), there exists a dominating function FF, with F(z)f(z)F(z)\geq f(z), satisfying the following properties: F(z)=o(z)F(z)=o(z) as zz\to\infty, FF is invertible and its inverse F1F^{-1} is a monotonically increasing function which grows super-linearly. It is shown, furthermore, that F1F^{-1} can be written in the form F1(y)=yσ(y)F^{-1}(y)=y\sigma(\sqrt{y}), with σ\sigma a monotonically increasing function such that σ(y)\sigma(\sqrt{y})\to\infty as yy\to\infty. Using the notation introduced in the current paragraph, see (3.12), and estimating ff by FF we find

(δtωLx22𝑑τ)2F(δtωνLx22𝑑τ).\left(\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\leq F\left(\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau\right).

Applying F1F^{-1} to both sides of the inequality above yields

F1([δtωLx22𝑑τ]2)δtωνLx22𝑑τ.F^{-1}\left(\left[\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right]^{2}\right)\leq\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau.

Finally, writing F1(y)=yσ(y)F^{-1}(y)=y\sigma(\sqrt{y}) implies the desired upper bound

(δtωLx22𝑑τ)2σ(δtωLx22𝑑τ)δtωνLx22𝑑τ.\left(\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\sigma\left(\int_{\delta}^{t}\|\omega\|_{L^{2}_{x}}^{2}\,d\tau\right)\leq\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau.

The previous lemma will allow us to derive a differential inequality for the energy dissipation term ζν(t)=ν0tωνLx22𝑑τ\zeta_{\nu}(t)=\nu\int_{0}^{t}\|\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau, starting from the vorticity equation. The next lemma will then be used to show that ζν(T)0\zeta_{\nu}(T)\to 0, as ν0\nu\to 0, i.e. absence of anomalous dissipation, see Definition 2.6.

Lemma 3.11.

Let 0a<T0\leq a<T, M>0M>0 and let σ:++\sigma:\mathbb{R}_{+}\to\mathbb{R}_{+} be a continuous, monotonically increasing function, such that limzσ(z)=\lim_{z\to\infty}\sigma(z)=\infty. Let {ζν}ν>0W1,1([a,T])\{\zeta_{\nu}\}_{\nu>0}\subset W^{1,1}([a,T]) be a family of monotonically increasing functions. If ζν\zeta_{\nu} satisfies the differential inequality

dζνdtMζν2σ(ζνν), a.e. t[a,T],\displaystyle\frac{d\zeta_{\nu}}{dt}\leq M-\zeta_{\nu}^{2}\sigma\left(\frac{\zeta_{\nu}}{\nu}\right),\quad\text{ a.e. }t\in[a,T],

then lim supν0ζν(T)=0\limsup_{\nu\to 0}\zeta_{\nu}(T)=0.

Proof.

Since ζν\zeta_{\nu} is monotonically increasing and ζνW1,1\zeta_{\nu}\in W^{1,1} we have, for almost every t[a,T]t\in[a,T], that:

0ddtζν(t)M[ζν(t)]2σ(ζν(t)ν).0\leq\frac{d}{dt}\zeta_{\nu}(t)\leq M-[\zeta_{\nu}(t)]^{2}\sigma\left(\frac{\zeta_{\nu}(t)}{\nu}\right).

We note that the function on the right-hand side is continuous as a function of tt. Letting tTt\to T, we deduce that

[ζν(T)]2σ(ζν(T)/ν)M,ν>0.[\zeta_{\nu}(T)]^{2}\sigma(\zeta_{\nu}(T)/\nu)\leq M,\quad\forall\,\nu>0.

From the monotonicity of σ\sigma it follows that the function zz2σ(z/ν)z\mapsto z^{2}\sigma(z/\nu) is increasing.

Let us assume now, by contradiction, that lim supν0ζν(T)2ϵ0>0\limsup_{\nu\to 0}\zeta_{\nu}(T)\geq 2\epsilon_{0}>0, so that there exists a sequence νk0\nu_{k}\to 0 with ζνk(T)ϵ0\zeta_{\nu_{k}}(T)\geq\epsilon_{0}, for all kk\in\mathbb{N}. In this case we have, in particular, that

ϵ02σ(ϵ0νk)[ζνk(T)]2σ(ζνk(T)νk)M,\displaystyle\epsilon_{0}^{2}\sigma\left(\frac{\epsilon_{0}}{\nu_{k}}\right)\leq[\zeta_{\nu_{k}}(T)]^{2}\sigma\left(\frac{\zeta_{\nu_{k}}(T)}{\nu_{k}}\right)\leq M, (3.13)

for all kk\in\mathbb{N}, i.e. ϵ02σ(ϵ0/νk)\epsilon_{0}^{2}\sigma(\epsilon_{0}/\nu_{k}) is uniformly bounded. On the other hand, letting kk\to\infty, and using the assumption that limzσ(z)=\lim_{z\to\infty}\sigma(z)=\infty leads to

limkϵ02σ(ϵ0νk)=,\lim_{k\to\infty}\epsilon_{0}^{2}\sigma\left(\frac{\epsilon_{0}}{\nu_{k}}\right)=\infty,

in contradiction with (3.13). Therefore, we must have lim supν0ζν(T)=0\limsup_{\nu\to 0}\zeta_{\nu}(T)=0 as claimed. ∎

We are now in a position to prove the following result, which encodes the “strong convergence \Rightarrow energy balance” part of Theorem 2.8.

Proposition 3.12.

Let uu be a physically realizable solution of the Euler equations (2.4) with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνf^{\nu} converges weakly to ff in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). If uνuu^{\nu}\to u strongly in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})), then uu is energy balanced.

Proof.

Step 1: We begin by proving that for any δ>0\delta>0, the strong convergence assumption implies that the “δ\delta-truncated” energy dissipation term is vanishingly small:

νδTωνLx22𝑑τ0,as ν0.\nu\int_{\delta}^{T}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau\to 0,\quad\text{as }\nu\to 0.

To see this, first recall that, by Lemma 3.7, for any δ>0\delta>0, we have ων(δ)Lx2<\|\omega^{\nu}(\delta)\|_{L^{2}_{x}}<\infty. We may thus consider the following enstrophy equation, valid for any ν>0\nu>0, and δ(0,t]\delta\in(0,t]:

12ων(t)Lx22=12ων(δ)Lx22νδtωνLx22𝑑τ+δtων,gν𝑑τ,\frac{1}{2}\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2}=\frac{1}{2}\|\omega^{\nu}(\delta)\|_{L^{2}_{x}}^{2}-\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau+\int_{\delta}^{t}\langle\omega^{\nu},g^{\nu}\rangle\,d\tau,

with gν=curlfνLt2Hx1g^{\nu}=\operatorname{curl}f^{\nu}\in L^{2}_{t}H^{-1}_{x}. Integrating by parts the last term and using the Cauchy-Schwarz inequality we find

ων(t)Lx22ων(δ)Lx222νδtωνLx22𝑑τ+2δtωνLx2fνLx2𝑑τ.\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2}\leq\|\omega^{\nu}(\delta)\|_{L^{2}_{x}}^{2}-2\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau+2\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}\|f^{\nu}\|_{L^{2}_{x}}\,d\tau.

By Young’s inequality we have

ων(t)Lx22ων(δ)Lx22νδtωνLx22𝑑τ+1νδtfνLx22𝑑τ.\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2}\leq\|\omega^{\nu}(\delta)\|_{L^{2}_{x}}^{2}-\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau+\frac{1}{\nu}\int_{\delta}^{t}\|f^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau. (3.14)

In the following we keep δ>0\delta>0 fixed, and we will only consider values tδt\geq\delta. We note that, using again Lemma 3.7, there exists a constant C>0C>0, depending only on u0νLx2\|u_{0}^{\nu}\|_{L^{2}_{x}} and fνLt2Lx2\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}, such that

ων(δ)L22Cνδ.\displaystyle\|\omega^{\nu}(\delta)\|^{2}_{L^{2}}\leq\frac{C}{\nu\delta}. (3.15)

Furthermore we have

1νδtfνL22𝑑τ1νfνLt2Lx22.\displaystyle\frac{1}{\nu}\int_{\delta}^{t}\|f^{\nu}\|^{2}_{L^{2}}\,d\tau\leq\frac{1}{\nu}\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}^{2}. (3.16)

Using (3.15), (3.16) in (3.14), and since u0νLx2\|u_{0}^{\nu}\|_{L^{2}_{x}}, fνLt2Lx2\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}} are clearly uniformly bounded, we can find a constant M=M(δ)>0M=M(\delta)>0 such that

ων(t)Lx22\displaystyle\|\omega^{\nu}(t)\|_{L^{2}_{x}}^{2} MννδtωνLx22𝑑τ,\displaystyle\leq\frac{M}{\nu}-\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}\|_{L^{2}_{x}}^{2}\,d\tau, (3.17)

for all ν>0\nu>0. To estimate the gradient term, we note that Lemma 3.10 implies the existence of a monotonically increasing continuous, nonnegative, function σ=σ(z)\sigma=\sigma(z), with limzσ(z)=\lim_{z\to\infty}\sigma(z)=\infty, such that

(δtων(s)Lx22𝑑τ)2σ(δtων(s)Lx22𝑑τ)δtων(s)Lx22𝑑τ,\displaystyle\left(\int_{\delta}^{t}\|\omega^{\nu}(s)\|_{L^{2}_{x}}^{2}\,d\tau\right)^{2}\sigma\left(\int_{\delta}^{t}\|\omega^{\nu}(s)\|_{L^{2}_{x}}^{2}\,d\tau\right)\leq\int_{\delta}^{t}\|\nabla\omega^{\nu}(s)\|_{L^{2}_{x}}^{2}\,d\tau, (3.18)

for all ν>0\nu>0. We introduce the shorthand notation

ζν,δ(t):=νδtων(τ)Lx22𝑑τ,\displaystyle\zeta_{\nu,\delta}(t):=\nu\int_{\delta}^{t}\|\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau, (3.19)

for tδt\geq\delta, and we conclude that

νδtων(τ)Lx22𝑑τ\displaystyle-\nu\int_{\delta}^{t}\|\nabla\omega^{\nu}(\tau)\|_{L^{2}_{x}}^{2}\,d\tau 1ν[ζν,δ(t)]2σ(ζν,δ(t)ν).\displaystyle\leq-\frac{1}{\nu}[\zeta_{\nu,\delta}(t)]^{2}\sigma\left(\frac{\zeta_{\nu,\delta}(t)}{\nu}\right).

We also note that ζν,δW1,1([δ,T])\zeta_{\nu,\delta}\in W^{1,1}([\delta,T]), with ddtζν,δ(t)=νων(t)L22\frac{d}{dt}\zeta_{\nu,\delta}(t)=\nu\|\omega^{\nu}(t)\|^{2}_{L^{2}}. Multiplication of (3.17) by ν\nu and substitution of the above estimate therefore yields

ddtζν,δ(t)\displaystyle\frac{d}{dt}\zeta_{\nu,\delta}(t) M[ζν,δ(t)]2σ(ζν,δ(t)ν).\displaystyle\leq M-[\zeta_{\nu,\delta}(t)]^{2}\sigma\left(\frac{\zeta_{\nu,\delta}(t)}{\nu}\right). (3.20)

Recall that MM depends on δ\delta but is independent of ν\nu and tt. By construction, see (3.19), tζν,δ(t)t\mapsto\zeta_{\nu,\delta}(t) is monotonically increasing, and ζν,δW1,1([δ,T])\zeta_{\nu,\delta}\in W^{1,1}([\delta,T]). Since δ>0\delta>0 is fixed, and ζν,δ\zeta_{\nu,\delta} satisfies the differential inequality (3.20), we can use Lemma 3.11 to obtain that

lim supν0νδTων(τ)Lx22𝑑τlim supν0ζν,δ(T)=0,\limsup_{\nu\to 0}\nu\int_{\delta}^{T}\|\omega^{\nu}(\tau)\|^{2}_{L^{2}_{x}}\,d\tau\equiv\limsup_{\nu\to 0}\zeta_{\nu,\delta}(T)=0,

as desired. This concludes step 1 of the proof.

Step 2: We are now in a position to show that the physically realizable solution uu satisfies the energy balance equation (2.6). First we note that the strong convergence uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x} implies strong convergence of uν(t)Lx2u(t)Lx2\|u^{\nu}(t)\|_{L^{2}_{x}}\to\|u(t)\|_{L^{2}_{x}} in L2([0,T])L^{2}([0,T]). Passing to subsequences as needed, without relabeling, we may assume that

uν(t)Lx2u(t)Lx2,a.e. t[0,T].\displaystyle\|u^{\nu}(t)\|_{L^{2}_{x}}\to\|u(t)\|_{L^{2}_{x}},\quad\text{a.e. }t\in[0,T]. (3.21)

Next, we recall the energy identity for uνu^{\nu}, see also (2.11):

12uν(t)2=12uν(δ)2νδtωνLx22𝑑τ+δtfν,uνLx2𝑑τ,\frac{1}{2}\|u^{\nu}(t)\|^{2}=\frac{1}{2}\|u^{\nu}(\delta)\|^{2}-\nu\int_{\delta}^{t}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau+\int_{\delta}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau, (3.22)

valid for 0δtT0\leq\delta\leq t\leq T

By (3.21) we can choose δ>0\delta>0 outside of a set of measure 0 so that uν(δ)Lx2u(δ)Lx2\|u^{\nu}(\delta)\|_{L^{2}_{x}}\to\|u(\delta)\|_{L^{2}_{x}}. From the weak convergence fνff^{\nu}\rightharpoonup f in Lt2Lx2L^{2}_{t}L^{2}_{x}, and the strong convergence uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x}, it follows that

δtfν,uνLx2𝑑τδtf,uLx2𝑑τ,(ν0),\int_{\delta}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau\to\int_{\delta}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau,\quad(\nu\to 0),

for any t[δ,T]t\in[\delta,T]. By Step 1, it follows that

νδtωνLx22𝑑τ0,(ν0),\nu\int_{\delta}^{t}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau\to 0,\quad(\nu\to 0),

uniformly for all t[δ,T]t\in[\delta,T].

Thus, we conclude that for almost every t[δ,T]t\in[\delta,T], we have

12u(t)Lx22\displaystyle\frac{1}{2}\|u(t)\|^{2}_{L^{2}_{x}} =limν012uν(t)Lx22=12u(δ)Lx22+δtf,uLx2𝑑τ.\displaystyle=\lim_{\nu\to 0}\frac{1}{2}\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}=\frac{1}{2}\|u(\delta)\|^{2}_{L^{2}_{x}}+\int_{\delta}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau.

By Remark 3.6 it holds that δu(δ)Lx22\delta\mapsto\|u(\delta)\|_{L^{2}_{x}}^{2}, and δδtf,uLx2𝑑τ\delta\mapsto\int_{\delta}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau are both right-continuous at δ=0\delta=0, see (3.4) and (3.5). Letting δ0\delta\to 0, we therefore conclude that

12u(t)Lx22=12u0Lx22+0tf,uLx2𝑑τ,\frac{1}{2}\|u(t)\|^{2}_{L^{2}_{x}}=\frac{1}{2}\|u_{0}\|^{2}_{L^{2}_{x}}+\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau,

for almost all t[0,T]t\in[0,T], so that uu is energy balanced.

Redefining uu on a set of times of measure zero it follows that the energy balance holds for all t[0,T]t\in[0,T], see also Remark 2.3. This concludes the proof. ∎

Corollary 3.13.

Let uu be a physically realizable solution of the Euler equations (2.4) with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνf^{\nu} converges weakly to ff in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). If uνuu^{\nu}\to u strongly in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})), then

limν0ν0TωνLx22𝑑τ=0.\displaystyle\lim_{\nu\to 0}\,\nu\int_{0}^{T}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau=0. (3.23)
Proof.

For the sake of contradiction, assume that the claim is false. Then there exists ϵ0>0\epsilon_{0}>0, and a sequence νn0\nu_{n}\to 0, such that

νn0TωνnLx22𝑑τϵ0>0.\nu_{n}\int_{0}^{T}\|\omega^{\nu_{n}}\|_{L^{2}_{x}}^{2}\,d\tau\geq\epsilon_{0}>0.

In step 1 of the proof of Proposition 3.12, we have already shown that for any δ>0\delta>0, we have

limν0νδTωνLx22𝑑τ=0.\lim_{\nu\to 0}\,\nu\int_{\delta}^{T}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau=0.

Thus, the assumed lower bound on the energy dissipation would imply that

lim infnνn0δωνnLx22𝑑τϵ0>0,\displaystyle\liminf_{n\to\infty}\nu_{n}\int_{0}^{\delta}\|\omega^{\nu_{n}}\|_{L^{2}_{x}}^{2}\,d\tau\geq\epsilon_{0}>0, (3.24)

for any δ>0\delta>0. Our aim is to find δ>0\delta>0 for which this lower bound fails, thus reaching the desired contradiction.

By energy balance for Navier-Stokes (2.11), we have

νn0δωνnLx22𝑑τ=12uνn(δ)Lx2212u0νnLx220δfνn,uνnLx2𝑑τ.\nu_{n}\int_{0}^{\delta}\|\omega^{\nu_{n}}\|^{2}_{L^{2}_{x}}\,d\tau=\frac{1}{2}\|u^{\nu_{n}}(\delta)\|_{L^{2}_{x}}^{2}-\frac{1}{2}\|u^{\nu_{n}}_{0}\|_{L^{2}_{x}}^{2}-\int_{0}^{\delta}\langle f^{\nu_{n}},u^{\nu_{n}}\rangle_{L^{2}_{x}}\,d\tau.

Since uνnuu^{\nu_{n}}\to u strongly in Lt2Lx2L^{2}_{t}L^{2}_{x}, and since fνnff^{\nu_{n}}{\rightharpoonup}f weakly in Lt2Lx2L^{2}_{t}L^{2}_{x}, it follows that

limn0δfνn,uνnLx2𝑑τ=0δf,uLx2𝑑τ.\lim_{n\to\infty}\int_{0}^{\delta}\langle f^{\nu_{n}},u^{\nu_{n}}\rangle_{L^{2}_{x}}\,d\tau=\int_{0}^{\delta}\langle f,u\rangle_{L^{2}_{x}}\,d\tau.

Furthermore, since u0νnu(0)u_{0}^{\nu_{n}}\to u(0) strongly in Lx2L^{2}_{x} by assumption, we have

limn12u0νnLx22=12u(0)Lx22.\lim_{n\to\infty}\frac{1}{2}\|u^{\nu_{n}}_{0}\|_{L^{2}_{x}}^{2}=\frac{1}{2}\|u(0)\|_{L^{2}_{x}}^{2}.

Finally, as shown in the proof of Proposition 3.12 (cf. equation 3.21), after passing to a subsequence we can ensure that

limnuνn(δ)Lx2=u(δ)Lx2,a.e. δ[0,T].\lim_{n\to\infty}\|u^{\nu_{n}}(\delta)\|_{L^{2}_{x}}=\|u(\delta)\|_{L^{2}_{x}},\quad\text{a.e. }\delta\in[0,T].

Thus, for almost every δ>0\delta>0, we conclude that

limnνn0δωνnLx22𝑑τ=12u(δ)Lx2212u(0)Lx220δf,u𝑑τ.\lim_{n\to\infty}\nu_{n}\int_{0}^{\delta}\|\omega^{\nu_{n}}\|_{L^{2}_{x}}^{2}\,d\tau=\frac{1}{2}\|u(\delta)\|^{2}_{L^{2}_{x}}-\frac{1}{2}\|u(0)\|^{2}_{L^{2}_{x}}-\int_{0}^{\delta}\langle f,u\rangle\,d\tau.

From Remark 3.6, it follows that the right-hand side in the last display tends to 0 as δ0\delta\to 0. In particular, given ϵ0>0\epsilon_{0}>0, we can find δ>0\delta>0, outside of a set of measure zero, such that

limnνn0δωνnLx22𝑑τ\displaystyle\lim_{n\to\infty}\nu_{n}\int_{0}^{\delta}\|\omega^{\nu_{n}}\|_{L^{2}_{x}}^{2}\,d\tau =12u(δ)Lx2212u(0)Lx220δf,u𝑑τ\displaystyle=\frac{1}{2}\|u(\delta)\|^{2}_{L^{2}_{x}}-\frac{1}{2}\|u(0)\|^{2}_{L^{2}_{x}}-\int_{0}^{\delta}\langle f,u\rangle\,d\tau
ϵ02.\displaystyle\leq\frac{\epsilon_{0}}{2}. (3.25)

This upper bound (3.25) is clearly in contradiction with (3.24), completing the proof. ∎

Remark 3.14.

Corollary 3.13 corresponds to the statement that strong convergence of the physical realization implies no anomalous dissipation. In Proposition 3.12 we use the strong convergence and the absence of anomalous dissipation to establish that there is no inviscid dissipation in the associated physically realizable solution. Proposition 3.8 corresponds to the statement that absence of inviscid dissipation implies strong convergence, which in turn implies no anomalous dissipation.

It is natural to ask whether absence of anomalous dissipation alone implies no inviscid dissipation. Going back to (2.11) we see that

0tfν,uνLx2𝑑τν0tων(τ)Lx22𝑑τ\displaystyle\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau-\nu\int_{0}^{t}\|\omega^{\nu}(\tau)\|^{2}_{L^{2}_{x}}\,d\tau =12(uν(t)Lx22u0νLx22)\displaystyle=\frac{1}{2}\left(\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}-\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}\right) (3.26)
0tfν,uνLx2𝑑τ.\displaystyle\leq\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau.

If fνff^{\nu}\to f strongly in Lt2Lx2L^{2}_{t}L^{2}_{x} and uνuu^{\nu}\rightharpoonup u weakly in Lt2Lx2L^{2}_{t}L^{2}_{x} then the right-hand-side of (3.26) converges to

0tf,uLx2𝑑τ,\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau,

as ν0\nu\to 0. If, additionally, we assume absence of anomalous dissipation then the left-hand-side converges to the same term. It follows that

lim supν0(uν(t)Lx22u0νLx22)=0tf,uLx2𝑑τ.\limsup_{\nu\to 0}\left(\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}-\|u_{0}^{\nu}\|_{L^{2}_{x}}^{2}\right)=\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau. (3.27)

However, even though u0νLx2u0Lx2\|u_{0}^{\nu}\|_{L^{2}_{x}}\to\|u_{0}\|_{L^{2}_{x}} we cannot conclude that uu satisfies energy balance unless we have uν(t)Lx22u(t)Lx22\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}\to\|u(t)\|^{2}_{L^{2}_{x}} a.e. t[0,T]t\in[0,T]. As we have already argued, this is equivalent to requiring that uνu^{\nu} converge strongly to uu in Lt2Lx2L^{2}_{t}L^{2}_{x}. Therefore, absence of anomalous dissipation does not rule out inviscid dissipation. Nevertheless, it is interesting in its own right to study the absence of anomalous dissipation and we refer the reader to [DeRosaPark2024] for results in this direction.

3.4. Improvement from Lt2Lx2L^{2}_{t}L^{2}_{x} to CtLx2C_{t}L^{2}_{x} convergence

Putting together Propositions 3.8 and 3.12, we have shown that Theorem 2.8(1) is equivalent to Theorem 2.8(2). We will now complete the proof of Theorem 2.8 by showing that these two equivalent statements are also equivalent to Theorem 2.8(3).

Proposition 3.15.

Let uu be a physically realizable solution of the Euler equations (2.4) with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνf^{\nu} converges strongly to ff in Lt2Lx2L^{2}_{t}L^{2}_{x}.

If uνuu^{\nu}\to u strongly in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})) then uνuu^{\nu}\to u converges strongly in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})), or, equivalently, uu is energy balanced.

Conversely, if either of the following equivalent assertions holds,

  • uu is energy balanced, (cf. Theorem 2.8(1)),

  • uνu^{\nu} converges to uu in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})), (cf. Theorem 2.8(2)),

then the convergence uνuu^{\nu}\to u is actually strong in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})), (cf. Theorem 2.8(3)).

Proof.

It is immediate that uνuu^{\nu}\to u strongly in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})) implies uνuu^{\nu}\to u strongly in Lt2Lx2L^{2}_{t}L^{2}_{x}. By Propositions 3.8 and 3.12 this is equivalent to uu being energy balanced.

It remains to show that Theorem 2.8(1) and Theorem 2.8(2) together imply Theorem 2.8(3). Let uu be energy-balanced and uνuu^{\nu}\to u strongly in Lt2Lx2L^{2}_{t}L^{2}_{x}.

Step 1: We start with the observation that, since uu is energy-balanced, then uC([0,T];L2(𝕋2))u\in C([0,T];L^{2}(\mathbb{T}^{2})). To see this first, recall Remark 2.3, in which we observed that the function tu(t)L2t\mapsto\|u(t)\|_{L^{2}} is continuous. In addition in Lemma 3.3 we showed that uC([0,T];w-L2(𝕋2))u\in C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})). It is now immediate that continuity of norms together with continuity in time into Lx2L^{2}_{x} with the weak topology implies continuity in time into Lx2L^{2}_{x}.

We now come to the heart of the proof.

Step 2: We claim that uν(t)Lx22u(t)Lx22\|u^{\nu}(t)\|_{L^{2}_{x}}^{2}\to\|u(t)\|_{L^{2}_{x}}^{2} uniformly on [0,T][0,T].

To see this, we note that, since uu is energy balanced and from the energy balance identity (2.11) for solutions of the Navier-Stokes equations, we have for any t[0,T]t\in[0,T]:

u(t)Lx22uν(t)Lx22\displaystyle\|u(t)\|^{2}_{L^{2}_{x}}-\|u^{\nu}(t)\|^{2}_{L^{2}_{x}} ={u0Lx22+20tf,uLx2𝑑τ}\displaystyle=\left\{\|u_{0}\|^{2}_{L^{2}_{x}}+2\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau\right\}
{u0νLx222ν0tωνLx22𝑑τ+20tfν,uνLx2𝑑τ}\displaystyle\qquad-\left\{\|u^{\nu}_{0}\|^{2}_{L^{2}_{x}}-2\nu\int_{0}^{t}\|\omega^{\nu}\|^{2}_{L^{2}_{x}}\,d\tau+2\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau\right\}
=u0Lx22u0νLx22\displaystyle=\|u_{0}\|^{2}_{L^{2}_{x}}-\|u^{\nu}_{0}\|^{2}_{L^{2}_{x}}
+2{0tf,uLx2𝑑τ0tfν,uνLx2𝑑τ}\displaystyle\qquad+2\left\{\int_{0}^{t}\langle f,u\rangle_{L^{2}_{x}}\,d\tau-\int_{0}^{t}\langle f^{\nu},u^{\nu}\rangle_{L^{2}_{x}}\,d\tau\right\}
+2ν0tων2𝑑τ.\displaystyle\qquad+2\nu\int_{0}^{t}\|\omega^{\nu}\|^{2}\,d\tau.

Bounding the integal terms on the right-hand side, it follows that

|u(t)Lx22uν(t)Lx22|\displaystyle\Big{|}\|u(t)\|^{2}_{L^{2}_{x}}-\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}\Big{|} |u0Lx22u0νLx22|\displaystyle\leq\Big{|}\|u_{0}\|^{2}_{L^{2}_{x}}-\|u^{\nu}_{0}\|^{2}_{L^{2}_{x}}\Big{|}
+20Tf(τ)fν(τ)Lx2u(τ)Lx2𝑑τ\displaystyle\qquad+2\int_{0}^{T}\|f(\tau)-f^{\nu}(\tau)\|_{L^{2}_{x}}\|u(\tau)\|_{L^{2}_{x}}\,d\tau
+20Tfν(τ)Lx2u(τ)uν(τ)Lx2𝑑τ\displaystyle\qquad+2\int_{0}^{T}\|f^{\nu}(\tau)\|_{L^{2}_{x}}\|u(\tau)-u^{\nu}(\tau)\|_{L^{2}_{x}}\,d\tau
+2ν0Tων2𝑑τ\displaystyle\qquad+2\nu\int_{0}^{T}\|\omega^{\nu}\|^{2}\,d\tau
=:(I)+(II)+(III)+(IV).\displaystyle=:(I)+(II)+(III)+(IV).

Since the right-hand side is independent of tt, we conclude that

supt[0,T]|u(t)Lx22uν(t)Lx22|(I)+(II)+(III)+(IV).\sup_{t\in[0,T]}\Big{|}\|u(t)\|^{2}_{L^{2}_{x}}-\|u^{\nu}(t)\|^{2}_{L^{2}_{x}}\Big{|}\leq(I)+(II)+(III)+(IV).

It remains to show that the terms on the right-hand side converge to 0 as ν0\nu\to 0.

We note that (I)(I) converges to 0 as ν0\nu\to 0, since by assumption, u0νu0u^{\nu}_{0}\to u_{0} strongly in Lx2L^{2}_{x}, and hence u0νLx2u0Lx2\|u^{\nu}_{0}\|_{L^{2}_{x}}\to\|u_{0}\|_{L^{2}_{x}}. Next, we can bound

(II)ffνLt2Lx2uLt2Lx20,(II)\leq\|f-f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\|u\|_{L^{2}_{t}L^{2}_{x}}\to 0,

by assumption on fνff^{\nu}\to f in Lt2Lx2L^{2}_{t}L^{2}_{x}. Furthermore, since fνLt2Lx2M\|f^{\nu}\|_{L^{2}_{t}L^{2}_{x}}\leq M is uniformly bounded, and since we also assume that uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x}, we similarly conclude that (III)0(III)\to 0. Finally, since uνuu^{\nu}\to u in Lt2Lx2L^{2}_{t}L^{2}_{x}, the convergence (IV)0(IV)\to 0, as ν0\nu\to 0, follows from Corollary 3.13.

Step 3: We finally claim that

supt[0,T]u(t)uν(t)Lx20, as ν0.\sup_{t\in[0,T]}\|u(t)-u^{\nu}(t)\|_{L^{2}_{x}}\to 0,\quad\text{ as }\nu\to 0.

We argue by contradiction. If this is not the case, then there exists a convergent sequence tnt[0,T]t_{n}\to t\in[0,T] and a sequence νn0\nu_{n}\to 0, such that

u(tn)uνn(tn)Lx2ϵ0>0.\|u(t_{n})-u^{\nu_{n}}(t_{n})\|_{L^{2}_{x}}\geq\epsilon_{0}>0.

But, by the uniform convergence of uνn()Lx2u()Lx2\|u^{\nu_{n}}({\,\cdot\,})\|_{L^{2}_{x}}\to\|u({\,\cdot\,})\|_{L^{2}_{x}} in C([0,T])C([0,T]), it follows that uνn(tn)Lx2u(t)Lx2\|u^{\nu_{n}}(t_{n})\|_{L^{2}_{x}}\to\|u(t)\|_{L^{2}_{x}}. From the convergence uνnuu^{\nu_{n}}\to u in C([0,T];wLx2)C([0,T];w-L^{2}_{x}), see the proof of Lemma 3.3, it follows that uνn(tn)u(t)u^{\nu_{n}}(t_{n}){\rightharpoonup}u(t) in Lx2L^{2}_{x}. Since norm convergence and weak convergence imply strong convergence, we conclude that

limnuνn(tn)u(t)Lx2=0,\lim_{n\to\infty}\|u^{\nu_{n}}(t_{n})-u(t)\|_{L^{2}_{x}}=0,

Owing to the fact that uC([0,T];Lx2)u\in C([0,T];L^{2}_{x}) (cf. Step 0 of this proof), this leads to a contradiction:

0\displaystyle 0 <ϵ0lim supnu(tn)uνn(tn)Lx2\displaystyle<\epsilon_{0}\leq\limsup_{n\to\infty}\|u(t_{n})-u^{\nu_{n}}(t_{n})\|_{L^{2}_{x}}
lim supnuνn(tn)u(t)Lx2\displaystyle\leq\limsup_{n\to\infty}\|u^{\nu_{n}}(t_{n})-u(t)\|_{L^{2}_{x}}
+lim supnu(t)u(tn)Lx2=0.\displaystyle\qquad+\limsup_{n\to\infty}\|u(t)-u(t_{n})\|_{L^{2}_{x}}=0.

Thus, we must have that uνnuu^{\nu_{n}}\to u in C([0,T];Lx2)C([0,T];L^{2}_{x}). ∎

Putting together Propositions 3.8, 3.12 and 3.15 we have completed the proof of Theorem 2.8.

4. Examples

Theorem 2.8 provides necessary and sufficient conditions for a physically realizable solution uu to be energy balanced. In the present section, we focus on specific classes of initial data and forcing for which these conditions are satisfied. More precisely, we will study solutions whose vorticity belongs to rearrangement invariant spaces, including LpL^{p} (p>1p>1), the Orlicz spaces Llog(L)αL\log(L)^{\alpha} (α>1/2\alpha>1/2), and the (modified) Lorentz spaces L(1,q)L^{(1,q)} (1q21\leq q\leq 2).

Besides exhibiting instances to which Theorem 2.8 applies, the goals of this section are two-fold:

  • (i)

    we show how the present work extends the main result of [LN2022, Thm. 2.4], where energy balance was shown for physically realizable solutions with LpL^{p}-control on the vorticity for p>1p>1;

  • (ii)

    we fill a gap in the proof of [LMPP2021a, Corollary 2.13], where it was asserted that certain bounds on decreasing rearrangements are preserved by the solution operator of the Navier-Stokes equations, in the absence of an external force. A detailed proof of this assertion has, so far, been missing from the literature.

In Section 4.1 we connect our main results with the recent work [LN2022] on LpL^{p} vorticity control. Section 4.2 contains basic definitions on rearrangement invariant spaces; we also include the statements of our main results on a priori vorticity control for solutions of the Navier-Stokes equations. In Section 4.3 we extend the discussion of energy balance from the rearrangement-invariant LpL^{p}-spaces (p>1p>1) to more general rearrangement-invariant spaces, including Orlicz and Lorentz spaces.

4.1. Solutions with vorticity in LpL^{p}, p>1p>1

Proposition 3.12 implies the following stronger version of [LN2022, Thm. 2.4]:

Corollary 4.1.

Let uu be a physically realizable solution of the incompressible Euler equations (2.4) with external forcing fL1((0,T);L2(𝕋2))f\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})). Consider a physical realization uνu^{\nu} of uu, with viscosity ν>0\nu>0 and forcing fνf^{\nu}, as in Definition 2.5. Suppose, in addition, that for some p>1p>1:

  1. (1)

    ω0curlu0Lp(𝕋2)\omega_{0}\equiv\operatorname{curl}u_{0}\in L^{p}(\mathbb{T}^{2}),

  2. (2)

    ω0νcurlu0νω0\omega_{0}^{\nu}\equiv\operatorname{curl}u_{0}^{\nu}\to\omega_{0} strongly in Lp(𝕋2)L^{p}(\mathbb{T}^{2}),

  3. (3)

    gν=curlfνg^{\nu}=\operatorname{curl}f^{\nu} is bounded in L2((0,T);Lp(𝕋2))L^{2}((0,T);L^{p}(\mathbb{T}^{2})).

Then uu is an energy balanced weak solution.

Remark 4.2.

To derive the corresponding result in [LN2022, Thm. 2.4] the authors assume that gνg^{\nu} is uniformly bounded in L1((0,T);Lp(𝕋2))L((0,T);L2(𝕋2))L^{1}((0,T);L^{p}(\mathbb{T}^{2}))\cap L^{\infty}((0,T);L^{2}(\mathbb{T}^{2})). We note that, in the most relevant range 1<p<21<p<2, this is strictly stronger than assumption (3) of Corollary 4.1.

Proof.

We begin by observing that, from the hypothesis that gνg^{\nu} is bounded in Lt2LxpL^{2}_{t}L^{p}_{x}, we have, using elliptic regularity and the Poincaré inequality, that fνf^{\nu} is bounded in Lt2Wx1,pL^{2}_{t}W^{1,p}_{x}. Therefore, since W1,p(𝕋2)W^{1,p}(\mathbb{T}^{2}) is continuously embedded in L2(𝕋2)L^{2}(\mathbb{T}^{2}) for p1p\geq 1, we obtain that {fν}Lt2Lx2\{f^{\nu}\}\subset L^{2}_{t}L^{2}_{x} is uniformly bounded. Thus it is precompact in Lt2Lx2L^{2}_{t}L^{2}_{x} with the weak topology. Since fνff^{\nu}\rightharpoonup f in Lt1Lx2L^{1}_{t}L^{2}_{x} it follows that fνf^{\nu} converges weakly to ff in Lt2Lx2L^{2}_{t}L^{2}_{x} as well.

Next, we make a small adaptation in the proof of [LN2022, Lemma 3.1] to show that the assumptions of Corollary 4.1 imply that uνu^{\nu} converges strongly to uu in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})). For convenience of the reader, recall the estimate [LN2022, (3.3)]:

ων(t)LpC(ω0νLp+0Tgν(s)Lp𝑑s).\|\omega^{\nu}(t)\|_{L^{p}}\leq C\left(\|\omega^{\nu}_{0}\|_{L^{p}}+\int_{0}^{T}\|g^{\nu}(s)\|_{L^{p}}ds\right).

Using the Cauchy-Schwarz inequality in the integral term above we conclude that ων\omega^{\nu} is bounded in LtLxpL^{\infty}_{t}L^{p}_{x}, uniformly with respect to ν\nu. Thus, as in [LN2022, Lemma 3.1], we use elliptic regularity and the Poincaré inequality to conclude that uνu^{\nu} is bounded in LtWx1,pL^{\infty}_{t}W^{1,p}_{x}. We then further deduce from the PDE (2.10), that tuν\partial_{t}u^{\nu} is bounded in Lt2HxML^{2}_{t}H^{-M}_{x} for some MM\in\mathbb{N}. Since the embedding of W1,p(𝕋2)W^{1,p}(\mathbb{T}^{2}) into L2(𝕋2)L^{2}(\mathbb{T}^{2}) is compact for p>1p>1, it follows by the Aubin-Lions-Simons Lemma, see [BoyerFabrie2013, Theorem II.5.16], that {uν}\{u^{\nu}\} is compact in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})). Passing to subsequences as needed without relabeling we find uνvu^{\nu}\to v strongly in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})). By hypothesis we already know that uνuu^{\nu}\rightharpoonup u weak-\ast LtLx2L^{\infty}_{t}L^{2}_{x}. Therefore v=uv=u and the whole family converges strongly in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})).

The desired result is now an immediate corollary of Proposition 3.12. ∎

4.2. A priori estimates in rearrangement-invariant spaces

Aiming to generalize the result of the previous section, which pertains to solutions with LpL^{p}-vorticity control, we consider solutions with vorticity in more general rearrangement invariant spaces in the present section. In order to treat such spaces we will need the following definition.

Definition 4.3.

Let fL1(𝕋2)f\in L^{1}(\mathbb{T}^{2}). The rearrangement invariant maximal function for ff is

s(f):=sup{E|f(x)|dx|E𝕋2,E measurable, |E|=s},\mathcal{M}_{s}(f):=\sup\left\{\int_{E}|f(x)|\,dx\,\Big{|}\,E\subset\mathbb{T}^{2},E\text{ measurable, }|E|=s\right\},

defined for 0s|𝕋2|=(2π)20\leq s\leq|\mathbb{T}^{2}|=(2\pi)^{2}.

Remark 4.4.

We mention in passing that, since the torus 𝕋2\mathbb{T}^{2} with the Haar measure is a strongly resonant measure space, the function s(f)\mathcal{M}_{s}(f) defined above corresponds to sf(s)sf^{\ast\ast}(s), where ff^{\ast\ast} is the standard maximal function of the non-increasing rearrangement function ff^{\ast}; see [BennettSharpley, Chapter 2, Section 3] for additional information.

In this section, we prove the following a priori estimate for solutions of the Navier-Stokes equations:

Proposition 4.5.

Let uνL((0,T);L2(𝕋2)u^{\nu}\in L^{\infty}((0,T);L^{2}(\mathbb{T}^{2}) be the unique solution of the Navier-Stokes equations (2.10), with forcing fνL1((0,T);L2(𝕋2))f^{\nu}\in L^{1}((0,T);L^{2}(\mathbb{T}^{2})) and initial data u0νL2(𝕋2)u^{\nu}_{0}\in L^{2}(\mathbb{T}^{2}), both assumed to be divergence-free. Assume that ω0ν=curlu0νL1(𝕋2)\omega_{0}^{\nu}=\operatorname{curl}u^{\nu}_{0}\in L^{1}(\mathbb{T}^{2}), and gν=curlfνL1((0,T);L1(𝕋2))g^{\nu}=\operatorname{curl}f^{\nu}\in L^{1}((0,T);L^{1}(\mathbb{T}^{2})). Then for any t[0,T]t\in[0,T], we have the following a priori estimate:

s(ων(t))s(ω0ν)+0ts(gν(τ))𝑑τ.\displaystyle\mathcal{M}_{s}(\omega^{\nu}(t))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau. (4.1)

Our proof of this fact is based on operator splitting: The idea is to approximate the solution of the vorticity form of the Navier-Stokes equations by a composition of (small) time-steps for the forced heat equation and a transport PDE, respectively. The necessary a priori bounds for the heat and the transport equations are readily derived based on explicit solution formulae. If the operator splitting scheme converges to the solution operator of the Navier-Stokes equations in a suitable norm, then the necessary bounds for the Navier-Stokes equations can be deduced from a limiting argument.

One difficulty with this approach is the potential lack of smoothness of the underlying solution; in the absence of such smoothness, the operator splitting scheme is not known to converge. We circumvent this difficulty via an additional mollification argument: we establish the required bounds for solutions of the mollified system, and extend these a priori bounds to rough solutions by a limiting argument. The details of this argument are contained in Sections 4.2.14.2.3.

Section 4.2.1 discusses basic a priori estimates for the heat and transport PDEs involved in the operator splitting scheme. Section 4.2.2 defines the operator splitting approximant, and provides relevant estimates for this approximant. Section 4.2.3 combines these results to prove Proposition 4.5.

4.2.1. Split estimates

Given initial data ω0νC(𝕋2)\omega_{0}^{\nu}\in C^{\infty}(\mathbb{T}^{2}) and forcing gνC(𝕋2×[0,T])g^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]), standard well-posedness theory of the Navier-Stokes equations implies that there exists a unique smooth solution ωνC(𝕋2×[0,T])\omega^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) of the vorticity formulation of the Navier-Stokes equations. Given such a smooth solution, our aim is to derive estimates on s(ων)\mathcal{M}_{s}(\omega^{\nu}) through operator splitting. To this end, we decompose the vorticity equation:

tων=uνων(E)+νΔων+gν(H),\displaystyle\partial_{t}\omega^{\nu}=\underbrace{-u^{\nu}\cdot\nabla\omega^{\nu}}_{(E)}+\underbrace{\nu\Delta\omega^{\nu}+g^{\nu}}_{(H)}, (4.2)

according to the two terms (E)(E) and (H)(H) on the right-hand side.

In the following we assume that the smooth initial data ω0ν\omega_{0}^{\nu} and smooth forcing gνg^{\nu} are fixed. We denote by ωνC(𝕋2×[0,T])\omega^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) the corresponding solution, and we denote by uνC(𝕋2×[0,T])u^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) the divergence-free velocity field satisfying ων=curluν\omega^{\nu}=\operatorname{curl}u^{\nu}.


As indicated in (4.2), the vorticity equation can be split into a transport equation and a forced heat equation. We next introduce the corresponding solution operators.

Transport equation.

Given t0,t[0,T]t_{0},t\in[0,T], tt0t\geq t_{0}, we denote by β0E(t;t0)β0\beta_{0}\mapsto E(t;t_{0})\beta_{0} the solution operator associated with the transport PDE

tβ(t)+uν(t)β(t)=0,β(t0)=β0,\displaystyle\partial_{t}\beta(t)+u^{\nu}(t)\cdot\nabla\beta(t)=0,\qquad\beta(t_{0})=\beta_{0}, (E)

so that E(t;t0)β0:=β(t)E(t;t_{0})\beta_{0}:=\beta(t). In (E), uν(t)=uν(,t)u^{\nu}(t)=u^{\nu}({\,\cdot\,},t) denotes the velocity field obtained by solving the Navier-Stokes equations with the fixed initial vorticity ω0ν\omega^{\nu}_{0} and the fixed vorticity forcing gνg^{\nu}. The following proposition gives a simple a priori estimate on rearrangements under E(t;t0)E(t;t_{0}):

Proposition 4.6.

Let t0,t[0,T]t_{0},t\in[0,T], tt0t\geq t_{0}. If uνu^{\nu} is smooth, and if β(t)=E(t,t0)β0\beta(t)=E(t,t_{0})\beta_{0} is a solution of the transport PDE (E) with data βL1(𝕋2)\beta\in L^{1}(\mathbb{T}^{2}) then, for all s0s\geq 0, we have

s(β(t))=s(β0).\mathcal{M}_{s}(\beta(t))=\mathcal{M}_{s}(\beta_{0}).
Proof.

Let ϕ:𝕋2×[t0,t]𝕋2\phi:\mathbb{T}^{2}\times[t_{0},t]\to\mathbb{T}^{2} denote the flow-map of uνu^{\nu}, i.e.

{dϕdτ(x,τ)=uν(ϕ(x,τ),τ),τ[t0,t],ϕ(x,t0)=x,x𝕋2.\left\{\begin{array}[]{ll}\displaystyle{\frac{d\phi}{d\tau}}(x,\tau)=u^{\nu}(\phi(x,\tau),\tau),&\tau\in[t_{0},t],\\ &\\ \phi(x,t_{0})=x,&x\in\mathbb{T}^{2}.\end{array}\right.

We recall that β(t)=β0[ϕ(,t)]1\beta(t)=\beta_{0}\circ[\phi(\cdot,t)]^{-1}. Since uνu^{\nu} is divergence-free, ϕ(,t)\phi(\cdot,t) is measure-preserving, i.e. |ϕ(,t)1(E)|=|E||\phi(\cdot,t)^{-1}(E)|=|E|. In particular, it follows that

s(β(t))\displaystyle\mathcal{M}_{s}(\beta(t)) =sup|E|=sE|β(x,t)|𝑑x=sup|E|=sϕ(,t)1(E)|β0(x)|𝑑x\displaystyle=\sup_{|E|=s}\int_{E}|\beta(x,t)|\,dx=\sup_{|E|=s}\int_{\phi(\cdot,t)^{-1}(E)}|\beta_{0}(x)|\,dx
=sup|E|=sE|β0(x)|𝑑x=s(β0).\displaystyle=\sup_{|E|=s}\int_{E}|\beta_{0}(x)|\,dx=\mathcal{M}_{s}(\beta_{0}).


Heat equation.

Let t,t0[0,T]t,t_{0}\in[0,T] with tt0t\geq t_{0}. We denote by β0H(t;t0)β0\beta_{0}\mapsto H(t;t_{0})\beta_{0} the solution operator associated with the forced heat equation, i.e. we set H(t;t0)β0:=β(t)H(t;t_{0})\beta_{0}:=\beta(t), where β\beta solves

tβ(t)=νΔβ(t)+gν(t),β(t0)=β0.\displaystyle\partial_{t}\beta(t)=\nu\Delta\beta(t)+g^{\nu}(t),\qquad\beta(t_{0})=\beta_{0}. (H)

In the following, we derive simple a priori estimates on rearrangements under H(t;t0)H(t;t_{0}). We first consider the action of the heat kernel.

Proposition 4.7.

For t>0t>0, let Gt:𝕋2G_{t}:\mathbb{T}^{2}\to\mathbb{R} denote the ν\nu-heat kernel on the 2-dimensional torus; more precisely, let

Gt(x):=14πνtk2e(x2πk)24νt.G_{t}(x):=\frac{1}{4\pi\nu t}\sum_{k\in\mathbb{Z}^{2}}e^{-\frac{(x-2\pi k)^{2}}{4\nu t}}.

Then for any βC(𝕋2)\beta\in C^{\infty}(\mathbb{T}^{2}) and s0s\geq 0, we have

s(Gtβ)s(β).\mathcal{M}_{s}\left(G_{t}\ast\beta\right)\leq\mathcal{M}_{s}(\beta).
Proof.

We note that GtL1=1\|G_{t}\|_{L^{1}}=1, and hence

s(Gtβ)\displaystyle\mathcal{M}_{s}\left(G_{t}\ast\beta\right) =sup|E|=sE|Gtβ|𝑑x\displaystyle=\sup_{|E|=s}\int_{E}|G_{t}\ast\beta|\,dx
sup|E|=s𝕋21E(x)𝕋2Gt(y)|β(xy)|𝑑y𝑑x\displaystyle\leq\sup_{|E|=s}\int_{\mathbb{T}^{2}}1_{E}(x)\int_{\mathbb{T}^{2}}G_{t}(y)|\beta(x-y)|\,dy\,dx
=sup|E|=s𝕋2Gt(y)(𝕋21E(x)|β(xy)|𝑑x)𝑑y\displaystyle=\sup_{|E|=s}\int_{\mathbb{T}^{2}}G_{t}(y)\left(\int_{\mathbb{T}^{2}}1_{E}(x)|\beta(x-y)|\,dx\right)\,dy
sup|E|=sGtL1supy𝕋2E|β(xy)|𝑑x\displaystyle\leq\sup_{|E|=s}\|G_{t}\|_{L^{1}}\;\sup_{y\in\mathbb{T}^{2}}\int_{E}|\beta(x-y)|\,dx
=sup|E|=sE|β(y)|𝑑y\displaystyle=\sup_{|E|=s}\int_{E}|\beta(y)|\,dy
=s(β).\displaystyle=\mathcal{M}_{s}(\beta).

As a consequence of the last proposition, we obtain

Proposition 4.8.

Let t,t0[0,T]t,t_{0}\in[0,T] be given, such that tt0t\geq t_{0}. Let β0C(𝕋2)\beta_{0}\in C^{\infty}(\mathbb{T}^{2}), gνC(𝕋2×[0,T])g^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]), and let

β(t)=H(t;t0)β0C(𝕋2×[0,T]),\beta(t)=H(t;t_{0})\beta_{0}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]),

be the solution of the forced heat equation (H), i.e.

tβ(t)=νΔβ(t)+gν(t),β(t0)=β0.\partial_{t}\beta(t)=\nu\Delta\beta(t)+g^{\nu}(t),\qquad\beta(t_{0})=\beta_{0}.

Then for any s0s\geq 0:

s(β(t))s(β0)+t0ts(gν(τ))𝑑τ.\mathcal{M}_{s}(\beta(t))\leq\mathcal{M}_{s}(\beta_{0})+\int_{t_{0}}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.
Proof.

By Duhamel’s formula, the solution of the forced heat equation is given by

β(,t)=Gtt0β0+t0tGtτgν(,τ)𝑑τ.\beta({\,\cdot\,},t)=G_{t-t_{0}}\ast\beta_{0}+\int_{t_{0}}^{t}G_{t-\tau}\ast g^{\nu}({\,\cdot\,},\tau)\,d\tau.

Thus,

s(β(t))\displaystyle\mathcal{M}_{s}(\beta(t)) =sup|E|=sE|β(,t)|𝑑x\displaystyle=\sup_{|E|=s}\int_{E}|\beta({\,\cdot\,},t)|\,dx
sup|E|=sE|Gtt0β0|𝑑x+t0t(sup|E|=sE|Gtτgν(,τ)|𝑑x)𝑑τ\displaystyle\leq\sup_{|E|=s}\int_{E}|G_{t-t_{0}}\ast\beta_{0}|\,dx+\int_{t_{0}}^{t}\left(\sup_{|E|=s}\int_{E}|G_{t-\tau}\ast g^{\nu}({\,\cdot\,},\tau)|\,dx\right)\,d\tau
=s(Gtt0β0)+t0ts(Gtτgν(τ))𝑑τ.\displaystyle=\mathcal{M}_{s}(G_{t-t_{0}}\ast\beta_{0})+\int_{t_{0}}^{t}\mathcal{M}_{s}(G_{t-\tau}\ast g^{\nu}(\tau))\,d\tau.
s(β0)+t0ts(gν(τ))𝑑τ,\displaystyle\leq\mathcal{M}_{s}(\beta_{0})+\int_{t_{0}}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau,

where the last inequality follows from Proposition 4.7. ∎

4.2.2. Operator splitting approximation

In view of the definitions of E(t;t0)E(t;t_{0}), as the solution operator of the transport PDE (E) for tt0t\geq t_{0}, and H(t;t0)H(t;t_{0}), as the solution operator of the forced heat equation (H) for tt0t\geq t_{0}, we now define an “operator splitting” approximation of ων\omega^{\nu} recursively as follows: Fix a time-step Δt\Delta t, and for n0n\in\mathbb{N}_{0} define tn=nΔtt_{n}=n\Delta t. Given initial data ω0ν\omega^{\nu}_{0}, we set ω0ν,Δ(t0):=ω0ν\omega^{\nu,\Delta}_{0}(t_{0}):=\omega^{\nu}_{0} at t=t0=0t=t_{0}=0. Then, we recursively define

{ωn+1/2ν,Δ(t):=E(t;tn)ωnν,Δ(tn),ωn+1ν,Δ(t):=H(t;tn)ωn+1/2ν,Δ(t),for t(tn,tn+1].\displaystyle\left\{\begin{aligned} \omega^{\nu,\Delta}_{n+1/2}(t)&:=E(t;t_{n})\omega^{\nu,\Delta}_{n}(t_{n}),\\ \omega^{\nu,\Delta}_{n+1}(t)&:=H(t;t_{n})\omega^{\nu,\Delta}_{n+1/2}(t),\end{aligned}\right.\qquad\text{for }t\in(t_{n},t_{n+1}]. (4.3)

We finally define ων,Δ\omega^{\nu,\Delta} piecewise in time, by setting

ων,Δ(t):=ωnν,Δ(t),for t(tn1,tn].\displaystyle\omega^{\nu,\Delta}(t):=\omega^{\nu,\Delta}_{n}(t),\quad\text{for }t\in(t_{n-1},t_{n}]. (4.4)

Before providing formal motivation for (4.3) (cf. the next remark), we would like to point out that

ωn+1ν,Δ(t)=H(t;tn)ωn+1/2ν,Δ(t),\omega^{\nu,\Delta}_{n+1}(t)=H(t;t_{n})\omega^{\nu,\Delta}_{n+1/2}(t),

depends on time tt through both the solution operator H(t;tn)H(t;t_{n}) and additionally through the data ωn+1/2ν,Δ(t)\omega^{\nu,\Delta}_{n+1/2}(t) to which this solution operator is applied. To avoid confusion, we point out that we could have equivalently defined ωn+1ν,Δ(t)\omega^{\nu,\Delta}_{n+1}(t) in two steps, by first setting for τ1,τ2[tn,tn+1]\tau_{1},\tau_{2}\in[t_{n},t_{n+1}]:

h(τ1,τ2):=H(τ1;tn)ωn+1/2ν,Δ(τ2),\displaystyle h(\tau_{1},\tau_{2}):=H(\tau_{1};t_{n})\omega^{\nu,\Delta}_{n+1/2}(\tau_{2}), (4.5)

and then setting ωn+1ν,Δ(t)=h(t,t)\omega^{\nu,\Delta}_{n+1}(t)=h(t,t).

Remark 4.9.

To motivate the above definition, we note that upon formally expanding in time and neglecting terms O(|ttn|2)O(|t-t_{n}|^{2}), we have

ωn+1/2ν,Δ(t)=E(t;tn)ωnν,Δ(tn)ωnν,Δ(tn)[ttn]uν(tn)ωnν,Δ(tn),\displaystyle\begin{aligned} \omega^{\nu,\Delta}_{n+1/2}(t)&=E(t;t_{n})\omega^{\nu,\Delta}_{n}(t_{n})\\ &\approx\omega^{\nu,\Delta}_{n}(t_{n})-[t-t_{n}]\,u^{\nu}(t_{n})\cdot\nabla\omega^{\nu,\Delta}_{n}(t_{n}),\end{aligned} (4.6)

and

ωn+1ν,Δ(t)=H(t;tn)ωn+1/2ν,Δ(t)ωn+1/2ν,Δ(t)+[ttn]{gν(t)+νΔωn+1/2ν,Δ(t)}.\displaystyle\begin{aligned} \omega^{\nu,\Delta}_{n+1}(t)&=H(t;t_{n})\omega^{\nu,\Delta}_{n+1/2}(t)\\ &\approx\omega^{\nu,\Delta}_{n+1/2}(t)+[t-t_{n}]\left\{g^{\nu}(t)+\nu\Delta\omega^{\nu,\Delta}_{n+1/2}(t)\right\}.\end{aligned} (4.7)

Inserting (4.6) into (4.7), rearranging and retaining only lowest order terms in [ttn][t-t_{n}], we find

ωn+1ν,Δ(t)ωnν,Δ(tn)+[ttn]{uν(tn)ωnν,Δ(tn)+gν(t)+νΔωnν,Δ(tn)}.\omega^{\nu,\Delta}_{n+1}(t)\approx\omega^{\nu,\Delta}_{n}(t_{n})+[t-t_{n}]\left\{-u^{\nu}(t_{n})\cdot\nabla\omega^{\nu,\Delta}_{n}(t_{n})+g^{\nu}(t)+\nu\Delta\omega^{\nu,\Delta}_{n}(t_{n})\right\}.

Equivalently, upon rearranging and stating this equation in terms of ων,Δ(t)\omega^{\nu,\Delta}(t), we find for t[tn,tn+1]t\in[t_{n},t_{n+1}],

ων,Δ(t)ων,Δ(tn)ttn=uν(tn)ων,Δ(tn)+gν(t)+νΔων,Δ(tn).\frac{\omega^{\nu,\Delta}(t)-\omega^{\nu,\Delta}(t_{n})}{t-t_{n}}=-u^{\nu}(t_{n})\cdot\nabla\omega^{\nu,\Delta}(t_{n})+g^{\nu}(t)+\nu\Delta\omega^{\nu,\Delta}(t_{n}).

Expanding the terms in this equation once more around tt, we have

ων,Δ(t)ων,Δ(tn)ttn\displaystyle\frac{\omega^{\nu,\Delta}(t)-\omega^{\nu,\Delta}(t_{n})}{t-t_{n}} =tων,Δ(t)+O(Δt),\displaystyle=\partial_{t}\omega^{\nu,\Delta}(t)+O(\Delta t),
uν(tn)ων,Δ(tn)\displaystyle-u^{\nu}(t_{n})\cdot\nabla\omega^{\nu,\Delta}(t_{n}) =uν(t)ων,Δ(t)+O(Δt),\displaystyle=-u^{\nu}(t)\cdot\nabla\omega^{\nu,\Delta}(t)+O(\Delta t),
νΔων,Δ(tn)\displaystyle\nu\Delta\omega^{\nu,\Delta}(t_{n}) =νΔων,Δ(t)+O(Δt).\displaystyle=\nu\Delta\omega^{\nu,\Delta}(t)+O(\Delta t).

Thus, formally up to terms of order O(Δt)O(\Delta t), the function ων,Δ(t)\omega^{\nu,\Delta}(t) defined by (4.3) solves the equation,

tων,Δ(t)=uν(t)ων,Δ(t)+gν(t)+νΔων,Δ(t)+O(Δt).\partial_{t}\omega^{\nu,\Delta}(t)=-u^{\nu}(t)\cdot\nabla\omega^{\nu,\Delta}(t)+g^{\nu}(t)+\nu\Delta\omega^{\nu,\Delta}(t)+O(\Delta t).

This provides the formal justification for our definition of the splitting approximant ων,Δ\omega^{\nu,\Delta}. To make this precise, a detailed analysis of the O(Δt)O(\Delta t) correction term is required. The detailed derivation will be provided in Appendix A.

Combining Propositions 4.6 and 4.8 we obtain the following a priori control on rearrangements for the operator splitting approximant ων,Δ\omega^{\nu,\Delta}:

Lemma 4.10.

Let ω0νC(𝕋2)\omega^{\nu}_{0}\in C^{\infty}(\mathbb{T}^{2}) be initial data for the Navier-Stokes equations in vorticity formulation with forcing gνC(𝕋2×[0,T])g^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]). Let ων,Δ\omega^{\nu,\Delta} be the operator splitting approximation (4.3) for a given time-step Δt>0\Delta t>0. Then for any time t[0,T]t\in[0,T] and for any s0s\geq 0, we have the following estimate

s(ων,Δ(t))s(ω0ν)+0ts(gν(τ))𝑑τ.\displaystyle\mathcal{M}_{s}(\omega^{\nu,\Delta}(t))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau. (4.8)
Proof.

Given t[0,T]t\in[0,T], we can choose nn\in\mathbb{N}, such that t[tn,tn+1]t\in[t_{n},t_{n+1}]. By definition, we have ων,Δ(t)=ωnν,Δ(t)\omega^{\nu,\Delta}(t)=\omega^{\nu,\Delta}_{n}(t). It thus suffices to prove that for nn\in\mathbb{N}, we have

s(ωnν,Δ(t))s(ω0ν)+0ts(gν(τ))𝑑τ,t[tn,tn+1].\displaystyle\mathcal{M}_{s}(\omega^{\nu,\Delta}_{n}(t))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau,\quad\forall t\in[t_{n},t_{n+1}]. (4.9)

We prove (4.9) by induction on nn. The claim is trivially true for n=0n=0. For the induction step and fixed t[tn,tn+1]t\in[t_{n},t_{n+1}], we recall that ωn+1/2ν,Δ(t)=E(t;tn)ωnν,Δ(tn)\omega^{\nu,\Delta}_{n+1/2}(t)=E(t;t_{n}){\omega}_{n}^{\nu,\Delta}(t_{n}), and ωn+1ν,Δ(t)=H(t;tn)ωn+1/2ν,Δ(t){\omega}^{\nu,\Delta}_{n+1}(t)=H(t;t_{n})\omega^{\nu,\Delta}_{n+1/2}(t). Since the advecting velocity field uνu^{\nu} is smooth for solutions of 2D Navier-Stokes with smooth initial data and forcing, Proposition 4.8 implies that

s(ωn+1ν,Δ(t))\displaystyle\mathcal{M}_{s}({\omega}^{\nu,\Delta}_{n+1}(t)) =s(H(t;tn)ωn+1/2ν,Δ(t))\displaystyle=\mathcal{M}_{s}\left(H(t;t_{n})\omega_{n+1/2}^{\nu,\Delta}(t)\right)
s(ωn+1/2ν,Δ(t))+tnts(gν(τ))𝑑τ.\displaystyle\leq\mathcal{M}_{s}(\omega_{n+1/2}^{\nu,\Delta}(t))+\int_{t_{n}}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.

Furthermore, since ωn+1/2ν,Δ(t)=E(t;tn)ωnν,Δ(tn)\omega_{n+1/2}^{\nu,\Delta}(t)=E(t;t_{n})\omega_{n}^{\nu,\Delta}(t_{n}), by Proposition 4.6 we have

s(ωn+1/2ν,Δ(t))=s(ωnν,Δ(tn)).\mathcal{M}_{s}(\omega_{n+1/2}^{\nu,\Delta}(t))=\mathcal{M}_{s}({\omega}_{n}^{\nu,\Delta}(t_{n})).

By the induction hypothesis, we have

s(ωnν,Δ(tn))s(ω0ν)+0tns(gν(τ))𝑑τ.\mathcal{M}_{s}({\omega}_{n}^{\nu,\Delta}(t_{n}))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t_{n}}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.

Combining these estimates yields (4.9). ∎

The previous result provides a priori bounds on the operator splitting approximant ων,Δ\omega^{\nu,\Delta}. The next result shows that the operator splitting approximant ων,Δων\omega^{\nu,\Delta}\to\omega^{\nu} converges to the solution, provided that the underlying forcing and solution are sufficiently regular:

Proposition 4.11.

Let u0νC(𝕋2)u^{\nu}_{0}\in C^{\infty}(\mathbb{T}^{2}) be smooth divergence-free initial data for the incompressible Navier-Stokes equations (2.10). Assume that that the forcing fνC(𝕋2×[0,T])f^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) is smooth. Let uνC(𝕋2×[0,T])u^{\nu}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) be the unique smooth solution with this data. Then the operator splitting approximant ων,Δ\omega^{\nu,\Delta} defined by (4.3) converges to ων\omega^{\nu}; more precisely, we have

limΔt0ωνων,ΔLtLx2=0.\lim_{\Delta t\to 0}\|\omega^{\nu}-\omega^{\nu,\Delta}\|_{L^{\infty}_{t}L^{2}_{x}}=0.
Remark 4.12.

The last proposition implies in particular that ων,Δ(,t)ων(,t)\omega^{\nu,\Delta}({\,\cdot\,},t)\to\omega^{\nu}({\,\cdot\,},t) converges in Lx1L^{1}_{x} for any t[0,T]t\in[0,T].

Proof of Proposition 4.11.

This is a direct consequence of the convergence result for operator splitting applied to the forced advection-diffusion PDE tβ+Uβ=νΔβ+g\partial_{t}\beta+U\cdot\nabla\beta=\nu\Delta\beta+g, which we have included in Appendix A for completeness; more precisely, we invoke Proposition A.6 with β:=ων\beta:=\omega^{\nu}, β0:=ω0ν\beta_{0}:=\omega^{\nu}_{0}, U:=uνU:=u^{\nu}, g:=gν=curl(fν)g:=g^{\nu}=\operatorname{curl}(f^{\nu}). This yields the claim. We emphasize that the convergence rate in this Proposition depends on certain CkC^{k}-norms of ω0ν\omega_{0}^{\nu}, uνu^{\nu}, fνf^{\nu} and on ν>0\nu>0, and hence this result applies only to smooth solutions of the Navier-Stokes equations. ∎

Given the convergence of operator splitting, Proposition 4.11, we next aim to derive an a priori estimate for the rearrangement invariant vorticity maximal functions s(ων)\mathcal{M}_{s}(\omega^{\nu}), where ων\omega^{\nu} is a solution of the vorticity form of the Navier-Stokes equations. To this end, we will need the following simple lemma:

Lemma 4.13.

If ωΔω\omega^{\Delta}\to\omega converges in L1(𝕋2)L^{1}(\mathbb{T}^{2}), then for any s0s\geq 0, we have

s(ω)=limΔs(ωΔ).\mathcal{M}_{s}(\omega)=\lim_{\Delta}\mathcal{M}_{s}(\omega^{\Delta}).
Proof.

The convergence ωΔω\omega^{\Delta}\to\omega in L1L^{1} implies the convergence of the rearrangements, since (see e.g. Thm. 1.D of [talenti])

|s(ω)s(ωΔ)|ωωΔL1(𝕋2)0.|\mathcal{M}_{s}(\omega)-\mathcal{M}_{s}(\omega^{\Delta})|\leq\|\omega-\omega^{\Delta}\|_{L^{1}(\mathbb{T}^{2})}\to 0.

4.2.3. Proof of Proposition 4.5

Proof.

Based on the results of the last sections, we finally prove that for solutions uνL((0,T);L2(𝕋2))L2((0,T);H1(𝕋2))u^{\nu}\in L^{\infty}((0,T);L^{2}(\mathbb{T}^{2}))\cap L^{2}((0,T);H^{1}(\mathbb{T}^{2})) of the Navier-Stokes equations (2.10) with additional vorticity control ω0ν=curl(u0ν)L1(𝕋2)\omega_{0}^{\nu}=\operatorname{curl}(u^{\nu}_{0})\in L^{1}(\mathbb{T}^{2}) and gν=curl(fν)L1((0,T);L1(𝕋2))g^{\nu}=\operatorname{curl}(f^{\nu})\in L^{1}((0,T);L^{1}(\mathbb{T}^{2})), we have the following a priori bound on the vorticity maximal function

s(ων(t))s(ω0ν)+0ts(gν(τ))𝑑τ.\mathcal{M}_{s}(\omega^{\nu}(t))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.

We aim to deduce this estimate from the corresponding estimate for suitable operator splitting approximants, equation (4.9).

To this end, we denote by ω0ν,ϵ\omega^{\nu,\epsilon}_{0} the mollification of the initial data with a smooth mollifier (e.g. applying the heat kernel for time ϵ\epsilon), and we denote by gν,ϵg^{\nu,\epsilon} the mollification of gνg^{\nu} in both space and time (where we extend gνg^{\nu} by zero for t[0,T]t\notin[0,T]). Let ων,ϵ\omega^{\nu,\epsilon} denote the solution of the corresponding vorticity equation

tων,ϵ+uν,ϵων,ϵ=νΔων,ϵ+gν,ϵ,ων,ϵ(t=0)=ω0ν,ϵ.\partial_{t}\omega^{\nu,\epsilon}+u^{\nu,\epsilon}\cdot\nabla\omega^{\nu,\epsilon}=\nu\Delta\omega^{\nu,\epsilon}+g^{\nu,\epsilon},\quad\omega^{\nu,\epsilon}(t=0)=\omega^{\nu,\epsilon}_{0}.

Finally, let ων,ϵ,Δ\omega^{\nu,\epsilon,\Delta} denote the operator splitting approximant of ων,ϵ\omega^{\nu,\epsilon} for a time-step Δt\Delta t. By Lemma 4.10, we have

s(ων,ϵ,Δ(t))s(ω0ν,ϵ)+0ts(gν,ϵ(τ))𝑑τ.\mathcal{M}_{s}(\omega^{\nu,\epsilon,\Delta}(t))\leq\mathcal{M}_{s}(\omega^{\nu,\epsilon}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu,\epsilon}(\tau))\,d\tau.

It follows from Proposition 4.11 that ων,ϵ,Δων,ϵ\omega^{\nu,\epsilon,\Delta}{\to}\omega^{\nu,\epsilon} in LtLx1L^{\infty}_{t}L^{1}_{x} as Δt0\Delta t\to 0. Lemma 4.13 thus implies that

s(ων,ϵ(t))s(ω0ν,ϵ)+0ts(gν,ϵ(τ))𝑑τ.\mathcal{M}_{s}(\omega^{\nu,\epsilon}(t))\leq\mathcal{M}_{s}(\omega^{\nu,\epsilon}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu,\epsilon}(\tau))\,d\tau.

Next, we note that mollification decreases the value of the vorticity maximal function (cf. the proof of Proposition 4.7), so that

s(ων,ϵ(t))s(ω0ν,ϵ)+0ts(gν,ϵ(τ))𝑑τs(ω0ν)+0ts(gν(τ))𝑑τ\displaystyle\begin{aligned} \mathcal{M}_{s}(\omega^{\nu,\epsilon}(t))&\leq\mathcal{M}_{s}(\omega^{\nu,\epsilon}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu,\epsilon}(\tau))\,d\tau\\ &\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\end{aligned} (4.10)

is uniformly bounded for any ϵ>0\epsilon>0. Finally, we note that the last estimate implies that, for each t[0,T]t\in[0,T], the family {ων,ϵ(t)|ϵ(0,1]}{\left\{\omega^{\nu,\epsilon}(t)\,\middle|\,\epsilon\in(0,1]\right\}} is weakly compact L1(𝕋2)L^{1}(\mathbb{T}^{2}) by the Dunford-Pettis theorem. Furthermore, ων,ϵων\omega^{\nu,\epsilon}\to\omega^{\nu} in C([0,T];w-H1(𝕋2))C([0,T];w\text{-}H^{-1}(\mathbb{T}^{2})), since uν,ϵuνu^{\nu,\epsilon}\to u^{\nu} in C([0,T];w-L2(𝕋2))C([0,T];w\text{-}L^{2}(\mathbb{T}^{2})) by classical well-posedness of the Navier-Stokes equations in dimension two. Therefore, it follows that the only weak L1L^{1}-limit point of ων,ϵ(t)\omega^{\nu,\epsilon}(t) is ων(t)\omega^{\nu}(t), and hence ων,ϵων\omega^{\nu,\epsilon}\to\omega^{\nu} in C([0,T];w-L1(𝕋2))C([0,T];w\text{-}L^{1}(\mathbb{T}^{2})), as ϵ0\epsilon\to 0. It is easy to see that s()\mathcal{M}_{s}({\,\cdot\,}) is weakly-L1L^{1} lower semi-continuous so, using (4.10), this gives

s(ων(t))lim infϵ0s(ων,ϵ(t))s(ω0ν)+0ts(gν(τ))𝑑τ.\mathcal{M}_{s}(\omega^{\nu}(t))\leq\liminf_{\epsilon\to 0}\mathcal{M}_{s}(\omega^{\nu,\epsilon}(t))\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.

This concludes the proof. ∎

4.3. The Lorentz space L(1,q)(𝕋2)L^{(1,q)}(\mathbb{T}^{2})

For 1q<1\leq q<\infty, the rearrangement-invariant Lorentz space L(1,q)(𝕋2)L^{(1,q)}(\mathbb{T}^{2}) is defined as the space

L(1,q)(𝕋2)={ωL1(𝕋2)|ωL(1,q)<},L^{(1,q)}(\mathbb{T}^{2})={\left\{\omega\in L^{1}(\mathbb{T}^{2})\,\middle|\,\|\omega\|_{L^{(1,q)}}<\infty\right\}},

with norm

ωL(1,q)\displaystyle\|\omega\|_{L^{(1,q)}} =(0|𝕋2||s(ω)|qdss)1/q.\displaystyle=\left(\int_{0}^{|\mathbb{T}^{2}|}\left|\mathcal{M}_{s}(\omega)\right|^{q}\frac{ds}{s}\right)^{1/q}.
Remark 4.14.

We comment in passing that the spaces L(p,q)L^{(p,q)} discussed in [LNT, Section 2.3], see also [Lions, (4.39)], were defined as the space of functions ff such s1/pf(s)Lq(ds/s)s^{1/p}f^{\ast\ast}(s)\in L^{q}(ds/s). Note that, since s(f)=sf(s)\mathcal{M}_{s}(f)=sf^{\ast\ast}(s), this definition coincides with the one above for p=1p=1.

It is well-known that for 1q21\leq q\leq 2, we have a continuous embedding L(1,q)(𝕋2)H1(𝕋2)L^{(1,q)}(\mathbb{T}^{2}){\hookrightarrow}H^{-1}(\mathbb{T}^{2}). This embedding is compact for q<2q<2, see for example [LNT, Theorem 2.3]. For q=2q=2, the space L(1,2)(𝕋2)L^{(1,2)}(\mathbb{T}^{2}) is the largest rearrangement invariant space with a continuous (but not compact) embedding in H1(𝕋2)H^{-1}(\mathbb{T}^{2}) [Lions]. The following proposition, due to P.L. Lions, provides sufficient conditions for a family of functions in L(1,2)L^{(1,2)} to be precompact in H1H^{-1}, see [Lions, Lemma 4.1]:

Proposition 4.15 (P.L. Lions).

A family {ων}νL(1,2)(𝕋2)\{\omega^{\nu}\}_{\nu}\subset L^{(1,2)}(\mathbb{T}^{2}) is precompact in H1(𝕋2)H^{-1}(\mathbb{T}^{2}), if the following conditions hold:

  • (i)

    There exists C>0C>0, such that ωνL(1,2)(𝕋2)C\|\omega^{\nu}\|_{L^{(1,2)}(\mathbb{T}^{2})}\leq C uniformly in ν\nu,

  • (ii)

    we have uniform decay

    limδ0{supν0δ|s(ων)|2dss}=0\lim_{\delta\to 0}\left\{\sup_{\nu}\int_{0}^{\delta}|\mathcal{M}_{s}(\omega^{\nu})|^{2}\frac{ds}{s}\right\}=0

We recall that C([0,T];w-X)C([0,T];w\text{-}X) denotes the space of continuous functions in time with values in XX endowed with the topology of weak convergence. We also recall the following lemma from [LNT]:

Lemma 4.16 ([LNT, Lemma 2.3]).

Let XX be a reflexive, separable Banach space. Let {fn}\{f_{n}\} be a bounded sequence in C([0,T];X)C([0,T];X). Then {fn}\{f_{n}\} is precompact in C([0,T];X)C([0,T];X) if and only if the following two conditions hold:

  • (a)

    {fn}\{f_{n}\} is precompact in C([0,T];w-X)C([0,T];w\text{-}X),

  • (b)

    For any t[0,T]t\in[0,T] and for any sequence tntt_{n}\to t, we have that {fn(tn)}\{f_{n}(t_{n})\} is precompact in XX.

As a consequence of Lemma 4.16 and Proposition 4.15 we obtain:

Corollary 4.17.

Let {ων}ν\{\omega^{\nu}\}_{\nu} be a family of functions in C([0,T];H1(𝕋2))C([0,T];H^{-1}(\mathbb{T}^{2})) and suppose that {tων}ν\{\partial_{t}\omega^{\nu}\}_{\nu} is bounded in L2((0,T);HM(𝕋2))L^{2}((0,T);H^{-M}(\mathbb{T}^{2})) for some M>1M>1. Then {ων}ν\{\omega^{\nu}\}_{\nu} is precompact in C([0,T];H1(𝕋2))C([0,T];H^{-1}(\mathbb{T}^{2})), if the following conditions hold:

  1. (1)

    we have

    supνsupt[0,T]ων(t)L(1,2)(𝕋2)<,\sup_{\nu}\sup_{t\in[0,T]}\|\omega^{\nu}(t)\|_{L^{(1,2)}(\mathbb{T}^{2})}<\infty,
  2. (2)

    we have the following uniform decay in ν\nu and tt:

    limδ0{supνsupt[0,T]0δ|s(ων(t))|2dss}=0.\lim_{\delta\to 0}\left\{\sup_{\nu}\sup_{t\in[0,T]}\int_{0}^{\delta}|\mathcal{M}_{s}(\omega^{\nu}(t))|^{2}\frac{ds}{s}\right\}=0.
Proof.

We will check that conditions (a) and (b) in Lemma 4.16 hold true with X=H1(𝕋2)X=H^{-1}(\mathbb{T}^{2}).

From the boundedness in LtLx(1,2)L^{\infty}_{t}L^{(1,2)}_{x}, it follows that {ων}\{\omega^{\nu}\} is bounded in LtHx1L^{\infty}_{t}H^{-1}_{x} because Lx(1,2)Hx1L^{(1,2)}_{x}{\hookrightarrow}H^{-1}_{x}. Since we assumed that {tων}\{\partial_{t}\omega^{\nu}\} bounded in L2((0,T);HM(𝕋2))L^{2}((0,T);H^{-M}(\mathbb{T}^{2})) for some, possibly large, M>1M>1, it follows that {ων}\{\omega^{\nu}\} is equicontinuous from [0,T][0,T] into HxMH^{-M}_{x}. We may now use [Lions, Lemma C.1] to verify that (a) holds.

To check condition (b) let {tν}\{t_{\nu}\} be a convergent sequence in [0,T][0,T] and consider {ων(,tν)}\{\omega^{\nu}(\cdot,t_{\nu})\}. Then, hypotheses (1) and (2) imply that conditions (i) and (ii) of Proposition 4.15 hold true, which in turn implies that {ων(,tν)}\{\omega^{\nu}(\cdot,t_{\nu})\} is precompact in Hx1H^{-1}_{x}. This verifies condition (b) of Lemma 4.16.

The proof is complete. ∎

Theorem 4.18.

Let uu be a physically realizable solution of the Euler equations (2.4), with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})) and divergence-free initial data u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνff^{\nu}\rightharpoonup f in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Assume that ω0ν=curl(u0ν)L(1,2)(𝕋2)\omega_{0}^{\nu}=\operatorname{curl}(u^{\nu}_{0})\in L^{(1,2)}(\mathbb{T}^{2}) and gν=curl(fν)L1([0,T];L(1,2)(𝕋2))g^{\nu}=\operatorname{curl}(f^{\nu})\in L^{1}([0,T];L^{(1,2)}(\mathbb{T}^{2})), for all ν\nu. If we have

  • uniform bounds

    supνω0νL(1,2)C,supνgνLt1Lx(1,2)C,\displaystyle\sup_{\nu}\|\omega_{0}^{\nu}\|_{L^{(1,2)}}\leq C,\;\sup_{\nu}\|g^{\nu}\|_{L^{1}_{t}L^{(1,2)}_{x}}\leq C, (4.11)
  • uniform decay of the vorticity initial data

    limδ0{supν0δ|s(ω0ν)|2dss}=0,\displaystyle\lim_{\delta\to 0}\left\{\sup_{\nu}\int_{0}^{\delta}|\mathcal{M}_{s}(\omega_{0}^{\nu})|^{2}\frac{ds}{s}\right\}=0, (4.12)
  • and uniform decay of the time-averaged forcing

    limδ0{supν0δ|0Ts(gν(τ))𝑑τ|2dss}=0,\displaystyle\lim_{\delta\to 0}\left\{\sup_{\nu}\int_{0}^{\delta}\left|\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\right|^{2}\frac{ds}{s}\right\}=0, (4.13)

then the family {uν}ν\{u^{\nu}\}_{\nu} is relatively compact in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})), and uu is an energy balanced solution.

Proof.

From the definition of a physical realization, see Definition 2.5, we already have uνuu^{\nu}\rightharpoonup u in weak-\ast LtLx2L^{\infty}_{t}L^{2}_{x}.

We will show that, under the assumptions of Theorem 4.18, the vorticities {ων=curl(uν)}ν}\{\omega^{\nu}=\operatorname{curl}(u^{\nu})\}_{\nu}\} are relatively compact in C([0,T];H1(𝕋2))C([0,T];H^{-1}(\mathbb{T}^{2})). This in turn implies that {uν}ν\{u^{\nu}\}_{\nu} is relatively compact in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})) and hence, by Proposition 3.12, that uu is energy balanced.

We will use Corollary 4.17 to show that {ων}ν}\{\omega^{\nu}\}_{\nu}\} is relatively compact in C([0,T];H1(𝕋2))C([0,T];H^{-1}(\mathbb{T}^{2})). To this end we begin by observing that, for each fixed ν>0\nu>0, ωνC([0,T];H1(𝕋2))\omega^{\nu}\in C([0,T];H^{-1}(\mathbb{T}^{2})). Indeed, this is an immediate consequence of the fact that uνC([0,T];L2(𝕋2))u^{\nu}\in C([0,T];L^{2}(\mathbb{T}^{2})), as noted in Remark 2.4. Next, in the proof of Lemma 3.3 we deduced an estimate on tuν\partial_{t}u^{\nu}, (3.3), from which it follows that {tων}\{\partial_{t}\omega^{\nu}\} is bounded in Lt2HxML^{2}_{t}H^{-M}_{x} for some, possibly large, M>1M>1.

It remains to show that our assumptions imply uniform control on ων(t)L(1,2)\|\omega^{\nu}(t)\|_{L^{(1,2)}} and that supνsupt[0,T]0δ|s(ων(t))|2dss0\displaystyle{\sup_{\nu}\sup_{t\in[0,T]}\int_{0}^{\delta}|\mathcal{M}_{s}(\omega^{\nu}(t))|^{2}\frac{ds}{s}}\to 0 as δ0\delta\to 0. By Proposition 4.5, we have

s(ων(t))\displaystyle\mathcal{M}_{s}(\omega^{\nu}(t)) s(ω0ν)+0ts(gν(τ))𝑑τ.\displaystyle\leq\mathcal{M}_{s}(\omega^{\nu}_{0})+\int_{0}^{t}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau.

We can thus bound

ων,ϵ(t)L(1,2)\displaystyle\|\omega^{\nu,\epsilon}(t)\|_{L^{(1,2)}} =(0|𝕋2||s(ων,ϵ(t))|2dss)1/2\displaystyle=\left(\int_{0}^{|\mathbb{T}^{2}|}|\mathcal{M}_{s}(\omega^{\nu,\epsilon}(t))|^{2}\,\frac{ds}{s}\right)^{1/2}
(0|𝕋2||s(ω0ν)|2dss)1/2\displaystyle\leq\left(\int_{0}^{|\mathbb{T}^{2}|}|\mathcal{M}_{s}(\omega_{0}^{\nu})|^{2}\,\frac{ds}{s}\right)^{1/2}
+(0|𝕋2||0Ts(gν(τ))𝑑τ|2dss)1/2\displaystyle\qquad+\left(\int_{0}^{|\mathbb{T}^{2}|}\left|\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\right|^{2}\,\frac{ds}{s}\right)^{1/2}

Minkowski’s integral inequality applied to G(τ,s):=s(gν(τ))G(\tau,s):=\mathcal{M}_{s}(g^{\nu}(\tau)) implies that

[0|𝕋2|[0TG(τ,s)𝑑τ]2dss]1/20T[0|𝕋2|G(τ,s)2dss]1/2𝑑τ,\displaystyle\left[\int_{0}^{|\mathbb{T}^{2}|}\left[\int_{0}^{T}G(\tau,s)\,d\tau\right]^{2}\,\frac{ds}{s}\right]^{1/2}\leq\int_{0}^{T}\left[\int_{0}^{|\mathbb{T}^{2}|}G(\tau,s)^{2}\,\frac{ds}{s}\right]^{1/2}\,d\tau,

and hence

ων,ϵ(t)L(1,2)ω0νL(1,2)+0tgν(τ)L(1,2)𝑑τ.\|\omega^{\nu,\epsilon}(t)\|_{L^{(1,2)}}\leq\|\omega_{0}^{\nu}\|_{L^{(1,2)}}+\int_{0}^{t}\|g^{\nu}(\tau)\|_{L^{(1,2)}}\,d\tau.

By our assumptions on ω0ν\omega_{0}^{\nu} and gνg^{\nu}, the right-hand side is bounded uniformly in ν\nu and t[0,T]t\in[0,T]. This is the first condition of Corollary 4.17.

Next, replacing the upper integration bound |𝕋2||\mathbb{T}^{2}| by δ>0\delta>0, the same argument implies that

(0δ|s(ων,ϵ(t))|2dss)1/2\displaystyle\left(\int_{0}^{\delta}|\mathcal{M}_{s}(\omega^{\nu,\epsilon}(t))|^{2}\frac{ds}{s}\right)^{1/2} (0δ|s(ω0ν)|2dss)1/2\displaystyle\leq\left(\int_{0}^{\delta}|\mathcal{M}_{s}(\omega^{\nu}_{0})|^{2}\frac{ds}{s}\right)^{1/2}
+(0δ|0Ts(gν(τ))𝑑τ|2dss)1/2.\displaystyle\qquad+\left(\int_{0}^{\delta}\left|\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\right|^{2}\frac{ds}{s}\right)^{1/2}.

The right-hand side is independent of tt. Furthermore, by our assumptions, the right-hand side converges to zero as δ0\delta\to 0, uniformly in ν\nu. Thus, the family {ων}ν\{\omega^{\nu}\}_{\nu} satisfies the assumptions of Corollary 4.17, and hence {ωtν}ϵ,ν>0\{\omega_{t}^{\nu}\}_{\epsilon,\nu>0} is precompact in C([0,T];H1(𝕋2))C([0,T];H^{-1}(\mathbb{T}^{2})).

This concludes the proof. ∎

In particular, the last theorem can be invoked if ω0ν\omega^{\nu}_{0} and gνg^{\nu} satisfy uniform bounds in L(1,q)L^{(1,q)} for 1q<21\leq q<2, as shown next.

Corollary 4.19.

Let uu be a physically realizable solution of the Euler equations (2.4), with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})) and divergence-free initial data u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνff^{\nu}\rightharpoonup f in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Fix 1q<21\leq q<2, and assume that ω0ν=curl(u0ν)L(1,q)(𝕋2)\omega_{0}^{\nu}=\operatorname{curl}(u^{\nu}_{0})\in L^{(1,q)}(\mathbb{T}^{2}) and gν=curl(fν)L1([0,T];L(1,q)(𝕋2))g^{\nu}=\operatorname{curl}(f^{\nu})\in L^{1}([0,T];L^{(1,q)}(\mathbb{T}^{2})) for all ν\nu. If we have uniform bounds

supνω0νLx(1,q)C,supνgνLt1Lx(1,q)C,\sup_{\nu}\|\omega_{0}^{\nu}\|_{L^{(1,q)}_{x}}\leq C,\;\sup_{\nu}\|g^{\nu}\|_{L^{1}_{t}L^{(1,q)}_{x}}\leq C,

then {uν}ν\{u^{\nu}\}_{\nu} is relatively compact in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})), and uu is an energy balanced solution.

Proof.

We note that the a priori L(1,2)L^{(1,2)}-bounds (4.11) on ω0ν\omega^{\nu}_{0} and gνg^{\nu} follow immediately from the assumed L(1,q)L^{(1,q)}-bounds. In the following, we will show that the a priori L(1,q)L^{(1,q)}-bounds on ω0ν\omega_{0}^{\nu} and gνg^{\nu} for q<2q<2 imply the uniform decay conditions (4.12) and (4.13) of Theorem 4.18: Indeed, since ss(ω0ν)s\mapsto\mathcal{M}_{s}(\omega^{\nu}_{0}) is a monotonically increasing function, we have

0δ|s(ω0ν)|2dss|δ(ω0ν)|2q0δ|s(ω0ν)|qdss|δ(ω0ν)|2qω0νL(1,q)q,\int_{0}^{\delta}\left|\mathcal{M}_{s}(\omega^{\nu}_{0})\right|^{2}\,\frac{ds}{s}\leq\left|\mathcal{M}_{\delta}(\omega^{\nu}_{0})\right|^{2-q}\,\int_{0}^{\delta}\left|\mathcal{M}_{s}(\omega^{\nu}_{0})\right|^{q}\,\frac{ds}{s}\leq\left|\mathcal{M}_{\delta}(\omega^{\nu}_{0})\right|^{2-q}\|\omega^{\nu}_{0}\|^{q}_{L^{(1,q)}},

and

|δ(ω0ν)|qlog(|𝕋2|δ)δ|𝕋2||s(ω0ν)|qdssω0νL(1,q)q.\left|\mathcal{M}_{\delta}(\omega^{\nu}_{0})\right|^{q}\,\log\left(\frac{|\mathbb{T}^{2}|}{\delta}\right)\leq\int_{\delta}^{|\mathbb{T}^{2}|}\left|\mathcal{M}_{s}(\omega^{\nu}_{0})\right|^{q}\frac{ds}{s}\leq\|\omega^{\nu}_{0}\|_{L^{(1,q)}}^{q}.

Combining these estimates yields

0δ|s(ω0ν)|2dssω0νL(1,q)2|log(|𝕋2|/δ)|(2q)/q.\int_{0}^{\delta}\left|\mathcal{M}_{s}(\omega^{\nu}_{0})\right|^{2}\,\frac{ds}{s}\leq\frac{\|\omega^{\nu}_{0}\|_{L^{(1,q)}}^{2}}{\left|\,\log(|\mathbb{T}^{2}|/\delta)\right|^{(2-q)/q}}.

Given that supνω0νL(1,q)C\sup_{\nu}\|\omega^{\nu}_{0}\|_{L^{(1,q)}}\leq C, the last upper bound decays to zero uniformly in ν\nu, as δ0\delta\to 0. This shows the uniform decay condition (4.12).

Similarly, replacing s(ω0ν)\mathcal{M}_{s}(\omega^{\nu}_{0}) by 0Ts(gν(τ))𝑑τ\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau, we can show that

0δ|0Ts(gν(τ))𝑑τ|2dss\displaystyle\int_{0}^{\delta}\left|\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\right|^{2}\,\frac{ds}{s} [0|𝕋2||0Ts(gν(τ))𝑑τ|qdss]2/q|log(|𝕋2|/δ)|(2q)/q\displaystyle\leq\frac{\left[\int_{0}^{|\mathbb{T}^{2}|}\left|\int_{0}^{T}\mathcal{M}_{s}(g^{\nu}(\tau))\,d\tau\right|^{q}\,\frac{ds}{s}\right]^{2/q}}{\left|\,\log(|\mathbb{T}^{2}|/\delta)\right|^{(2-q)/q}}
[0T(0|𝕋2||s(gν(τ))|qdss)1/q𝑑τ]2|log(|𝕋2|/δ)|(2q)/q\displaystyle\leq\frac{\left[\int_{0}^{T}\left(\int_{0}^{|\mathbb{T}^{2}|}\left|\mathcal{M}_{s}(g^{\nu}(\tau))\right|^{q}\,\frac{ds}{s}\right)^{1/q}\,d\tau\right]^{2}}{\left|\,\log(|\mathbb{T}^{2}|/\delta)\right|^{(2-q)/q}}
=[0Tgν(τ)L(1,q)𝑑τ]2log(|𝕋2|/δ)(2q)/q,\displaystyle=\frac{\left[\int_{0}^{T}\|g^{\nu}(\tau)\|_{L^{(1,q)}}\,d\tau\right]^{2}}{\log(|\mathbb{T}^{2}|/\delta)^{(2-q)/q}},

where we have used Minkowski’s integral inequality to pass to the second line. The boundedness assumption on gνg^{\nu} now implies the uniform decay condition (4.13). The claim thus follows from Theorem 4.18. ∎

We close this subsection with a result for the Orlicz spaces L(logL)αL(\log L)^{\alpha}, with α>2\alpha>2. We briefly recall that these spaces are defined as

L(logL)α(𝕋2)={fL1(𝕋2)|𝕋2|f|[log+(|f|)]α𝑑x<}.L(\log L)^{\alpha}(\mathbb{T}^{2})=\{f\in L^{1}(\mathbb{T}^{2})\,\big{|}\,\int_{\mathbb{T}^{2}}|f|[\log^{+}(|f|)]^{\alpha}\,dx<\infty\}.

These are Banach spaces under the Luxembourg norm, given by

uL(logL)α:=inf{λ|𝕋2(|u(x)|λ)[log+(|u(x)|λ)]α𝑑x1},\|u\|_{L(\log L)^{\alpha}}:=\inf\left\{\lambda\,\bigg{|}\,\int_{\mathbb{T}^{2}}\left(\frac{|u(x)|}{\lambda}\right)\left[\log^{+}\left(\frac{|u(x)|}{\lambda}\right)\right]^{\alpha}\,dx\leq 1\right\},

see e.g. [BennettSharpley].

Corollary 4.20.

Let uu be a physically realizable solution of the Euler equations (2.4), with forcing fL2((0,T);L2(𝕋2))f\in L^{2}((0,T);L^{2}(\mathbb{T}^{2})) and divergence-free initial data u0L2(𝕋2)u_{0}\in L^{2}(\mathbb{T}^{2}). Let uνu^{\nu} be a physical realization of uu and assume that the forcing fνff^{\nu}\rightharpoonup f in L2((0,T);L2(𝕋2))L^{2}((0,T);L^{2}(\mathbb{T}^{2})). Fix α>12\alpha>\displaystyle{\frac{1}{2}}, and assume that ω0ν=curl(u0ν)L(logL)α(𝕋2)\omega_{0}^{\nu}=\operatorname{curl}(u^{\nu}_{0})\in L(\log L)^{\alpha}(\mathbb{T}^{2}) and gν=curl(fν)L1([0,T];L(logL)α(𝕋2))g^{\nu}=\operatorname{curl}(f^{\nu})\in L^{1}([0,T];L(\log L)^{\alpha}(\mathbb{T}^{2})) for all ν\nu. If we have uniform bounds

supνω0νL(logL)xαC,supνgνLt1L(logL)xαC,\sup_{\nu}\|\omega_{0}^{\nu}\|_{L(\log L)^{\alpha}_{x}}\leq C,\;\sup_{\nu}\|g^{\nu}\|_{L^{1}_{t}L(\log L)^{\alpha}_{x}}\leq C,

then {uν}ν\{u^{\nu}\}_{\nu} is relatively compact in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})), and uu is an energy balanced solution.

Proof.

The result is an immediate consequence of Corollary 4.19 and the embedding

L(logL)α(𝕋2)L(1,1/α)(𝕋2),L(\log L)^{\alpha}(\mathbb{T}^{2})\subset L^{(1,1/\alpha)}(\mathbb{T}^{2}),

which was established in [LNT, Lemma 2.1]. ∎

5. Conclusion

The primary objective of this work is to find suitable conditions on the regularity of the forcing to characterize those physically realizable weak solutions of the 2D Euler equations which are energy balanced, combining previous work by Lanthaler, Mishra and Parés-Pulido in [LMPP2021a] with work by Lopes Filho and Nussenzveig Lopes in [LN2022]. In [LMPP2021a], the authors establish, for flows without forcing, equivalence between strong convergence of the viscous approximation in Lt2Lx2L^{2}_{t}L^{2}_{x} and conservation of energy. We have proved the corresponding statement for flows with forcing, assuming {fν}\{f^{\nu}\} converges strongly in Lt2Lx2L^{2}_{t}L^{2}_{x}. In addition, we prove that these equivalent conditions are also equivalent to strong convergence of the viscous velocities in C([0,T];L2(𝕋2))C([0,T];L^{2}(\mathbb{T}^{2})). Furthermore, even if we consider only initial vorticities in LpL^{p}, 1<p<21<p<2, and the direction “strong convergence \Longrightarrow energy balance”, our conditions on the forcing are weaker than those in [LN2022]. Additionally, we provide examples of flows with vorticity in rearrangement-invariant spaces, for which the sufficient conditions are shown to hold. To this end, we develop novel a priori estimates for the rearrangement-invariant maximal vorticity function of solutions of the incompressible Navier-Stokes equations, and under minimal regularity assumptions. Our result fills a gap in the proof of [LMPP2021a, Corollary 2.13], where such bounds were asserted without proof. In short, we have proven energy conservation of physically realizable solutions with initial vorticity belonging to arbitrary rearrangement-invariant spaces with compact embedding in H1H^{-1}, and under natural assumptions on the external force. In particular, this sharpens and extends the results of [LN2022] in the forced setting, going beyond LpL^{p} vorticity control.

We describe a few avenues for future work. Firstly, the present work only considers deterministic forcing. Since investigations of driven turbulence often employ stochastic forcing, it would be of interest to extend the present work to the stochastic setting. Second, it would be interesting to consider the effects of a boundary, and investigate potential connections of the characterization of energy conservation in the present work with the Kato criterion [kato1984seminar]. Thirdly, combining the ideas of the present work with those of [ciampa2022energy] could provide a proof of energy conservation in the two-dimensional plane, going beyond vorticity in Llog(L)αL\log(L)^{\alpha} as assumed in [ciampa2022energy]. We plan to explore some of these research directions in the future.

Acknowledgments

The authors would like to thank Siddhartha Mishra for several insightful discussions which contributed significantly to this work. MCLF and HJNL are thankful for the generous hospitality of the “Forschungsinstitut für Mathematik” (FIM) at ETH Zürich during a research stay in Summer 2022. The research of SL is supported by Postdoc.Mobility grant P500PT-206737 from the Swiss National Science Foundation. MCLF was partially supported by CNPq, through grant # 304990/2022-1, and FAPERJ, through grant # E-26/201.209/2021. HJNL acknowledges the support of CNPq, through grant # 305309/2022-6, and of FAPERJ, through grant # E-26/201.027/2022.

Appendix A Operator splitting for advection-diffusion

In this appendix, our goal is to provide a self-contained proof of convergence for an operator splitting approximation of a simple advection-diffusion equation. We fix a smooth divergence-free vector field UC(𝕋2×[0,T])U\in C^{\infty}(\mathbb{T}^{2}\times[0,T]), a scalar forcing gC(𝕋2×[0,T])g\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) with zero mean 𝕋2g(x,t)𝑑x=0\int_{\mathbb{T}^{2}}g(x,t)\,dx=0 for all t[0,T]t\in[0,T], and we consider the following advection-diffusion PDE:

{tβ+Uβ=νΔβ+g,in 𝕋2×(0,T),β(t=0)=β0,on 𝕋2×{0}.\displaystyle\left\{\begin{array}[]{ll}\partial_{t}\beta+U\cdot\nabla\beta=\nu\Delta\beta+g,&\text{in }\mathbb{T}^{2}\times(0,T),\\ \beta(t=0)=\beta_{0},&\text{on }\mathbb{T}^{2}\times\{0\}.\end{array}\right. (A.3)

We will assume throughout that the initial data β0C(𝕋2)\beta_{0}\in C^{\infty}(\mathbb{T}^{2}), and β0\beta_{0} has zero mean, so that 𝕋2β0(x)𝑑x=0\int_{\mathbb{T}^{2}}\beta_{0}(x)\,dx=0.

Remark A.1.

While several convergence results for operator splitting approximations of advection-diffusion-reaction equations are available in the literature, those results mostly focus on higher-order (e.g. Strang-) operator splitting procedures. We have not been able to find a simple reference for the exact PDE (A.3) and the low-order splitting of relevance for the present work. Our goal in this appendix is therefore to provide a self-contained proof of convergence of a low-order operator splitting scheme for (A.3). From a numerical analysis point of view, the estimates in this appendix mostly follow standard procedure. We claim no originality, but include them here for completeness.

A.1. Operator splitting

A.1.1. Heat equation

Let t,t0[0,T]t,t_{0}\in[0,T] with tt0t\geq t_{0}. In the following, we will denote by β0H(t;t0)β0\beta_{0}\mapsto H(t;t_{0})\beta_{0} the solution operator of the following forced heat equation:

{tβ(t)=νΔβ(t)+g(t),β(,t0)=β0,\displaystyle\left\{\begin{aligned} &\partial_{t}\beta(t)=\nu\Delta\beta(t)+g(t),\\ &\beta({\,\cdot\,},t_{0})=\beta_{0},\end{aligned}\right. (A.4)

i.e. tH(t;t0)β0t\mapsto H(t;t_{0})\beta_{0} solves (A.4) over the time interval [t0,T][t_{0},T]. We recall that g(x,t)g(x,t) depends explicitly on tt, and therefore the solution operator H(t;t0)H(t;t_{0}) depends on both the initial time t0t_{0}, as well as on tt.

A.1.2. Transport equation

Similarly, we denote by β0E(t;t0)β\beta_{0}\mapsto E(t;t_{0})\beta the solution operator of the transport equation:

{tβ+Uβ=0,β(,t)=β0.\displaystyle\left\{\begin{aligned} &\partial_{t}\beta+U\cdot\nabla\beta=0,\\ &\beta({\,\cdot\,},t)=\beta_{0}.\end{aligned}\right. (A.5)

So that tE(t;t0)β0t\mapsto E(t;t_{0})\beta_{0} solves (A.5).

A.1.3. Splitting scheme

We expect that the solution operator S(t;t0)S(t;t_{0}) for the advection-diffusion equation (A.3) can be approximated by

S(t;t0)H(t;t0)E(t;t0)+O(Δt2),for |tt0|Δt,S(t;t_{0})\approx H(t;t_{0})E(t;t_{0})+O(\Delta t^{2}),\quad\text{for }|t-t_{0}|\leq\Delta t,

and hence, repeated application of HH and EE over small time-steps of size Δt\Delta t are expected to converge to the true solution operator as Δt0\Delta t\to 0. Our goal is to make this intuition precise, in the following. To this end, we will show that the corresponding “operator splitting approximant” βΔ\beta^{\Delta} converges as Δt0\Delta t\to 0:

To be more precise, given initial data β0\beta_{0} at t=0t=0, a finite time-horizon T>0T>0 and a small time-step Δt=TN>0\Delta t=\displaystyle{\frac{T}{N}}>0, we set tn:=nΔtt_{n}:=n\Delta t for n=0,1,,Nn=0,1,\ldots,N, and we define a sequence βnΔ(t){\beta}^{\Delta}_{n}(t) recursively, by

β0Δ(0):=β0,\displaystyle{\beta}^{\Delta}_{0}(0):=\beta_{0}, (A.6)

and

{βn+1/2Δ(t):=E(t;tn)βnΔ(tn),βn+1Δ(t):=H(t;tn)βn+1/2Δ(t),for t(tn,tn+1]\displaystyle\left\{\begin{aligned} {\beta}^{\Delta}_{n+1/2}(t)&:=E(t;t_{n}){\beta}^{\Delta}_{n}(t_{n}),\\ {\beta}^{\Delta}_{n+1}(t)&:=H(t;t_{n}){\beta}^{\Delta}_{n+1/2}(t),\end{aligned}\right.\qquad\text{for }t\in(t_{n},t_{n+1}]

We furthermore define βΔ{\beta}^{\Delta} for any t[0,T]t\in[0,T] by

βΔ(t):=βnΔ(t),for t(tn1,tn].\displaystyle{\beta}^{\Delta}(t):={\beta}^{\Delta}_{n}(t),\quad\text{for }t\in(t_{n-1},t_{n}]. (A.7)

A.2. Convergence

Let tβ(t)t\mapsto\beta(t) denote the exact solution to (A.3) with initial data β0C(𝕋2×[0,T])\beta_{0}\in C^{\infty}(\mathbb{T}^{2}\times[0,T]), satisfying 𝕋2β0(x)𝑑x=0\int_{\mathbb{T}^{2}}\beta_{0}(x)\,dx=0. Our first goal is to show that the operator splitting approximation converges βΔβ{\beta}^{\Delta}\to\beta as Δt0\Delta t\to 0.

To this end, we will need to derive several basic stability estimates. We start with the following lemma:

Lemma A.2.

Let σ0\sigma\geq 0. Let t,t0[0,T]t,t_{0}\in[0,T] with tt0t\geq t_{0}. Assume smooth forcing gC(𝕋2×[0,T])g\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) in the forced heat equation (A.4), and 𝕋2g(x)𝑑x=0\int_{\mathbb{T}^{2}}g(x)\,dx=0. If |tt0|Δt|t-t_{0}|\leq\Delta t, then

H(t;t0)β0Hxσβ0Hxσ+ΔtgLtHxσ.\displaystyle\|H(t;t_{0}){\beta_{0}}\|_{H^{\sigma}_{x}}\leq\|\beta_{0}\|_{H^{\sigma}_{x}}+\Delta t\|g\|_{L^{\infty}_{t}H^{\sigma}_{x}}. (A.8)

Furthermore, if kk\in\mathbb{N} and Δ\Delta is the Laplacian then

ΔkH(t;t0)β0Lx2Δkβ0Lx2+ΔtΔkgLtLx2.\displaystyle\|\Delta^{k}H(t;t_{0}){\beta_{0}}\|_{L^{2}_{x}}\leq\|\Delta^{k}\beta_{0}\|_{L^{2}_{x}}+\Delta t\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}. (A.9)
Proof.

Both (A.8) and (A.9) follow in a straightforward manner from the representation of the solution in terms of the heat kernel GtG_{t}:

H(t;t0)β0=Gtt0β0+t0tGtτg(,τ)𝑑τ.H(t;t_{0})\beta_{0}=G_{t-t_{0}}\ast\beta_{0}+\int_{t_{0}}^{t}G_{t-\tau}\ast g({\,\cdot\,},\tau)\,d\tau.

Lemma A.3 (Stability estimate for βΔ{\beta}^{\Delta}).

Assume that β0\beta_{0}, gg and UU are smooth functions, div(U)=0\operatorname{div}(U)=0 and β0\beta_{0}, gg have zero mean. Let βΔ\beta^{\Delta} be defined by (A.7). For any σ>0\sigma>0, there exists a constant C=C(T,σ,g,U,β0)>0C=C(T,\sigma,g,U,\beta_{0})>0, such that

supt[0,T]βΔ(t)HxσC,\sup_{t\in[0,T]}\|{\beta}^{\Delta}(t)\|_{H^{\sigma}_{x}}\leq C,

and

maxn:tn<Tsupt[tn,tn+1]E(t;tn)βnΔ(tn)HxσC,\max_{n:\,t_{n}<T}\sup_{t\in[t_{n},t_{n+1}]}\|E(t;t_{n})\beta^{\Delta}_{n}(t_{n})\|_{H^{\sigma}_{x}}\leq C,

uniformly as Δt0\Delta t\to 0.

Proof.

Recall that βΔ(t)\beta^{\Delta}(t) is defined piecewise as βΔ(t)=βnΔ(t)\beta^{\Delta}(t)=\beta^{\Delta}_{n}(t) over intervals t(tn1,tn]t\in(t_{n-1},t_{n}] with tn=nΔtt_{n}=n\Delta t for n=1,2,n=1,2,\dots. By definition, we have the recursion

βnΔ(t)=H(t;tn)E(t;tn)βn1Δ(tn).\beta^{\Delta}_{n}(t)=H(t;t_{n})E(t;t_{n})\beta^{\Delta}_{n-1}(t_{n}).

To provide a uniform bound on βΔ(t)Hxσ\|\beta^{\Delta}(t)\|_{H^{\sigma}_{x}} for all t[0,T]t\in[0,T], we will proceed in three steps: In a first step, we derive a general HxσH^{\sigma}_{x}-estimate over short time-intervals of length Δt\Delta t. Then, we use this estimate to bound the values βΔ(tn)Hxσ\|\beta^{\Delta}(t_{n})\|_{H^{\sigma}_{x}} at the interval endpoints. Finally, we combine the short-time estimate of Step 1 and the uniform bound on the βΔ(tn)Hxσ\|\beta^{\Delta}(t_{n})\|_{H^{\sigma}_{x}} from Step 2, to conclude that there exists a uniform bound on βΔ(t)Hxσ\|\beta^{\Delta}(t)\|_{H^{\sigma}_{x}} for all t[0,T]t\in[0,T].

Step 1: (short-time estimate) We begin by claiming that, for any σ>0\sigma>0, H(t;t0)E(t;t0)β0Hxσ\|H(t;t_{0})E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}} is bounded on short time intervals |tt0|<Δt|t-t_{0}|<\Delta t in terms of β0Hσ\|\beta_{0}\|_{H^{\sigma}} and Δt\Delta t. By interpolation, it suffices to prove the claim for σ=2k\sigma=2k, where kk\in\mathbb{N}. We first recall that for any t0tt_{0}\leq t, and β0C\beta_{0}\in C^{\infty} with zero-mean, we have (cf. Lemma A.2), (A.8):

H(t;t0)E(t;t0)β0HxσE(t;t0)β0Hxσ+ΔtgLtHxσ.\displaystyle\|H(t;t_{0})E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}}\leq\|E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}}+\Delta t\|g\|_{L^{\infty}_{t}H^{\sigma}_{x}}. (A.10)

Let us denote β~(t):=E(t;t0)β0\widetilde{\beta}(t):=E(t;t_{0})\beta_{0}. To estimate β~Hxσ=E(t;t0)β0Hxσ\|\widetilde{\beta}\|_{H^{\sigma}_{x}}=\|E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}} for σ=2k\sigma=2k, we multiply (A.5) by Δ2kβ~\Delta^{2k}\widetilde{\beta}, with Δ\Delta the Laplacian, to find

ddt12Δkβ~Lx22\displaystyle\frac{d}{dt}\frac{1}{2}\|\Delta^{k}\widetilde{\beta}\|_{L^{2}_{x}}^{2} |Uβ~,Δ2kβ~Lx2|\displaystyle\leq\left|\langle U\cdot\nabla\widetilde{\beta},\Delta^{2k}\widetilde{\beta}\rangle_{L^{2}_{x}}\right|
=|Δk(Uβ~),Δkβ~Lx2|\displaystyle=\left|\langle\Delta^{k}\left(U\cdot\nabla\widetilde{\beta}\right),\Delta^{k}\widetilde{\beta}\rangle_{L^{2}_{x}}\right|
|U(Δkβ~),Δkβ~Lx2|+CkUWx2k,β~Hx2k2,\displaystyle\leq\left|\langle U\cdot\nabla(\Delta^{k}\widetilde{\beta}),\Delta^{k}\widetilde{\beta}\rangle_{L^{2}_{x}}\right|+C_{k}\|U\|_{W^{2k,\infty}_{x}}\|\widetilde{\beta}\|_{H^{2k}_{x}}^{2},

with a constant CkC_{k} depending only on kk. Since UU is divergence-free, the first term vanishes on account of the periodic boundary conditions,

|U(Δkβ~),Δkβ~Lx2|=|𝕋2div(12[Δkβ~]2U)|=0.\left|\langle U\cdot\nabla(\Delta^{k}\widetilde{\beta}),\Delta^{k}\widetilde{\beta}\rangle_{L^{2}_{x}}\right|=\left|\int_{\mathbb{T}^{2}}\operatorname{div}\left(\frac{1}{2}\left[\Delta^{k}\widetilde{\beta}\right]^{2}U\right)\right|=0.

Since β0\beta_{0}, gg are assumed to have zero mean, it follows that also 𝕋2β~(x,t)𝑑x=0\int_{\mathbb{T}^{2}}\widetilde{\beta}(x,t)\,dx=0 at later times, and hence we may use the Poincaré inequality to obtain equivalence of norms:

Δkβ~Lx2β~Hx2kCkΔkβ~Lx2.\|\Delta^{k}\widetilde{\beta}\|_{L^{2}_{x}}\leq\|\widetilde{\beta}\|_{H^{2k}_{x}}\leq C_{k}\|\Delta^{k}\widetilde{\beta}\|_{L^{2}_{x}}.

Gronwall’s lemma applied to the differential inequality

ddtΔkβ~Lx22k,UΔkβ~Lx22,\frac{d}{dt}\|\Delta^{k}\widetilde{\beta}\|^{2}_{L^{2}_{x}}\lesssim_{k,U}\|\Delta^{k}\widetilde{\beta}\|^{2}_{L^{2}_{x}},

implies that Δkβ~(t)Lx22eC|tt0|Δkβ0Lx22\|\Delta^{k}\widetilde{\beta}(t)\|^{2}_{L^{2}_{x}}\leq e^{C|t-t_{0}|}\|\Delta^{k}\beta_{0}\|_{L^{2}_{x}}^{2}, with CC depending only on kk and UU. Therefore,

Δkβ~(t)L22(1+C|tt0|)Δkβ0L22,\|\Delta^{k}\widetilde{\beta}(t)\|^{2}_{L^{2}}\leq(1+C|t-t_{0}|)\|\Delta^{k}\beta_{0}\|^{2}_{L^{2}}, (A.11)

where the implied constant in the second estimate is uniform for |tt0|T|t-t_{0}|\leq T. By the equivalence of norms and upon rewriting σ=2k\sigma=2k, we conclude that there exists a constant C=C(σ,U)>0C=C(\sigma,U)>0, such that

E(t;t0)β0HxσC(1+Δt)1/2β0Hσ,\displaystyle\|E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}}\leq C(1+\Delta t)^{1/2}\|\beta_{0}\|_{H^{\sigma}}, (A.12)

whenever |tt0|Δt|t-t_{0}|\leq\Delta t. Combining (A.10) and (A.12), we have shown that there exists a constant C=C(σ,U,g)>0C=C(\sigma,U,g)>0, such that

H(t;t0)E(t;t0)β0HxσC(1+Δt)1/2β0Hσ+CΔt,\displaystyle\|H(t;t_{0})E(t;t_{0})\beta_{0}\|_{H^{\sigma}_{x}}\leq C(1+\Delta t)^{1/2}\|\beta_{0}\|_{H^{\sigma}}+C\Delta t, (A.13)

if |tt0|ΔtT|t-t_{0}|\leq\Delta t\leq T.

Furthermore, from (A.9) and (A.11), we also have

ΔkH(t;t0)E(t;t0)β0Lx2(1+CΔt)1/2Δkβ0Lx2+ΔkgLtLx2Δt,\displaystyle\|\Delta^{k}H(t;t_{0})E(t;t_{0})\beta_{0}\|_{L^{2}_{x}}\leq(1+C\Delta t)^{1/2}\|\Delta^{k}\beta_{0}\|_{L^{2}_{x}}+\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}\Delta t, (A.14)

if |tt0|ΔtT|t-t_{0}|\leq\Delta t\leq T.

Step 2: (estimate for t=tnt=t_{n}) Recall that βΔ{\beta}^{\Delta} is defined recursively by application of H(t;tn)E(t;tn)H(t;t_{n})E(t;t_{n}) over short time-intervals of length Δt\Delta t. Using (A.14), we thus arrive at the recursive estimate

Δkβn+1Δ(tn+1)Lx2(1+CΔt)1/2ΔkβnΔ(tn)Lx2+ΔkgLtLx2Δt.\|\Delta^{k}{\beta}_{n+1}^{\Delta}(t_{n+1})\|_{L^{2}_{x}}\leq(1+C\Delta t)^{1/2}\|\Delta^{k}{\beta}_{n}^{\Delta}(t_{n})\|_{L^{2}_{x}}+\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}\Delta t.

Iterating this inequality backwards until n=0n=0 yields

Δkβn+1Δ(tn+1)Lx2\displaystyle\|\Delta^{k}{\beta}_{n+1}^{\Delta}(t_{n+1})\|_{L^{2}_{x}} (1+CΔt)(n+1)/2Δkβ0Δ(0)Lx2\displaystyle\leq(1+C\Delta t)^{(n+1)/2}\|\Delta^{k}{\beta}_{0}^{\Delta}(0)\|_{L^{2}_{x}} (A.15)
+ΔkgLtLx2Δtj=0n(1+CΔt)j/2.\displaystyle+\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}\Delta t\sum_{j=0}^{n}(1+C\Delta t)^{j/2}. (A.16)

Recall tn=nΔtt_{n}=n\Delta t, with n=0,1,,Nn=0,1,\ldots,N. Then, the first term on the right-hand-side of (A.15) is bounded by

(1+CTN)N/2Δkβ0Δ(0)Lx2,\left(1+C\frac{T}{N}\right)^{N/2}\|\Delta^{k}{\beta}_{0}^{\Delta}(0)\|_{L^{2}_{x}},

which, in turn, converges to eCT/2Δkβ0Δ(0)Lx2e^{CT/2}\|\Delta^{k}{\beta}_{0}^{\Delta}(0)\|_{L^{2}_{x}} as NN\to\infty.

The second term on the right-hand-side of (A.15) is bounded by

ΔkgLtLx2TNj=0N1(1+CTN)j/2,\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}\frac{T}{N}\sum_{j=0}^{N-1}\left(1+C\frac{T}{N}\right)^{j/2},

which converges, as NN\to\infty, to

2ΔkgLtLx2C(eCT/21).2\frac{\|\Delta^{k}g\|_{L^{\infty}_{t}L^{2}_{x}}}{C}(e^{CT/2}-1).

Therefore it follows that

ΔkβnΔ(tn)Lx2T,U,g,kΔkβ0Lx2+1,\|\Delta^{k}{\beta}_{n}^{\Delta}(t_{n})\|_{L^{2}_{x}}\lesssim_{T,U,g,k}\|\Delta^{k}\beta_{0}\|_{L^{2}_{x}}+1,

for all n=0,1,,Nn=0,1,\ldots,N, with an implied constant depending only on T,U,g,kT,U,g,k. Thus, from the equivalence of norms we have that there exists a constant C>0C>0 depending only on TT, σ\sigma, UU, gg and on β0Δ(t0)=β0{\beta}^{\Delta}_{0}(t_{0})=\beta_{0}, such that

maxtnTβΔ(tn)HxσC.\max_{t_{n}\leq T}\|{\beta}^{\Delta}(t_{n})\|_{H^{\sigma}_{x}}\leq C.

Step 3: (conclusion) Appealing once more to the inequality (A.13), i.e.

βΔ(t)Hxσ\displaystyle\|\beta^{\Delta}(t)\|_{H^{\sigma}_{x}} =H(t;tn)E(t;tn)βΔ(tn)Hxσ\displaystyle=\|H(t;t_{n})E(t;t_{n})\beta^{\Delta}(t_{n})\|_{H^{\sigma}_{x}}
C(1+Δt)βΔ(tn)Hσ+CΔt,\displaystyle\leq C(1+\Delta t)\|\beta^{\Delta}(t_{n})\|_{H^{\sigma}}+C\Delta t,

it follows that there exists a constant C=C(T,σ,U,g,β0)>0C=C(T,\sigma,U,g,\beta_{0})>0, such that

supt[0,T]βΔ(t)HxσC.\sup_{t\in[0,T]}\|\beta^{\Delta}(t)\|_{H^{\sigma}_{x}}\leq C.

This is the claimed upper bound. ∎

The solution operator of the heat equation H(t;t0)H(t;t_{0}) evaluated at t=t0t=t_{0} is identity, H(t0;t0)=IH(t_{0};t_{0})=I. Formally expanding in tt, we expect that H(t;t0)=I+O(Δt)H(t;t_{0})=I+O(\Delta t). The following lemma formalizes this fact (with a very crude estimate for the first-order correction term):

Lemma A.4.

Let β0C(𝕋2)\beta_{0}\in C^{\infty}(\mathbb{T}^{2}) be a smooth function, and let t0[0,T]t_{0}\in[0,T]. Then for any t[t0,t0+Δt]t\in[t_{0},t_{0}+\Delta t], and σ0\sigma\geq 0, we have

H(t;t0)β0β0HxσΔtνβ0Hx2+σ+Δt(1+νΔt)gLtHx2+σ.\displaystyle\begin{aligned} \|H(t;t_{0})\beta_{0}-\beta_{0}\|_{H^{\sigma}_{x}}&\leq\Delta t\,\nu\|\beta_{0}\|_{H_{x}^{2+\sigma}}\\ &\qquad+\Delta t(1+\nu\Delta t)\|g\|_{L^{\infty}_{t}H_{x}^{2+\sigma}}.\end{aligned} (A.17)

If GτG_{\tau} denotes the heat kernel, then we similarly have

Gtt0β0β0HxσΔtνβ0Hx2+σ\displaystyle\|G_{t-t_{0}}\ast\beta_{0}-\beta_{0}\|_{H^{\sigma}_{x}}\leq\Delta t\,\nu\|\beta_{0}\|_{H_{x}^{2+\sigma}} (A.18)
Proof.

Let h(t):=H(t;t0)β0h(t):=H(t;t_{0})\beta_{0}. By definition of H(t;t0)H(t;t_{0}), we have

th(t)=νΔh(t)+g(t),\partial_{t}h(t)=\nu\Delta h(t)+g(t),

and h(t0)=β0h(t_{0})=\beta_{0}. Integration in time yields

H(t;t0)β0β0=νt0tΔh(τ)𝑑τ+t0tg(τ)𝑑τ,H(t;t_{0})\beta_{0}-\beta_{0}=\nu\int_{t_{0}}^{t}\Delta h(\tau)\,d\tau+\int_{t_{0}}^{t}g(\tau)\,d\tau,

and we can estimate

H(t;t0)β0β0Hxσ\displaystyle\|H(t;t_{0})\beta_{0}-\beta_{0}\|_{H^{\sigma}_{x}} νt0tΔh(τ)Hxσ𝑑τ+t0tg(τ)Hxσ𝑑τ\displaystyle\leq\nu\int_{t_{0}}^{t}\|\Delta h(\tau)\|_{H^{\sigma}_{x}}\,d\tau+\int_{t_{0}}^{t}\|g(\tau)\|_{H^{\sigma}_{x}}\,d\tau
ΔtνhLtHx2+σ+ΔtgLtHxσ.\displaystyle\leq\Delta t\,\nu\|h\|_{L^{\infty}_{t}H_{x}^{2+\sigma}}+\Delta t\,\|g\|_{L^{\infty}_{t}H_{x}^{\sigma}}.

Taking into account (A.8), we have

h(t)Hx2+σ=H(t;t0)β0H2+σβ0H2+σ+ΔtgLtHxσ.\|h(t)\|_{H^{2+\sigma}_{x}}=\|H(t;t_{0})\beta_{0}\|_{H^{2+\sigma}}\leq\|\beta_{0}\|_{H^{2+\sigma}}+\Delta t\|g\|_{L^{\infty}_{t}H^{\sigma}_{x}}.

Upon substitution of this bound, we thus obtain the (rough) estimate

H(t;t0)β0β0HxσΔtνβ0Hx2+σ+Δt(1+νΔt)gLtHx2+σ.\|H(t;t_{0})\beta_{0}-\beta_{0}\|_{H^{\sigma}_{x}}\leq\Delta t\nu\|\beta_{0}\|_{H_{x}^{2+\sigma}}+\Delta t(1+\nu\Delta t)\|g\|_{L^{\infty}_{t}H_{x}^{2+\sigma}}.

The estimate for Gtt0β0G_{t-t_{0}}\ast\beta_{0} is the special case where g0g\equiv 0. ∎

Using the last lemma, we next show that the operator splitting approximant βΔ{\beta}^{\Delta} is an approximate solution of the relevant equation, up to an O(Δt)O(\Delta t) error.

Lemma A.5.

Assume that β0\beta_{0}, UU and gg are smooth, div(U)=0\operatorname{div}(U)=0 and that β0,g\beta_{0},g have zero mean. Then the function βΔ{\beta}^{\Delta} defined by (A.7) is continuous, and with the potential exception of finitely many break-points t=t0,t1,t=t_{0},t_{1},\dots, the function βΔ\beta^{\Delta} solves the following PDE:

tβΔ+UβΔ=νΔβΔ+g+FΔ,\displaystyle\partial_{t}{\beta}^{\Delta}+U\cdot\nabla{\beta}^{\Delta}=\nu\Delta{\beta}^{\Delta}+g+F^{\Delta}, (A.19)

where FΔF^{\Delta} can be estimated by

FΔLtLx2CΔt,\displaystyle\|F^{\Delta}\|_{L^{\infty}_{t}L^{2}_{x}}\leq C\Delta t, (A.20)

with a constant C=C(T,ν,U,g,β0)>0C=C(T,\nu,U,g,\beta_{0})>0 which is bounded uniformly in Δt\Delta t.

Proof.

Continuity of βΔ{\beta}^{\Delta} is straight-forward. We therefore focus on (A.19). To this end, we consider t[tn,tn+1]t\in[t_{n},t_{n+1}] for n0n\geq 0. By definition, we have

βΔ(t)=βn+1Δ(t)=H(t;tn)[E(t;tn)βnΔ(tn)],t[tn,tn+1].\beta^{\Delta}(t)={\beta}^{\Delta}_{n+1}(t)=H(t;t_{n})\left[E(t;t_{n}){\beta}_{n}^{\Delta}(t_{n})\right],\quad\forall\,t\in[t_{n},t_{n+1}].

For τ1,τ2[tn,tn+1]\tau_{1},\tau_{2}\in[t_{n},t_{n+1}], we now define,

h(τ1,τ2):=H(τ1;tn)[E(τ2;tn)βnΔ(tn)],h(\tau_{1},\tau_{2}):=H(\tau_{1};t_{n})\left[E(\tau_{2};t_{n}){\beta}_{n}^{\Delta}(t_{n})\right],

so that βn+1Δ(t)=h(t,t){\beta}^{\Delta}_{n+1}(t)=h(t,t). Note that the dependency on the spatial variable has been suppressed in this notation, i.e. hh is considered as a mapping

h:[tn,tn+1]×[tn,tn+1]C(𝕋2).h:[t_{n},t_{n+1}]\times[t_{n},t_{n+1}]\to C^{\infty}(\mathbb{T}^{2}).

We next observe that

tβn+1Δ(t)=(τ1h)(t,t)+(τ2h)(t,t).\displaystyle\partial_{t}\beta^{\Delta}_{n+1}(t)=(\partial_{\tau_{1}}h)(t,t)+(\partial_{\tau_{2}}h)(t,t). (A.21)

We will compute the partial derivatives with respect to τ1,τ2\tau_{1},\tau_{2} on the right-hand side of (A.21).

Calculation for τ1\partial_{\tau_{1}}: With τ2\tau_{2} fixed, let β¯(τ2):=E(τ2;tn)β0\overline{\beta}(\tau_{2}):=E(\tau_{2};t_{n})\beta_{0}. By definition of H(τ1;tn)H(\tau_{1};t_{n}), the function τ1h(τ1,τ2)=H(τ1;tn)β¯(τ2)\tau_{1}\mapsto h(\tau_{1},\tau_{2})=H(\tau_{1};t_{n})\overline{\beta}(\tau_{2}) solves the heat equation with initial data β¯(τ2)\overline{\beta}(\tau_{2}). In particular,

τ1h(τ1,τ2)=g(τ1)+νΔh(τ1,τ2).\displaystyle\partial_{\tau_{1}}h(\tau_{1},\tau_{2})=g(\tau_{1})+\nu\Delta h(\tau_{1},\tau_{2}). (A.22)

Calculation for τ2\partial_{\tau_{2}}: To compute the derivative with respect to τ2\tau_{2}, we freeze τ1\tau_{1}, and note that H(τ1;tn)H(\tau_{1};t_{n}) is explicitly given by

H(τ1;tn)β¯(τ2)=Gτ1tnβ¯(τ2)+tnτ1Gτ1τg(τ)𝑑τ,H(\tau_{1};t_{n})\overline{\beta}(\tau_{2})=G_{\tau_{1}-t_{n}}\ast\overline{\beta}(\tau_{2})+\int_{t_{n}}^{\tau_{1}}G_{\tau_{1}-\tau}\ast g(\tau)\,d\tau,

in terms of the heat kernel GτG_{\tau}. Since the second term is independent of τ2\tau_{2}, upon taking a partial derivative of h(τ1,τ2)=H(τ1;tn)β¯(τ2)h(\tau_{1},\tau_{2})=H(\tau_{1};t_{n})\overline{\beta}(\tau_{2}) with respect to τ2\tau_{2}, we obtain,

τ2h(τ1,τ2)=Gτ1tnτ2β¯(τ2).\partial_{\tau_{2}}h(\tau_{1},\tau_{2})=G_{\tau_{1}-t_{n}}\ast\partial_{\tau_{2}}\overline{\beta}(\tau_{2}).

By definition, τ2β¯(τ2)=E(τ2;tn)βnΔ(tn)\tau_{2}\mapsto\overline{\beta}(\tau_{2})=E(\tau_{2};t_{n})\beta_{n}^{\Delta}(t_{n}) solves the transport equation, and hence,

τ2β¯(τ2)=U(τ2)β¯(τ2).\partial_{\tau_{2}}\overline{\beta}(\tau_{2})=-U(\tau_{2})\cdot\nabla\overline{\beta}(\tau_{2}).

Substitution of this identity in our equation for τ2h\partial_{\tau_{2}}h yields,

τ2h(τ1,τ2)=Gτ1tn(U(τ2)β¯(τ2)).\displaystyle\partial_{\tau_{2}}h(\tau_{1},\tau_{2})=-G_{\tau_{1}-t_{n}}\ast\left(U(\tau_{2})\cdot\nabla\overline{\beta}(\tau_{2})\right). (A.23)

Deriving the equation for βnΔ\beta^{\Delta}_{n}: Combining (A.21), (A.22) and (A.23), and recalling that βnΔ(t)=h(t,t)\beta^{\Delta}_{n}(t)=h(t,t), we obtain,

tβnΔ(t)=Gttn(U(t)β¯(t))+g(t)+νΔβnΔ(t).\partial_{t}\beta^{\Delta}_{n}(t)=-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right)+g(t)+\nu\Delta\beta^{\Delta}_{n}(t).

Setting

FΔ(t):=U(t)βnΔ(t)Gttn(U(t)β¯(t)),F^{\Delta}(t):=U(t)\cdot\nabla\beta^{\Delta}_{n}(t)-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right),

we thus have

tβnΔ(t)+U(t)βnΔ(t)=g(t)+νΔβnΔ(t)+FΔ(t).\partial_{t}\beta^{\Delta}_{n}(t)+U(t)\cdot\nabla\beta^{\Delta}_{n}(t)=g(t)+\nu\Delta\beta^{\Delta}_{n}(t)+F^{\Delta}(t).

This is (A.19).

To conclude the proof of Lemma A.5, it thus remains to derive the bound (A.20) on FΔ(t)F^{\Delta}(t).

Estimate for FΔ(t)F^{\Delta}(t): We can bound,

FΔ(t)Lx2\displaystyle\|F^{\Delta}(t)\|_{L^{2}_{x}} =U(t)βnΔ(t)Gttn(U(t)β¯(t))Lx2\displaystyle=\|U(t)\cdot\nabla\beta^{\Delta}_{n}(t)-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right)\|_{L^{2}_{x}}
=U(t)H(t;t0)β¯(t)Gttn(U(t)β¯(t))Lx2\displaystyle=\|U(t)\cdot\nabla H(t;t_{0})\overline{\beta}(t)-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right)\|_{L^{2}_{x}}
U(t){H(t;t0)[β¯(t)]β¯(t)}Lx2\displaystyle\leq\|U(t)\cdot\nabla\left\{H(t;t_{0})\left[\overline{\beta}(t)\right]-\overline{\beta}(t)\right\}\|_{L^{2}_{x}}
+U(t)β¯(t)Gttn(U(t)β¯(t))Lx2\displaystyle\qquad+\|U(t)\cdot\nabla\overline{\beta}(t)-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right)\|_{L^{2}_{x}}
=:(I)+(II).\displaystyle=:(I)+(II).

Applying Lemma A.4, bound (A.17), to the first term, we obtain

(I)\displaystyle(I) =U(t){H(t;tn)[β¯(t)]β¯(t)}Lx2\displaystyle=\big{\|}U(t)\cdot\nabla\left\{H(t;t_{n})\left[\overline{\beta}(t)\right]-\overline{\beta}(t)\right\}\big{\|}_{L^{2}_{x}}
U(t)LH(t;tn)[β¯(t)]β¯(t)H1\displaystyle\leq\|U(t)\|_{L^{\infty}}\left\|H(t;t_{n})[\overline{\beta}(t)]-\overline{\beta}(t)\right\|_{H^{1}}
U(t)L{Δtνβ¯(t)H3+Δt(1+νΔt)g(t)H3}.\displaystyle\leq\|U(t)\|_{L^{\infty}}\left\{\Delta t\,\nu\left\|\overline{\beta}(t)\right\|_{H^{3}}+\Delta t\,(1+\nu\Delta t)\left\|g(t)\right\|_{H^{3}}\right\}.

By Lemma A.3, we can bound β¯(t)H3\|\overline{\beta}(t)\|_{H^{3}} by a constant C=C(T,g,U,β0)>0C=C(T,g,U,\beta_{0})>0, uniform in time and in Δt\Delta t, i.e.

supt[tn,tn+1]β¯(t)H3C.\displaystyle\sup_{t\in[t_{n},t_{n+1}]}\|\overline{\beta}(t)\|_{H^{3}}\leq C. (A.24)

Enlarging the constant CC, if necessary, we thus have

(I)CΔt,(I)\leq C\Delta t,

where C=C(T,ν,g,U,β0)C=C(T,\nu,g,U,\beta_{0}) is independent of tt and Δt\Delta t.

Finally, by Lemma A.4, bound (A.18), we have

(II)\displaystyle(II) =U(t)β¯(t)Gttn(U(t)β¯(t))Lx2\displaystyle=\|U(t)\cdot\nabla\overline{\beta}(t)-G_{t-t_{n}}\ast\left(U(t)\cdot\nabla\overline{\beta}(t)\right)\|_{L^{2}_{x}}
ΔtνU(t)β¯(t)H2\displaystyle\leq\Delta t\,\nu\|U(t)\cdot\nabla\overline{\beta}(t)\|_{H^{2}}
ΔtνU(t)Wx2,β¯(t)H3.\displaystyle\leq\Delta t\,\nu\|U(t)\|_{W^{2,\infty}_{x}}\|\overline{\beta}(t)\|_{H^{3}}.

Invoking (A.24), it follows that

(II)CΔt,(II)\leq C\Delta t,

where C=C(T,ν,g,U,β0)>0C=C(T,\nu,g,U,\beta_{0})>0.

We conclude that

FΔ(t)Lx2(I)+(II)CΔt,\|F^{\Delta}(t)\|_{L^{2}_{x}}\leq(I)+(II)\leq C\Delta t,

for a constant C=C(T,ν,g,U,β0)>0C=C(T,\nu,g,U,\beta_{0})>0, independent of tt and Δt\Delta t. The claimed bound on FΔ(t)F^{\Delta}(t) thus follows upon taking the supremum over all t[tn,tn+1]t\in[t_{n},t_{n+1}] and all nn\in\mathbb{N} such that tnTt_{n}\leq T. This concludes our proof of Lemma A.5. ∎

We can finally state the following convergence result for the operator splitting approximation:

Proposition A.6 (Convergence of operator splitting).

Let β0C(𝕋2)\beta_{0}\in C^{\infty}(\mathbb{T}^{2}) be initial data for the advection-diffusion equation (A.3), with forcing gC(𝕋2×[0,T])g\in C^{\infty}(\mathbb{T}^{2}\times[0,T]) and divergence-free advecting velocity field UC(𝕋2×[0,T])U\in C^{\infty}(\mathbb{T}^{2}\times[0,T]). Assume that 𝕋2β0(x)𝑑x=0\int_{\mathbb{T}^{2}}\beta_{0}(x)\,dx=0 and 𝕋2g(x)𝑑x=0\int_{\mathbb{T}^{2}}g(x)\,dx=0. Let β\beta be the solution of the advection-diffusion PDE (A.3), and let βΔ{\beta}^{\Delta} be given by the operator splitting approximation (A.6),(A.1.3). Then

limΔt0ββΔLtLx2=0.\lim_{\Delta t\to 0}\|\beta-{\beta}^{\Delta}\|_{L^{\infty}_{t}L^{2}_{x}}=0.
Proof.

We define r~Δ:=βΔβ\widetilde{r}^{\Delta}:={\beta}^{\Delta}-\beta. Then, by Lemma A.5, r~Δ\widetilde{r}^{\Delta} solves

tr~Δ+Ur~Δ=νΔr~Δ+FΔ,\partial_{t}\widetilde{r}^{\Delta}+U\cdot\nabla\widetilde{r}^{\Delta}=\nu\Delta\widetilde{r}^{\Delta}+F^{\Delta},

where FΔLx2CΔt\|F^{\Delta}\|_{L^{2}_{x}}\leq C\Delta t for some constant CC independent of Δt\Delta t. Multiplying by r~Δ\widetilde{r}^{\Delta}, integrating over space and employing the Poincaré inequality, r~ΔLx2Cr~ΔLx2\|\widetilde{r}^{\Delta}\|_{L^{2}_{x}}\leq C\|\nabla\widetilde{r}^{\Delta}\|_{L^{2}_{x}}, readily yields

ddtr~ΔLx22\displaystyle\frac{d}{dt}\|\widetilde{r}^{\Delta}\|_{L^{2}_{x}}^{2} 2νr~ΔLx22+2FΔLx2r~ΔLx2\displaystyle\leq-2\nu\|\nabla\widetilde{r}^{\Delta}\|_{L^{2}_{x}}^{2}+2\|F^{\Delta}\|_{L^{2}_{x}}\|\widetilde{r}^{\Delta}\|_{L^{2}_{x}}
2νr~ΔLx22+Cνr~ΔLx22+(Cν)1FΔLx22\displaystyle\leq-2\nu\|\nabla\widetilde{r}^{\Delta}\|_{L^{2}_{x}}^{2}+C\nu\|\widetilde{r}^{\Delta}\|_{L^{2}_{x}}^{2}+(C\nu)^{-1}\|F^{\Delta}\|_{L^{2}_{x}}^{2}
(Cν)1FΔLx22Cν1Δt2,\displaystyle\leq(C\nu)^{-1}\|F^{\Delta}\|_{L^{2}_{x}}^{2}\leq C\nu^{-1}\Delta t^{2},

for some constant C>0C>0, independent of Δt\Delta t. Noting that r~(t=0)=0\widetilde{r}(t=0)=0, we conclude that

supt[0,T]r~Δ(t)Lx22CTν1Δt20,\sup_{t\in[0,T]}\|\widetilde{r}^{\Delta}(t)\|_{L^{2}_{x}}^{2}\leq CT\nu^{-1}\Delta t^{2}\to 0,

as Δt0\Delta t\to 0. This shows that

limΔt0ββΔLtLx2=0,\lim_{\Delta t\to 0}\|\beta-{\beta}^{\Delta}\|_{L^{\infty}_{t}L^{2}_{x}}=0,

i.e. the solution computed by operator splitting converges to the exact solution. ∎

References