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

On the convergence of trajectory statistical solutions

Anne C. Bronzi Instituto de Matemática, Estatística e Computação Científica, Universidade Estadual de Campinas (UNICAMP), Brazil [email protected] Cecilia F. Mondaini Department of Mathematics, Drexel University, USA [email protected]  and  Ricardo M. S. Rosa Instituto de Matemática, Universidade Federal do Rio de Janeiro (UFRJ), Brazil [email protected]
Abstract.

In this work, a recently introduced general framework for trajectory statistical solutions is considered, and the question of convergence of families of such solutions is addressed. Conditions for the convergence are given which rely on natural assumptions related to a priori estimates for the individual solutions of typical approximating problems. The first main result is based on the assumption that the superior limit of suitable families of compact subsets of carriers of the family of trajectory statistical solutions be included in the set of solutions of the limit problem. The second main result is a version of the former in the case in which the approximating family is associated with a well-posed system. These two results are then applied to the inviscid limit of incompressible Navier-Stokes system in two and three spatial dimensions, showing, in particular, the existence of trajectory statistical solutions to the two- and three-dimensional Euler equations, in the context of weak and dissipative solutions, respectively. Another application of the second main result is on the Galerkin approximations of statistical solutions of the three-dimensional Navier-Stokes equations.

Key words and phrases:
statistical solution; trajectory statistical solution; convergence; Navier-Stokes equations; Euler equations; Galerkin approximation
2020 Mathematics Subject Classification:
76D06, 35Q30, 35Q31, 35Q35, 60B05, 60B10, 82M10

1. Introduction

The theory of statistical solutions for the Navier-Stokes equations, initiated by Foias and Prodi in the early 1970s [33, 34, 39], and followed by Vishik and Fursikov later in the decade [64, 65], has seen many advances in relation to the theory of turbulence in fluid flow problems [35, 52, 22, 7, 26, 36, 43, 38, 37, 5, 23, 57, 68, 50, 59, 8]. The concept has also been adapted to a number of other specific equations, or specific classes of equations, for which a well-defined evolution semigroup does not exist or is not known to exist [47, 44, 3, 46, 60, 12]. Later on, [40] developed a slightly different formulation of the statistical solutions in [64, 65] which is compatible with those in [33, 34, 39], in the way that projections of the solutions in this new formulation become solutions in the sense of the latter.

More recently, inspired by [40], the authors introduced, in [13], a general framework applicable to a wide range of equations and containing the main results on the existence of statistical solutions for an associated initial-valued problem, based on natural and amenable conditions. This greatly facilitates the application of the notion of statistical solution both to the Navier-Stokes equations and to other systems. In this manuscript, we address the problem of convergence of families of statistical solutions, within this general framework.

Approximating a given problem is of fundamental importance in every branch of Mathematics, both pure and applied, and in many scientific fields and applications. This is no different in regards to statistical solutions. Be it with the aim of proving the existence of solutions, or in relation to asymptotic analysis, numerical computations and other investigations on the nature of the given problem. With this in mind, we present a couple of general results on the limit of trajectory statistical solutions and apply the results to the Euler equations as inviscid limits of the Navier-Stokes equations and to the Galerkin approximations of the Navier-Stokes equations. The inviscid limit is treated both in the two- and three-dimensional cases. In particular, our results show the existence of trajectory statistical solution to the limit Euler equations. The Galerkin approximation is considered in the three-dimensional viscous case, yielding, in particular, a different proof of existence of statistical solutions for the three-dimensional Navier-Stokes equations than given in [40, 13], see Remark 4.4.

Our two main results concern conditions for the limit of trajectory statistical solutions, of a given family of problems, to be a trajectory statistical solution of the limit problem. The first result, Theorem 3.1, concerns the more general case of a family of trajectory statistical solutions associated with an approximate problem which is not necessarily well-posed. The second result, Theorem 3.2, concerns the special case in which the family of trajectory statistical solutions of the approximate problem is associated with a well-defined semigroup of individual solutions, although, at the limit, the well-posedness may not stand.

We recall that a 𝒰\mathcal{U}-trajectory statistical solution, as introduced in [13], is a Borel probability measure ρ\rho which is tight and is carried by a Borel subset of a set 𝒰\mathcal{U} on a space 𝒳=𝒞loc(I,X)\mathcal{X}=\mathcal{C}_{loc}(I,X) of continuous functions from a real interval II into a Hausdorff space XX, with 𝒳\mathcal{X} endowed with the topology of uniform convergence on compact subsets of II (see Definition 3.1). The term solution here is in regards to the problem of finding such Borel probability measure carried inside 𝒰\mathcal{U}. This is not a problem per se, because any Dirac measure carried at an individual element in 𝒰\mathcal{U} is a trajectory statistical solution, but the problem becomes nontrivial when associated with an initial condition Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}, where Πt0:𝒳X\Pi_{t_{0}}:\mathcal{X}\rightarrow X is the projection at a “initial” time t0It_{0}\in I and μ0\mu_{0} is a given “initial” measure on XX. In applications, 𝒰\mathcal{U} is, for example, associated with the set of solutions of a given differential equation, with values in a phase space XX, such as the Leray-Hopf weak solutions of the 3D Navier-Stokes equations.

Suppose now that we have a family {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} of 𝒰ε\mathcal{U}_{\varepsilon}-trajectory statistical solutions, for some index set \mathcal{E} and with 𝒰ε𝒳\mathcal{U}_{\varepsilon}\subset\mathcal{X}. Think of 𝒰ε\mathcal{U}_{\varepsilon} as the space of solutions of a given differential equation depending on a parameter ε\varepsilon (e.g. the Leray-Hopf weak solutions of the 3D Navier-Stokes equations with the viscosity as the parameter, or a numerical approximation of a given equation depending on a certain discretization parameter) and of 𝒰\mathcal{U} as a set of solutions of the limit problem (e.g. dissipative solutions of the Euler equations). We establish, in Theorem 3.1, suitable conditions on the net {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} so that it converges (passing to a subnet if necessary), in a certain sense, to a 𝒰\mathcal{U}-trajectory statistical solution ρ\rho. More precisely, Theorem 3.1 imposes a uniform tightness assumption on {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon} along with a suitable condition relating the limit of compact subsets of 𝒰ε\mathcal{U}_{\varepsilon} to 𝒰\mathcal{U}, i.e. concerning the limit of approximate individual solutions.

Theorem 3.2 treats the special case in applications where each member in the family of approximating equations possesses a well-defined solution operator Sε:X𝒳S_{\varepsilon}:X\to\mathcal{X}. Namely, SεS_{\varepsilon} takes each initial datum u0,εXu_{0,\varepsilon}\in X to the unique trajectory uε=uε(t)u_{\varepsilon}=u_{\varepsilon}(t) in 𝒳\mathcal{X} of the corresponding approximate equation satisfying the initial condition uε(t0)=u0,εu_{\varepsilon}(t_{0})=u_{0,\varepsilon}. In this case, a set of assumptions is required from the family of operators {Sε}ε\{S_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} to guarantee that, given any initial probability measure μ0\mu_{0} on XX and suitable measures {με}ε\{\mu_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} on XX approximating μ0\mu_{0} in a certain sense, the family of measures ρε=Sεμε\rho_{\varepsilon}=S_{\varepsilon}\mu_{\varepsilon}, ε\varepsilon\in\mathcal{E}, on 𝒳\mathcal{X} has a convergent subnet to a 𝒰\mathcal{U}-trajectory statistical solution ρ\rho such that Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

In applications, such assumptions translate to verifying the following for the approximating equations: (i) continuous dependence of solutions of each approximating equation with respect to initial data lying in compact sets; (ii) suitable parameter-uniform a priori estimates; and (iii) convergence of individual solutions of the approximating systems starting from a fixed compact set towards a solution of the limit equation. This latter condition is weaker than the corresponding condition in Theorem 3.1 and more natural in this context, and for this reason we do not apply Theorem 3.1 directly to prove Theorem 3.2; see Remark 3.6.

The applications are treated in detail in Section 4. The first example, in Section 4.2, concerns the inviscid limit of the Navier-Stokes equations to the Euler equations in the two-dimensional case, illustrating an application of Theorem 3.2. We consider, more precisely, the set of weak solutions of the 2D Euler equations with periodic boundary conditions and initial vorticity in LrL^{r}, with 2r<2\leq r<\infty. In this case, a semigroup is not known to exist at the inviscid limit, but the viscous approximation has a well defined semigroup associated to weak solutions of the 2D Navier-Stokes equations under this setting (see Theorem 4.1).

The second example, in Section 4.3, is the inviscid limit of the Navier-Stokes equations to the Euler equations in three dimensions, where we consider periodic boundary conditions and initial data in L2L^{2}. For the 3D Euler equations, we consider the corresponding set of dissipative solutions, whereas for the 3D Navier-Stokes equations we consider the set of Leray-Hopf weak solutions. This provides an application of Theorem 3.1, given that no semigroup is available under this setting, neither at the limit, nor for the approximating family (see Theorem 4.2). Here, the verification of item (iii), namely the convergence of individual solutions of the 3D Navier-Stokes equations towards a solution of 3D Euler is in general a delicate issue, but is nevertheless known to hold within the context of dissipative solutions for Euler. See a more detailed discussion on this in Section 4.1, and also 4.4.

Our final example is given in Section 4.4 and concerns a spectral spatial discretization of the 3D Navier-Stokes equations given by the standard Galerkin approximation. Since the initial-value problem associated to the Galerkin system is well-posed, this gives another application of Theorem 3.2. Moreover, as we point out in Remark 4.2 and Remark 4.3 below, the flexibility provided by the framework of Theorem 3.2 with the approximating initial measures {με}ε\{\mu_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} allows not only for the natural example given by Galerkin projections of a given initial measure μ0\mu_{0} for the limiting system, namely με=μm=Pmμ0\mu_{\varepsilon}=\mu_{m}=P_{m}\mu_{0}, but also the example of Monte Carlo approximations of μ0\mu_{0}, provided the required conditions from Theorem 3.2 are met.

Regarding the first two applications, we first note that the two-dimensional inviscid limit that we present here for the periodic case has also been treated by Wagner and Wiedemann [67], in the whole space, with no forcing term. The result in [67] uses the general framework given in [13], and we take this opportunity to show how the general framework developed here, specifically in Theorem 3.2, can be applied to simplify the corresponding proof. Moreover, differently than [67], our application allows the presence of a time-dependent forcing term.

In regard to the application for the inviscid limit in 3D, we point out that a similar question was tackled in [31], although in the context of a certain notion of statistical solution in phase space called Friedman-Keller statistical solution [29, 30], which particularly takes into account the temporal evolution of multi-point spatial correlations in the flow. Specifically, [31] shows two main results. The first concerns an equivalence between the definitions of Friedman-Keller statistical solution and Foias-Prodi statistical solution from [33, 34, 39] for the 3D Navier-Stokes equations. Secondly, it is shown that, under a certain statistical scaling assumption, any suitable sequence of Friedman-Keller statistical solutions for the 3D Navier-Stokes equations converges, up to a subsequence, towards a corresponding one for the 3D Euler equations under the inviscid limit, in a certain time-averaged sense.

While the statements of our general convergence results, Theorem 3.1 and Theorem 3.2, are given with respect to statistical solutions in trajectory space, we note that a general notion of statistical solution in phase space was given in our previous work, [13]. These consist of time-parametrized families of Borel probability measures on XX satisfying suitable regularity conditions and a Liouville-type equation associated with an evolution equation du/dt=F(t,u)\mathrm{d}u/\mathrm{d}t=F(t,u). We also show in [13] that any suitably regular trajectory statistical solution ρ\rho on 𝒳\mathcal{X} yields a phase-space statistical solution via the time projections μt=Πtρ\mu_{t}=\Pi_{t}\rho, tIt\in I, called a projected statistical solution, although the converse might not necessarily hold. As such, if {Πtρε}tI\{\Pi_{t}\rho_{\varepsilon}\}_{t\in I}, ε\varepsilon\in\mathcal{E}, is a collection of projected statistical solutions for which {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} satisfies our conditions guaranteeing convergence to a trajectory statistical solution ρ\rho in ε\varepsilon, then it follows immediately from the continuity of the projection operator Πt\Pi_{t}, together with the notion of convergence in the space of probability measures we consider, that Πtρε\Pi_{t}\rho_{\varepsilon} converges to Πtρ\Pi_{t}\rho in ε\varepsilon, for each tIt\in I. Therefore, under such conditions and provided ρ\rho is sufficiently regular, the projected statistical solutions {Πtρε}tI\{\Pi_{t}\rho_{\varepsilon}\}_{t\in I}, ε\varepsilon\in\mathcal{E}, converge to a projected statistical solution of the limit problem, namely {Πtρ}tI\{\Pi_{t}\rho\}_{t\in I}.

Moreover, we recall that, as shown in one of the applications from [13], any suitable projected statistical solution of the Navier-Stokes equations is also a Foias-Prodi statistical solution, and hence it corresponds to a Friedman-Keller statistical solution as in [31]. Thus, in comparison to the aforementioned inviscid convergence result from [31], it would be interesting to investigate: (a) whether a projected statistical solution of the Euler equations corresponds to a solution in the Friedman-Keller sense, or, more generally, whether the notion of phase-space statistical solutions of the Euler equations, in the sense of [13], is equivalent to the concept of Friedman-Keller statistical solutions for such equations; and (b) what is the relation between the required assumptions and type of convergence from the result in [31] and the pointwise-in-time convergence for projected statistical solutions implied by our current results. We leave such investigation for future work.

We emphasize that, besides these illustrative applications, the framework is quite general and several other limiting problems fit within the scope of our theory. For example, the results apply to various different numerical discretizations; to the viscous approximations of the inviscid magnetohydrodynamic (MHD) equations; to the NSE-α\alpha and MHD-α\alpha models as regularized approximations of the NSE and MHD equations, respectively; to compressible approximations of the incompressible 3D NSE equations; and to approximations of other models such as reaction-diffusion equations and nonlinear wave equations [11, 13].

To further connect with the existing literature, we mention that there are a number of previous results concerning the limit of statistical solutions in various senses, with the majority given in the context of the Navier-Stokes equations. Those include the inviscid limit of the Navier-Stokes equations (or a damped version of it) towards the Euler equations in [18, 23], besides the previously mentioned work [67]; the vanishing α\alpha limit of the α\alpha-Navier-Stokes equations [66, 11]; the Navier-Stokes equations as the limit of the viscoelastic Navier-Stokes-Voigt model [58]; and the infinite Prandtl number limit of Rayleigh-Bénard convection [68]. Our results, however, instead of focusing on a specific model, apply to a general framework that can be more easily verified in specific cases.

It is also worth mentioning a yet another notion of statistical solution based on the concept of Markov selection, developed in [17], for general class of evolutionary problems, and in [28], in the context of weak solutions of the barotropic Navier-Stokes systems, inspired by previous works in the context of stochastic equations [32, 10]. The connection of this type of statistical solution with our theory and the limiting process of such solutions is not clear but is currently under investigation.

This manuscript is organized as follows. Section 2 recalls the necessary background on certain functional analytical and measure theoretical tools, including the notions of convergence in spaces of probability measures that we consider. Our general results on convergence of statistical solutions are given in Section 3. Finally, Section 4 presents our applications of these general results, namely the convergence of trajectory statistical solutions in the 2D and 3D inviscid limits, and the Galerkin approximations.

2. Preliminaries

In this section, we briefly recall the basic topological and measure theoretical concepts underlying our framework. For further details, we refer to e.g. [2, 9, 13, 41, 63].

2.1. Functional setting

Given a topological vector space XX, we denote its dual by XX^{\prime} and the corresponding duality product as ,X,X\langle\cdot,\cdot\rangle_{X^{\prime},X}. We employ the notation XwX_{\textrm{\rm w}} to indicate that XX is endowed with its weak topology, whereas XwX_{\textrm{\rm w}^{*}}^{\prime} stands for XX^{\prime} endowed with the weak-star topology. Notice that, for any topological vector space XX, the space XwX_{\textrm{\rm w}^{*}}^{\prime} is always a Hausdorff locally convex topological vector space ([27, Section 1.11.1]). Further, if XX is in particular a Banach space, we denote its norm by X\|\cdot\|_{X}, and by X\|\cdot\|_{X^{\prime}} the usual operator norm in the dual space XX^{\prime}.

For any Hausdorff space XX and interval II\subset\mathbb{R}, we denote by 𝒞(I,X)\mathcal{C}(I,X) the space of continuous paths in XX defined on II, i.e. the space of all functions u:IXu:I\rightarrow X which are continuous. When 𝒞(I,X)\mathcal{C}(I,X) is endowed with the compact-open topology, we denote 𝒳=𝒞loc(I,X)\mathcal{X}=\mathcal{C}_{\textrm{\rm loc}}(I,X). Here, the subscript “loc” refers to the fact that this topology is based on compact subintervals of II. We recall that when XX is a uniform space, the compact-open topology in 𝒞loc(I,X)\mathcal{C}_{\textrm{\rm loc}}(I,X) coincides with the topology of uniform convergence on compact subsets [48, Theorem 7.11]. In particular, this holds when XX is a topological vector space, which is the case in both applications presented in Section 4.

For any tIt\in I, let Πt:𝒳X\Pi_{t}:\mathcal{X}\rightarrow X be the “projection” map at time tt, defined by

Πtu=u(t), for all u𝒳.\displaystyle\Pi_{t}u=u(t),\quad\mbox{ for all }u\in\mathcal{X}. (2.1)

Moreover, given any subset III^{\prime}\subset I, define ΠI:𝒞loc(I,X)𝒞loc(I,X)\Pi_{I^{\prime}}:\mathcal{C}_{\textrm{\rm loc}}(I,X)\to\mathcal{C}_{\textrm{\rm loc}}(I^{\prime},X) to be the restriction operator

ΠIu=u|I, for all u𝒞loc(I,X).\displaystyle\Pi_{I^{\prime}}u=u|_{I^{\prime}},\quad\mbox{ for all }u\in\mathcal{C}_{\textrm{\rm loc}}(I,X). (2.2)

It is readily verified that Πt\Pi_{t} and ΠI\Pi_{I^{\prime}} are continuous with respect to the compact-open topology.

We also consider the space of bounded and continuous real-valued functions on XX, denoted by 𝒞b(X)\mathcal{C}_{b}(X). When XX is a subset of m\mathbb{R}^{m}, mm\in\mathbb{N}, we further consider the space 𝒞c(X)\mathcal{C}_{c}^{\infty}(X) of infinitely differentiable real-valued functions on XX which are compactly supported in the interior of XX.

When XX is a Hausdorff topological space, (Γ,)(\Gamma,\preccurlyeq) is a directed set, and {Aγ}γΓ\{A_{\gamma}\}_{\gamma\in\Gamma} is a net in XX, we recall the definitions of topological inferior and superior limits (see [16, Definition 2.2.3], where they are called interior and exterior limits, and also [6, Definition 5.2.1 and Proposition 5.2.2], where they are called lower and upper closed limits, respectively):

lim infγAγ={λΛAλ¯:Λ is a cofinal subset of Γ},\liminf_{\gamma}A_{\gamma}=\bigcap\left\{\overline{\bigcup_{\lambda\in\Lambda}A_{\lambda}}:\Lambda\mbox{ is a cofinal subset of }\Gamma\right\},
lim supγAγ={λΛAλ¯:Λ is a terminal subset of Γ}.\limsup_{\gamma}A_{\gamma}=\bigcap\left\{\overline{\bigcup_{\lambda\in\Lambda}A_{\lambda}}:\Lambda\mbox{ is a terminal subset of }\Gamma\right\}.

In these definitions, the overline denotes the closure of the set under the topology of XX; a subset Λ\Lambda is cofinal in Γ\Gamma when for every γΓ\gamma\in\Gamma, there exists λΛ\lambda\in\Lambda such that γλ\gamma\preccurlyeq\lambda (like a subsequence); and a subset Λ\Lambda is terminal in Γ\Gamma when there exists γΓ\gamma\in\Gamma such that Λ={λΓ;γλ}\Lambda=\{\lambda\in\Gamma;\;\gamma\preccurlyeq\lambda\} (like the tails of a sequence). Note that these definitions of topological inferior and superior limits are different from the set-theoretic limits, where no topological closure is used in the definitions.

When both limits agree, we define the result as the topological limit of the net:

limγAγ=lim supγAγ=lim infγAγ.\lim_{\gamma}A_{\gamma}=\limsup_{\gamma}A_{\gamma}=\liminf_{\gamma}A_{\gamma}.

In this work, we only use the superior limit, for which the following characterization is useful, relating it to the classical definition for sequences (see [42, Theorem 2.10]):

lim supγAγ=γΓγβAβ¯.\displaystyle\limsup_{\gamma}A_{\gamma}=\bigcap_{\gamma\in\Gamma}\overline{\bigcup_{\gamma\preccurlyeq\beta}A_{\beta}}. (2.3)

2.2. Elements of measure theory

Consider a topological space XX and let X\mathcal{B}_{X} denote the σ\sigma-algebra of Borel sets in XX. We denote by (X)\mathcal{M}(X) the set of finite Borel measures in XX, and by 𝒫(X)\mathcal{P}(X) its subset of all Borel probability measures in XX.

Given a family 𝔎\mathfrak{K} of Borel sets in XX, we say that a measure μ(X)\mu\in\mathcal{M}(X) is inner regular with respect to 𝔎\mathfrak{K} if for every set AXA\in\mathcal{B}_{X},

μ(A)=sup{μ(K):K𝔎,KA}.\mu(A)=\sup\{\mu(K):K\in\mathfrak{K},\;K\subset A\}.

A measure μ(X)\mu\in\mathcal{M}(X) is tight or Radon if μ\mu is inner regular with respect to the family of all compact subsets of XX. Moreover, a measure μ(X)\mu\in\mathcal{M}(X) is called outer regular if for every set AXA\in\mathcal{B}_{X},

μ(A)=inf{μ(U):U is open,AU}.\mu(A)=\inf\{\mu(U):U\mbox{ is open},\;A\subset U\}.

A net {μγ}γΓ\{\mu_{\gamma}\}_{\gamma\in\Gamma} of measures in (X)\mathcal{M}(X) is said to be uniformly tight if for every ε>0\varepsilon>0 there exists a compact set KXK\subset X such that

μγ(XK)<ε, for all γΓ.\mu_{\gamma}(X\setminus K)<\varepsilon,\quad\mbox{ for all }\gamma\in\Gamma.

Here we follow the terminology in [9], and we remark that such concept of uniform tightness does not necessarily imply tightness of each measure μγ\mu_{\gamma}. In what follows, however, we consider uniformly tight families of tight measures.

We denote the set of all measures μ(X)\mu\in\mathcal{M}(X) which are tight by (X,tight)\mathcal{M}(X,\rm{tight}), and its subset encompassing all tight Borel probability measures by 𝒫(X,tight)\mathcal{P}(X,\rm{tight}).

Let us recall some useful facts regarding these definitions. First, every tight finite Borel measure on a Hausdorff space XX is outer regular, see [2, Theorem 12.4]. Furthermore, if XX is a Polish space then every finite Borel measure on XX is tight, see [2, Theorem 12.7]. We will make use of this latter result in Section 4 in connection with the fact that in every separable Banach space XX, the Borel σ\sigma-algebras generated by the strong and weak topologies coincide, i.e. X=Xw\mathcal{B}_{X}=\mathcal{B}_{X_{\textrm{\rm w}}}, see e.g. [55, Section 2.2].

Now let XX and YY be Hausdorff spaces and consider a Borel measurable function F:XYF:X\rightarrow Y. Then every measure μ\mu on X\mathcal{B}_{X} induces a measure FμF\mu on Y\mathcal{B}_{Y} known as the push-forward of μ\mu by FF and defined as

Fμ(E)=μ(F1(E)),for all EY.F\mu(E)=\mu(F^{-1}(E)),\quad\mbox{for all }E\in\mathcal{B}_{Y}.

Moreover, if φ:Y\varphi:Y\rightarrow\mathbb{R} is a FμF\mu-integrable function then φF\varphi\circ F is μ\mu-integrable and the following change of variables formula holds

Xφ(F(x))𝑑μ(x)=Yφ(y)𝑑Fμ(y),\int_{X}\varphi(F(x))d\mu(x)=\int_{Y}\varphi(y)dF\mu(y), (2.4)

see e.g. [2]. Clearly, if μ\mu is a tight measure and FF is a continuous function then the push-forward measure FμF\mu is also tight.

Next, we present a generalization for nets of the continuity of a finite measure with respect to a decreasing sequence of measurable sets (see [41, Theorem 1.8] or [2, Lemma 4.51]). The following proof is based on similar ideas from [49, Proposition 10].

Lemma 2.1.

(Continuity from above) Let XX be a compact Hausdorff space and let μ(X)\mu\in\mathcal{M}(X) be an outer regular measure. Then, for any monotone decreasing net (Eγ)γΓ(E_{\gamma})_{\gamma\in\Gamma} of compact sets in XX, E=γΓEγE=\bigcap_{\gamma\in\Gamma}E_{\gamma} is a compact set and

μ(E)=limγΓμ(Eγ).\mu(E)=\lim_{\gamma\in\Gamma}\mu(E_{\gamma}).
Proof.

It is clear that E=γΓEγE=\bigcap_{\gamma\in\Gamma}E_{\gamma} is a closed subset of a compact set so that it is compact. Since μ\mu is outer regular then, for every ε>0\varepsilon>0, there exists an open set UXU\subset X such that EUE\subset U and μ(U)<μ(E)+ε\mu(U)<\mu(E)+\varepsilon. Observe that the compactness of XX implies the compactness of UcU^{c} and, since each EγE_{\gamma} is compact, we also have that EγcE_{\gamma}^{c} is an open set on XX. Furthermore, UcγΓEγcU^{c}\subset\bigcup_{\gamma\in\Gamma}E_{\gamma}^{c} so that there exists a finite subset {γ1,,γN}Γ\{\gamma_{1},\ldots,\gamma_{N}\}\subset\Gamma such that i=1NEγiU\bigcap_{i=1}^{N}E_{\gamma_{i}}\subset U. Since (Γ,)(\Gamma,\preccurlyeq) is a directed set then there exists γ¯Γ\bar{\gamma}\in\Gamma such that γiγ¯\gamma_{i}\preccurlyeq\bar{\gamma}, for all i=1,,Ni=1,\ldots,N. Since the net (Eγ)γΓ(E_{\gamma})_{\gamma\in\Gamma} is monotone decreasing, we have that Eγ¯EγiE_{\bar{\gamma}}\subset E_{\gamma_{i}} for every i=1,,Ni=1,\ldots,N. Hence, Eγ¯i=1NEγiUE_{\bar{\gamma}}\subset\bigcap_{i=1}^{N}E_{\gamma_{i}}\subset U. Therefore, infγΓμ(Eγ)μ(U)<μ(E)+ε\inf_{\gamma\in\Gamma}\mu(E_{\gamma})\leq\mu(U)<\mu(E)+\varepsilon. Since ε>0\varepsilon>0 is arbitrary we conclude that infγΓμ(Eγ)μ(E)\inf_{\gamma\in\Gamma}\mu(E_{\gamma})\leq\mu(E). On the other hand, it is clear that μ(E)infγΓμ(Eγ)\mu(E)\leq\inf_{\gamma\in\Gamma}\mu(E_{\gamma}). Thus,

limγΓμ(Eγ)=infγΓμ(Eγ)=μ(E),\lim_{\gamma\in\Gamma}\mu(E_{\gamma})=\inf_{\gamma\in\Gamma}\mu(E_{\gamma})=\mu(E),

as desired. ∎

2.3. Topologies for measure spaces and related results

We recall the definitions of two specific topologies in (X)\mathcal{M}(X), for any topological space XX. First, the weak-star topology is the smallest one for which the mappings μμ(f)=Xf(x)𝑑μ(x)\mu\mapsto\mu(f)=\int_{X}f(x)d\mu(x) are continuous, for every bounded and continuous real-valued function ff on XX, i.e. f𝒞b(X)f\in\mathcal{C}_{b}(X). If a net {μα}α\{\mu_{\alpha}\}_{\alpha} converges to μ\mu with respect to this topology, we denote μαwμ\mu_{\alpha}\stackrel{{\scriptstyle w*}}{{\rightharpoonup}}\mu. A less common topology is the one defined by Topsoe in [63], which is the smallest one for which the mappings μμ(f)\mu\mapsto\mu(f) are upper semicontinuous, for every bounded and upper semi-continuous real-valued function ff on XX. Topsoe calls this topology the “weak topology”, but in order to avoid any confusion we call it here the weak-star semicontinuity topology on (X)\mathcal{M}(X). We denote convergence of a net {μα}α\{\mu_{\alpha}\}_{\alpha} to μ\mu with respect to this latter topology as μαwscμ\mu_{\alpha}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu.

From these definitions, it is not difficult to see that the weak-star topology is in general weaker than the weak-star semi-continuity topology. Moreover, according to Lemma 2.2 below, if XX is a completely regular Hausdorff space, then these two topologies coincide when restricted to the space (X,tight)\mathcal{M}(X,\rm{tight}).

The following lemma summarizes some properties and useful characterizations for these topologies, see [63, Theorem 8.1].

Lemma 2.2.

Let XX be a Hausdorff space. For a net {μα}α\{\mu_{\alpha}\}_{\alpha} in (X)\mathcal{M}(X) and μ(X)\mu\in\mathcal{M}(X), consider the following statements:

  1. (i)

    μαwscμ\mu_{\alpha}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu;

  2. (ii)

    lim supμα(f)μ(f)\limsup\mu_{\alpha}(f)\leq\mu(f), for all bounded upper semicontinuous function ff;

  3. (iii)

    lim infμα(f)μ(f)\liminf\mu_{\alpha}(f)\geq\mu(f), for all bounded lower semicontinuous function ff;

  4. (iv)

    limαμα(X)=μ(X)\lim_{\alpha}\mu_{\alpha}(X)=\mu(X) and lim supμα(F)μ(F)\limsup\mu_{\alpha}(F)\leq\mu(F), for all closed set FXF\subset X;

  5. (v)

    limαμα(X)=μ(X)\lim_{\alpha}\mu_{\alpha}(X)=\mu(X) and lim infμα(G)μ(G)\liminf\mu_{\alpha}(G)\geq\mu(G), for all open set GXG\subset X;

  6. (vi)

    μαwμ\mu_{\alpha}\stackrel{{\scriptstyle w*}}{{\rightharpoonup}}\mu.

Then the first five statements are equivalent and each of them implies the last one.

Furthermore, if XX is also completely regular and μ(X,tight)\mu\in\mathcal{M}(X,\rm{tight}), then all six statements are equivalent.

We next state a result of compactness on the space of tight measures (X,tight)\mathcal{M}(X,\rm{tight}) that is essential for our main result. For a proof of this fact, see [63, Theorem 9.1].

Theorem 2.1.

Let XX be a Hausdorff space and let {μα}α\{\mu_{\alpha}\}_{\alpha} be a net in (X,tight)\mathcal{M}(X,\rm{tight}) such that lim supμα(X)<\limsup\mu_{\alpha}(X)<\infty. If {μα}α\{\mu_{\alpha}\}_{\alpha} is uniformly tight, then it is compact with respect to the weak-star semi-continuity topology in (X,tight)\mathcal{M}(X,\rm{tight}).

An important property of the space (X,tight)\mathcal{M}(X,\rm{tight}) is that it is Hausdorff when endowed with the weak-star semi-continuity (respectively, weak-star) topology whenever XX is a Hausdorff (respectively, completely regular Hausdorff) space. This result was proved by Topsoe [63] but we also refer the reader to [13, Section 2.4] for a more detailed proof. As a consequence of this Hausdorff property and the definition of the weak-star topology, one immediately obtains the following characterization of the equality of two measures in (X,tight)\mathcal{M}(X,\rm{tight}) for any completely regular space XX.

Proposition 2.1.

Let XX be a completely regular Hausdorff space and μ1,μ2(X,tight)\mu_{1},\mu_{2}\in\mathcal{M}(X,\rm{tight}). Then

μ1=μ2 if and only if Xφ(x)dμ1(x)=Xφ(x)dμ2(x) for all φ𝒞b(X).\mu_{1}=\mu_{2}\,\,\mbox{ if and only if }\,\,\int_{X}\varphi(x){\text{\rm d}}\mu_{1}(x)=\int_{X}\varphi(x){\text{\rm d}}\mu_{2}(x)\,\,\mbox{ for all }\varphi\in\mathcal{C}_{b}(X). (2.5)

3. Convergence of trajectory statistical solutions

This section contains our main results regarding convergence of trajectory statistical solutions in the sense defined in [13], which we recall in Definition 3.1 below. Throughout the section, we let XX be a Hausdorff space and denote as before by 𝒳=𝒞loc(I,X)\mathcal{X}=\mathcal{C}_{loc}(I,X) the space of continuous functions from II into XX endowed with the topology of uniform convergence on compact sets. We also denote by X\mathcal{B}_{X} and 𝒳\mathcal{B}_{\mathcal{X}} the Borel σ\sigma-algebras of XX and 𝒳\mathcal{X}, respectively.

As pointed out in [13], we note that the terminology “trajectory statistical solutions” refers to the fact that these are measures on 𝒳\mathcal{X} carried by a measurable subset of a fixed set 𝒰𝒳\mathcal{U}\subset\mathcal{X} which, in applications, would consist of the set of trajectories, i.e. solutions in a certain sense, of a given evolution equation. As such, these trajectory statistical solutions represent the probability distribution of all possible individual trajectories of the equation. At this abstract level, however, we do not specify the evolution equation, fixing only its corresponding set of solutions 𝒰\mathcal{U}.

Definition 3.1.

Let 𝒰\mathcal{U} be a subset of 𝒳\mathcal{X}. We say that a Borel probability measure ρ\rho on 𝒳\mathcal{X} is a 𝒰\mathcal{U}-trajectory statistical solution if

  1. (i)

    ρ\rho is tight;

  2. (ii)

    ρ\rho is carried by a Borel subset of 𝒳\mathcal{X} included in 𝒰\mathcal{U}, i.e. there exists 𝒱𝒳\mathcal{V}\in\mathcal{B}_{\mathcal{X}} such that 𝒱𝒰\mathcal{V}\subset\mathcal{U} and ρ(𝒳𝒱)=0\rho(\mathcal{X}\setminus\mathcal{V})=0.

From now on, we fix the following convention regarding notation. We use calligraphic capital letters to denote subsets of 𝒳\mathcal{X} (e.g. 𝒦\mathcal{K}, 𝒰\mathcal{U}, 𝒜\mathcal{A}, etc.), and plain capital letters for subsets of XX (e.g. KK, AA, etc.). The letters \mathcal{E} or Γ\Gamma are used as index sets of nets, where the indices are usually represented by the letters α\alpha, β\beta, ε\varepsilon, or γ\gamma.

We prove below two theorems on the convergence of trajectory statistical solutions, as described in more details in the Introduction. One is for arbitrary trajectory statistical solutions, suitable to approximations which are not necessarily well-posed, and the other is for approximations which have a well-defined solution semigroup.

The first result stems from the compactness of sets of uniformly tight Borel probability measures. The main point is to localize the carrier of the limit trajectory statistical solution. The second one simplifies the necessary conditions in the case there is a well-defined solution semigroup.

Theorem 3.1.

Let XX be a Hausdorff space, II be an interval in \mathbb{R}, and {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} be a family of 𝒰ε\mathcal{U}_{\varepsilon}-trajectory statistical solutions on subsets 𝒰ε𝒳\mathcal{U}_{\varepsilon}\subset\mathcal{X}, carried by Borel subsets 𝒱ε𝒳\mathcal{V}_{\varepsilon}\subset\mathcal{X}, where 𝒳=𝒞loc(I,X)\mathcal{X}=\mathcal{C}_{\textrm{\rm loc}}(I,X). Let 𝒰𝒳\mathcal{U}\subset\mathcal{X} and suppose there is a sequence {𝒦n}n\{\mathcal{K}_{n}\}_{n\in\mathbb{N}} of compact sets in 𝒳\mathcal{X} such that, for all nn\in\mathbb{N},

  1. (A1)

    ρε(𝒳𝒦n)<δn\rho_{\varepsilon}(\mathcal{X}\setminus\mathcal{K}_{n})<\delta_{n}, for all ε\varepsilon\in\mathcal{E}, with δn0\delta_{n}\rightarrow 0;

  2. (A2)

    lim supε(𝒱ε𝒦n)𝒰\limsup_{\varepsilon\in\mathcal{E}}(\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})\subset\mathcal{U} with respect to the topological superior limit in 𝒳\mathcal{X}.

Then, there exists a 𝒰\mathcal{U}-trajectory statistical solution ρ\rho which is a weak-star semicontinuity limit of a subnet of {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}}. Moreover, if the interval II is closed and bounded on the left with left endpoint t0t_{0} and Πt0ρεwscμ0\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} for some tight Borel probability measure μ0\mu_{0} on XX, then Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

Proof.

From the assumption that ρε(𝒳𝒦n)<δn\rho_{\varepsilon}(\mathcal{X}\setminus\mathcal{K}_{n})<\delta_{n}, for all ε\varepsilon\in\mathcal{E}, with compact sets 𝒦n𝒳\mathcal{K}_{n}\subset\mathcal{X} and δn0\delta_{n}\rightarrow 0, it follows immediately that {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} is a uniformly tight family of probability measures in (𝒳,tight)\mathcal{M}(\mathcal{X},\rm{tight}). Therefore, from the compactness result in Theorem 2.1, there exists a subnet of {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} which converges in the weak-star semicontinuity topology to a probability measure ρ(𝒳,tight)\rho\in\mathcal{M}(\mathcal{X},\rm{tight}). To deduce that ρ\rho is a 𝒰\mathcal{U}-trajectory statistical solution, it remains to show that ρ\rho is carried by a Borel subset of 𝒰\mathcal{U}, which is the main component of this proof.

Using the carriers 𝒱ε𝒰ε\mathcal{V}_{\varepsilon}\subset\mathcal{U}_{\varepsilon} of each 𝒰ε\mathcal{U}_{\varepsilon}-trajectory statistical solution, define the set

𝒱=n=1lim supε(𝒱ε𝒦n)=n=1εεγ𝒱γ𝒦n¯,\mathcal{V}=\bigcup_{n=1}^{\infty}\limsup_{\varepsilon\in\mathcal{E}}(\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})=\bigcup_{n=1}^{\infty}\bigcap_{\varepsilon\in\mathcal{E}}\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}, (3.1)

cf. (2.3). Thanks to condition (A2), it follows that 𝒱𝒰\mathcal{V}\subset\mathcal{U}. Moreover, since an arbitrary intersection of closed sets is closed and since each set εγ𝒱γ𝒦n¯\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}, ε\varepsilon\in\mathcal{E}, is a closed subset of the compact subset 𝒦n\mathcal{K}_{n}, the sets εεγ𝒱γ𝒦n¯\bigcap_{\varepsilon\in\mathcal{E}}\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}, nn\in\mathbb{N}, are compact, and 𝒱\mathcal{V} is a σ\sigma-compact subset of 𝒳\mathcal{X}, hence Borel.

We now claim that ρ\rho is carried by 𝒱\mathcal{V}. Since \mathcal{E} is a directed set, then given any β\beta\in\mathcal{E} there exists α\alpha\in\mathcal{E} such that βα\beta\preccurlyeq\alpha and εα\varepsilon\preccurlyeq\alpha. For any such α\alpha, we have, by assumption,

ρα(εγ𝒱γ𝒦n¯)ρα(𝒱α𝒦n)=ρα(𝒦n)1δn,\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq\rho_{\alpha}\left(\mathcal{V}_{\alpha}\cap\mathcal{K}_{n}\right)=\rho_{\alpha}(\mathcal{K}_{n})\geq 1-\delta_{n},

for any given n.n\in\mathbb{N}. Taking the supremum in α\alpha, we find that

supβαρα(εγ𝒱γ𝒦n¯)supβα,εαρα(εγ𝒱γ𝒦n¯)1δn.\sup_{\beta\preccurlyeq\alpha}\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq\sup_{\beta\preccurlyeq\alpha,\,\varepsilon\preccurlyeq\alpha}\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq 1-\delta_{n}.

Since β\beta\in\mathcal{E} above is arbitrary, we take the infimum of this expression over β\beta\in\mathcal{E} and find that

lim supβρβ(εγ𝒱γ𝒦n¯)1δn.\limsup_{\beta\in\mathcal{E}}\rho_{\beta}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq 1-\delta_{n}.

In view of Lemma 2.2, statement (iv), and the fact that ρβwscρ\rho_{\beta}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\rho in 𝒳\mathcal{X}, we obtain that

ρ(εγ𝒱γ𝒦n¯)lim supβρβ(εγ𝒱γ𝒦n¯)1δn,for all ε and n.\displaystyle\rho\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq\limsup_{\beta\in\mathcal{E}}\rho_{\beta}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq 1-\delta_{n},\quad\mbox{for all }\varepsilon\in\mathcal{E}\mbox{ and }n\in\mathbb{N}. (3.2)

Clearly, if ε1ε2\varepsilon_{1}\preccurlyeq\varepsilon_{2} then ε2γ𝒱γ𝒦n¯ε1γ𝒱γ𝒦n¯\overline{\bigcup_{\varepsilon_{2}\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\subset\overline{\bigcup_{\varepsilon_{1}\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}. Thus, for each nn\in\mathbb{N}, the net {εγ𝒱γ𝒦n¯}ε\left\{\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right\}_{\varepsilon\in\mathcal{E}} is a monotone decreasing net of compact sets in the compact Hausdorff space 𝒦n\mathcal{K}_{n}. Moreover, since ρ\rho is a tight Borel probability measure on a Hausdorff space then ρ\rho is outer regular, see Section 2.2. Hence, Lemma 2.1 applies, and we deduce that

ρ(lim supε(𝒱ε𝒦n))=ρ(εεγ𝒱γ𝒦n¯)=limερ(εγ𝒱γ𝒦n¯)1δn\displaystyle\rho\left(\limsup_{\varepsilon\in\mathcal{E}}(\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})\right)=\rho\left(\bigcap_{\varepsilon\in\mathcal{E}}\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)=\lim_{\varepsilon\in\mathcal{E}}\rho\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}\mathcal{V}_{\gamma}\cap\mathcal{K}_{n}}\right)\geq 1-\delta_{n}

for all nn\in\mathbb{N}.

Hence, from (3.1),

ρ(𝒱)=ρ(n=1lim supε(𝒱ε𝒦n))ρ(lim supε(𝒱ε𝒦n))1δn, for all n.\rho(\mathcal{V})=\rho\left(\bigcup_{n=1}^{\infty}\limsup_{\varepsilon}(\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})\right)\geq\rho\left(\limsup_{\varepsilon\in\mathcal{E}}(\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})\right)\geq 1-\delta_{n},\quad\mbox{ for all }n\in\mathbb{N}.

Since δn0\delta_{n}\rightarrow 0, by taking nn\to\infty we find that ρ(𝒱)=1\rho(\mathcal{V})=1. This completes the proof that the Borel set 𝒱𝒰\mathcal{V}\subset\mathcal{U} carries ρ\rho and that ρ\rho is a 𝒰\mathcal{U}-trajectory statistical solution.

For the second part of the statement, let us now suppose that II is closed and bounded on the left, with left endpoint t0t_{0}, and that Πt0ρεwscμ0\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}. Since Πt0:𝒳X\Pi_{t_{0}}:\mathcal{X}\to X is a continuous mapping, the set Kn=Πt0𝒦nK_{n}=\Pi_{t_{0}}\mathcal{K}_{n} is compact in XX and Πt0Kn=Πt01Πt0𝒦n𝒦n\Pi_{t_{0}}K_{n}=\Pi_{t_{0}}^{-1}\Pi_{t_{0}}\mathcal{K}_{n}\supset\mathcal{K}_{n}, so that

Πt0ρε(XKn)=ρε(Πt01(XKn))=ρε(𝒳Πt01Kn)ρε(𝒳𝒦n)<δn,\Pi_{t_{0}}\rho_{\varepsilon}(X\setminus K_{n})=\rho_{\varepsilon}(\Pi_{t_{0}}^{-1}(X\setminus K_{n}))=\rho_{\varepsilon}(\mathcal{X}\setminus\Pi_{t_{0}}^{-1}K_{n})\leq\rho_{\varepsilon}(\mathcal{X}\setminus\mathcal{K}_{n})<\delta_{n},

showing that Πt0ρε\Pi_{t_{0}}\rho_{\varepsilon} is tight on XX. Similarly, Πt0ρ\Pi_{t_{0}}\rho is also tight on XX. Let us show that Πt0ρεwscΠt0ρ\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\Pi_{t_{0}}\rho on XX using condition (v) of Lemma 2.2.

First, for the whole space XX, since Πt01X=𝒳\Pi_{t_{0}}^{-1}X=\mathcal{X}, we have

Πt0ρε(X)=ρε(Πt01X)=ρε(𝒳)ρ(𝒳)=ρ(Πt01X)=Πt0ρ(X).\Pi_{t_{0}}\rho_{\varepsilon}(X)=\rho_{\varepsilon}(\Pi_{t_{0}}^{-1}X)=\rho_{\varepsilon}(\mathcal{X})\rightarrow\rho(\mathcal{X})=\rho(\Pi_{t_{0}}^{-1}X)=\Pi_{t_{0}}\rho(X).

Now, for an open set GXG\subset X, the set Πt01G\Pi_{t_{0}}^{-1}G is open in 𝒳\mathcal{X}, so that

lim infεΠt0ρε(G)=lim infερε(Πt01(G))ρ(Πt01(G))=Πt0ρ(G).\liminf_{\varepsilon}\Pi_{t_{0}}\rho_{\varepsilon}(G)=\liminf_{\varepsilon}\rho_{\varepsilon}(\Pi_{t_{0}}^{-1}(G))\geq\rho(\Pi_{t_{0}}^{-1}(G))=\Pi_{t_{0}}\rho(G).

Thus, from Lemma 2.2, we deduce that Πt0ρεwscΠt0ρ\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\Pi_{t_{0}}\rho on XX.

On the other hand, by assumption, we have Πt0ρεwscμ0\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}. Since (X,tight)\mathcal{M}(X,\rm{tight}) is Hausdorff with respect to the weak-star semicontinuity topology (see Section 2.3), the weak-star semicontinuity limit has to be unique. Thus, we deduce that Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}. This concludes the proof. ∎

Remark 3.1.

The hypothesis, in Theorem 3.1, of the existence of compact subsets 𝒦n𝒳\mathcal{K}_{n}\subset\mathcal{X} with ρε(𝒳𝒦n)<δn\rho_{\varepsilon}(\mathcal{X}\setminus\mathcal{K}_{n})<\delta_{n} and δn0\delta_{n}\rightarrow 0 is, in fact, the condition of uniform tightness of the family {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}}, and this by itself guarantees, from the compactness given in Theorem 2.1, the existence of the weak limit ρ\rho. The importance of making this condition explicit in the statement of the theorem is only to relate the sets 𝒦n\mathcal{K}_{n} to the set 𝒰\mathcal{U}, via the condition that lim supε(𝒱ε𝒦n)𝒰\limsup_{\varepsilon\in\mathcal{E}}(\,\mathcal{V}_{\varepsilon}\cap\mathcal{K}_{n})\subset\mathcal{U}, with respect to the topological superior limit in 𝒳\mathcal{X}. With this extra condition, we prove that the limit measure ρ\rho is carried by a Borel set included in 𝒰\mathcal{U}, and hence ρ\rho is a 𝒰\mathcal{U}-trajectory statistical solution. The localization of the carrier of ρ\rho is in fact the main point of Theorem 3.1, since the existence of a limit measure ρ\rho follows directly from the underlying uniform tightness condition.

Remark 3.2.

From the proof of Theorem 3.1, we have, more precisely, that ρ\rho is carried by the σ\sigma-compact Borel subset 𝒱\mathcal{V} defined in (3.1), which depends on the sets 𝒱ε\mathcal{V}_{\varepsilon}, ε\varepsilon\in\mathcal{E}, and also the sequence of compact sets 𝒦n𝒳\mathcal{K}_{n}\subset\mathcal{X}, nn\in\mathbb{N}. In this regard, we note that a carrier set is in general not unique.

We turn to our second main result, relating to applications where the net of approximating trajectory statistical solutions is induced by a well-defined solution operator SεS_{\varepsilon}, ε\varepsilon\in\mathcal{E}.

Theorem 3.2.

Let XX be a Hausdorff space and let II be an interval in \mathbb{R} closed and bounded on the left with left endpoint t0t_{0}. Let 𝒰\mathcal{U} be a subset of 𝒳=𝒞loc(I,X)\mathcal{X}=\mathcal{\mathcal{C}_{\textrm{\rm loc}}}(I,X). Consider a net {Sε}ε\{S_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} of measurable functions Sε:X𝒳S_{\varepsilon}:X\rightarrow\mathcal{X} and a net {με}ε\{\mu_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} of tight Borel probability measures on XX. Set 𝒰ε=Sε(X)\mathcal{U}_{\varepsilon}=S_{\varepsilon}(X) and define Pε=Πt0Sε:XXP_{\varepsilon}=\Pi_{t_{0}}S_{\varepsilon}:X\rightarrow X. Assume that

  1. (H1)

    Pεμεwscμ0P_{\varepsilon}\mu_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}, for some tight Borel probability measure μ0\mu_{0} on XX.

Moreover, suppose that there exists a sequence {Kn}n\{K_{n}\}_{n\in\mathbb{N}} of compact sets in XX with the following properties:

  1. (H2)

    The map Sε|Kn:Kn𝒳S_{\varepsilon}|_{K_{n}}:K_{n}\rightarrow\mathcal{X} is continuous for all nn\in\mathbb{N} and all ε\varepsilon\in\mathcal{E}, with the topology inherited from XX;

  2. (H3)

    με(XKn)δn\mu_{\varepsilon}(X\setminus K_{n})\leq\delta_{n}, for all ε\varepsilon\in\mathcal{E} and all nn\in\mathbb{N}, where δn0,\delta_{n}\rightarrow 0, when nn\rightarrow\infty;

  3. (H4)

    For each nn, there exists a compact set 𝒦n\mathcal{K}_{n} in 𝒳\mathcal{X} with Sε(Kn)𝒦nS_{\varepsilon}(K_{n})\subset\mathcal{K}_{n}, for all ε\varepsilon\in\mathcal{E}.

  4. (H5)

    For each nn\in\mathbb{N},

    lim supεSε(Kn)𝒰.\limsup_{\varepsilon\in\mathcal{E}}S_{\varepsilon}(K_{n})\subset\mathcal{U}.

Then, each ρε=Sεμε\rho_{\varepsilon}=S_{\varepsilon}\mu_{\varepsilon} is a 𝒰ε\mathcal{U}_{\varepsilon}-trajectory statistical solution and the net {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} has a convergent subnet, with respect to the weak-star semicontinuity topology, to a 𝒰\mathcal{U}-trajectory statistical solution ρ\rho such that Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

Remark 3.3.

Note that, given {𝐮0,ε}εX\{\mathbf{u}_{0,\varepsilon}\}_{\varepsilon\in\mathcal{E}}\subset X, the family Pε𝐮0,ε=Πt0Sε𝐮0,εP_{\varepsilon}\mathbf{u}_{0,\varepsilon}=\Pi_{t_{0}}S_{\varepsilon}\mathbf{u}_{0,\varepsilon} represents a collection of approximating initial data associated with the solution operators {Sε}ε\{S_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} in a given application. Hence, condition (H1) is a natural assumption that guarantees that the initial measures Πt0Sεμε\Pi_{t_{0}}S_{\varepsilon}\mu_{\varepsilon} for the approximations converge to the limit initial measure. This should be checked in each application. In some cases, one has simply Πt0Sε=I\Pi_{t_{0}}S_{\varepsilon}=I, i.e. the identity operator, as in the application to the inviscid limit of the 2D Navier-Stokes equations given in Section 4.2. However, this is not the case in certain applications, for example, when the approximating systems evolve in some lower-dimensional approximation of XX, as in the application to Galerkin approximations of the Navier-Stokes equations presented in Section 4.4. Indeed, in the Galerkin case we have Πt0Sεu=PNu\Pi_{t_{0}}S_{\varepsilon}u=P_{N}u, uXu\in X, with PNP_{N} denoting the projection onto the space spanned by NN\in\mathbb{N} basis vectors of the vector space XX.

Remark 3.4.

Condition (H2) pertains to the regularity of the approximation semigroup and is used for measurability purposes, guaranteeing that ρε\rho_{\varepsilon} is carried by a Borel subset of 𝒰ε\mathcal{U}_{\varepsilon} and, hence, is a statistical solution. Condition (H3) guarantees that the initial measures με\mu_{\varepsilon} are uniformly exhausted by the sets KnK_{n}, which is then used together with (H4) to prove the uniform tightness of ρε=Sεμε\rho_{\varepsilon}=S_{\varepsilon}\mu_{\varepsilon} with respect to the compact sets {𝒦n}\{\mathcal{K}_{n}\}. Condition (H5) certifies that the family of operators are, in fact, an approximation of the limit problem, i.e. the approximations do converge to a solution of the limit problem, as expressed by the space 𝒰\mathcal{U}, so that the limit measure ρ\rho is a 𝒰\mathcal{U}-statistical solution.

Remark 3.5.

In the particular case where PεP_{\varepsilon} is the identity operator and με=μ0\mu_{\varepsilon}=\mu_{0} for all ε\varepsilon\in\mathcal{E}, for some μ0\mu_{0} tight Borel probability measure, assumptions (H1) and (H3) of Theorem 3.2 are immediately satisfied. Several applications would fit into this setting. This is indeed the case in our application to the inviscid limit of the 2D Navier-Stokes equations in Section 4.2.

Proof of Theorem 3.2.

From the definition of ρε\rho_{\varepsilon} as SεμεS_{\varepsilon}\mu_{\varepsilon}, we have

ρε(𝒳Sε(Kn))=με(Sε1(𝒳Sε(Kn)))=με(XSε1(Sε(Kn))).\rho_{\varepsilon}(\mathcal{X}\setminus S_{\varepsilon}(K_{n}))=\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{X}\setminus S_{\varepsilon}(K_{n})))=\mu_{\varepsilon}(X\setminus S_{\varepsilon}^{-1}(S_{\varepsilon}(K_{n}))).

Since Sε1(Sε(Kn))KnS_{\varepsilon}^{-1}(S_{\varepsilon}(K_{n}))\supset K_{n}, it follows that XSε1(Sε(Kn))XKnX\setminus S_{\varepsilon}^{-1}(S_{\varepsilon}(K_{n}))\subset X\setminus K_{n}, so that

ρε(𝒳Sε(Kn))με(XKn).\rho_{\varepsilon}(\mathcal{X}\setminus S_{\varepsilon}(K_{n}))\leq\mu_{\varepsilon}(X\setminus K_{n}).

Using (H3), we obtain

ρε(𝒳Sε(Kn))δn,\rho_{\varepsilon}(\mathcal{X}\setminus S_{\varepsilon}(K_{n}))\leq\delta_{n}, (3.3)

for all nn\in\mathbb{N} and arbitrary ε\varepsilon. Thus,

ρε(𝒳mSε(Km))ρε(𝒳Sε(Kn))δn,\rho_{\varepsilon}(\mathcal{X}\setminus\cup_{m\in\mathbb{N}}S_{\varepsilon}(K_{m}))\leq\rho_{\varepsilon}(\mathcal{X}\setminus S_{\varepsilon}(K_{n}))\leq\delta_{n},

for all nn. Since δn0\delta_{n}\rightarrow 0, this implies that

ρε(𝒳nSε(Kn))=0.\rho_{\varepsilon}(\mathcal{X}\setminus\cup_{n\in\mathbb{N}}S_{\varepsilon}(K_{n}))=0.

Therefore, ρε\rho_{\varepsilon} is carried by the set

𝒱ε=nSε(Kn).\mathcal{V}_{\varepsilon}=\bigcup_{n\in\mathbb{N}}S_{\varepsilon}(K_{n}). (3.4)

From (H2) we see that Sε(Kn)S_{\varepsilon}(K_{n}) is compact in 𝒳\mathcal{X}, so that 𝒱ε\mathcal{V}_{\varepsilon} is a σ\sigma-compact set in 𝒳\mathcal{X}. In particular, it is a Borel set in 𝒳\mathcal{X}.

Let us now show that each ρε\rho_{\varepsilon} is tight. Using (H3), we see that, for every Borel set 𝒜𝒳\mathcal{A}\subset\mathcal{X},

ρε(𝒜)\displaystyle\rho_{\varepsilon}(\mathcal{A}) =με(Sε1(𝒜))\displaystyle=\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A}))
=με(Sε1(𝒜)Kn)+με(Sε1(𝒜)Kn)\displaystyle=\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})+\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\setminus K_{n})
με(Sε1(𝒜)Kn)+με(XKn)\displaystyle\leq\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})+\mu_{\varepsilon}(X\setminus K_{n})
με(Sε1(𝒜)Kn)+δn.\displaystyle\leq\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})+\delta_{n}.

On the other hand,

με(Sε1(𝒜)Kn)με(Sε1(𝒜))=ρε(𝒜).\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\leq\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A}))=\rho_{\varepsilon}(\mathcal{A}).

Thus,

ρε(𝒜)=supn{με(Sε1(𝒜)Kn)}.\rho_{\varepsilon}(\mathcal{A})=\sup_{n}\{\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\}. (3.5)

Being tight, each με\mu_{\varepsilon} is continuous from below with respect to compact sets. Thus,

με(Sε1(𝒜)Kn)=sup{με(F):FSε1(𝒜)Kn,F compact in X}.\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})=\sup\{\mu_{\varepsilon}(F):F\subset S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n},F\mbox{ compact in }X\}.

Since FSε1(Sε(F))F\subset S_{\varepsilon}^{-1}(S_{\varepsilon}(F)), we have the bound

με(Sε1(𝒜)Kn)sup{με(Sε1(Sε(F))):FSε1(𝒜)Kn,F compact in X}.\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\leq\sup\{\mu_{\varepsilon}(S_{\varepsilon}^{-1}(S_{\varepsilon}(F))):F\subset S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n},F\mbox{ compact in }X\}.

From the continuity of the restriction Sε|KnS_{\varepsilon}|_{K_{n}} of SεS_{\varepsilon} to the compact set KnK_{n}, we have Sε(F)S_{\varepsilon}(F) compact, for every compact set FKnF\subset K_{n}. Thus, we bound the right hand side extending Sε(F)S_{\varepsilon}(F) to any compact set 𝒦𝒜\mathcal{K}^{\prime}\subset\mathcal{A}, i.e.

με(Sε1(𝒜)Kn)sup{με(Sε1(𝒦)):𝒦𝒜,𝒦 compact in 𝒳}.\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\leq\sup\{\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{K}^{\prime})):\mathcal{K}^{\prime}\subset\mathcal{A},\,\mathcal{K}^{\prime}\mbox{ compact in }\mathcal{X}\}.

Back to ρε=Sεμε\rho_{\varepsilon}=S_{\varepsilon}\mu_{\varepsilon}, this means

με(Sε1(𝒜)Kn)sup{ρε(𝒦):𝒦𝒜,𝒦 compact in 𝒳}.\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\leq\sup\{\rho_{\varepsilon}(\mathcal{K}^{\prime}):\mathcal{K}^{\prime}\subset\mathcal{A},\,\mathcal{K}^{\prime}\mbox{ compact in }\mathcal{X}\}.

Since the right hand side does not depend on nn, this gives

supnμε(Sε1(𝒜)Kn)sup{ρε(𝒦):𝒦𝒜,𝒦 compact in 𝒳}.\sup_{n}\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\leq\sup\{\rho_{\varepsilon}(\mathcal{K}^{\prime}):\mathcal{K}^{\prime}\subset\mathcal{A},\,\mathcal{K}^{\prime}\mbox{ compact in }\mathcal{X}\}.

Plugging this back into (3.5) yields

ρε(𝒜)supn{με(Sε1(𝒜)Kn)}sup{ρε(𝒦):𝒦𝒜,𝒦 compact in 𝒳}ρε(𝒜).\rho_{\varepsilon}(\mathcal{A})\leq\sup_{n}\{\mu_{\varepsilon}(S_{\varepsilon}^{-1}(\mathcal{A})\cap K_{n})\}\leq\sup\{\rho_{\varepsilon}(\mathcal{K}^{\prime}):\mathcal{K}^{\prime}\subset\mathcal{A},\,\mathcal{K}^{\prime}\mbox{ compact in }\mathcal{X}\}\leq\rho_{\varepsilon}(\mathcal{A}).

In other words,

ρε(𝒜)=sup{ρε(𝒦):𝒦𝒜,𝒦 compact in 𝒳}.\rho_{\varepsilon}(\mathcal{A})=\sup\{\rho_{\varepsilon}(\mathcal{K}^{\prime}):\mathcal{K}^{\prime}\subset\mathcal{A},\,\mathcal{K}^{\prime}\mbox{ compact in }\mathcal{X}\}.

This shows that ρε\rho_{\varepsilon} is tight. Thus, ρε\rho_{\varepsilon} is a tight measure carried by the Borel subset 𝒱ε\mathcal{V}_{\varepsilon} of 𝒰ε\mathcal{U}_{\varepsilon}, which means that ρε\rho_{\varepsilon} is a 𝒰ε\mathcal{U}_{\varepsilon}-statistical solution.

Now, using (H4) and the previous estimate (3.3), we see that

ρε(𝒳𝒦n)ρε(𝒳Sε(Kn))δn.\rho_{\varepsilon}(\mathcal{X}\setminus\mathcal{K}_{n})\leq\rho_{\varepsilon}(\mathcal{X}\setminus S_{\varepsilon}(K_{n}))\leq\delta_{n}.

This implies that the family {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} is uniformly tight.

Since {ρε}ε\{\rho_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} is uniformly tight, it follows from Theorem 2.1 that there exists a subnet which converges in the weak-star semicontinuity topology to a tight probability measure ρ(𝒳,tight)\rho\in\mathcal{M}(\mathcal{X},\rm{tight}). Now we show directly that ρ\rho is carried by

𝒱=n=1lim supεSε(Kn).\mathcal{V}=\bigcup_{n=1}^{\infty}\limsup_{\varepsilon\in\mathcal{E}}S_{\varepsilon}(K_{n}).

Note that, by the definition (2.3) of lim supε\limsup_{\varepsilon} as an intersection of closed sets, we have that lim supεSε(Kn)\limsup_{\varepsilon}S_{\varepsilon}(K_{n}) is a Borel set and so is 𝒱\mathcal{V}. Moreover, by assumption (H5), each lim supεSε(Kn)\limsup_{\varepsilon}S_{\varepsilon}(K_{n}) is a subset of 𝒰\mathcal{U}, hence 𝒱𝒰\mathcal{V}\subset\mathcal{U} as well. Let us now mimic the proof in Theorem 3.1 and show that ρ\rho is carried by 𝒱\mathcal{V}.

From the estimate (3.3), we find that

ρε(Sε(Kn))1δn.\rho_{\varepsilon}(S_{\varepsilon}(K_{n}))\geq 1-\delta_{n}.

For every β\beta\in\mathcal{E}, there exists α\alpha\in\mathcal{E} such that βα\beta\preccurlyeq\alpha and εα\varepsilon\preccurlyeq\alpha. Thus,

ρα(εγSγ(Kn)¯)ρα(Sα(Kn))1δn,\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq\rho_{\alpha}(S_{\alpha}(K_{n}))\geq 1-\delta_{n},

for arbitrary n.n\in\mathbb{N}. Taking the supremum over α\alpha\in\mathcal{E}, for βα\beta\preccurlyeq\alpha,

supα,βαρα(εγSγ(Kn)¯)supα,βα,εαρα(εγSγ(Kn)¯)1δn.\sup_{\alpha\in\mathcal{E},\;\beta\preccurlyeq\alpha}\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq\sup_{\alpha\in\mathcal{E},\;\beta\preccurlyeq\alpha,\varepsilon\preccurlyeq\alpha}\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq 1-\delta_{n}.

Taking, now, the infimum over β\beta\in\mathcal{E},

lim supβρβ(εγSγ(Kn)¯)=infβsupα,βαρα(εγSγ(Kn)¯)1δn.\limsup_{\beta\in\mathcal{E}}\rho_{\beta}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)=\inf_{\beta\in\mathcal{E}}\sup_{\alpha\in\mathcal{E},\;\beta\preccurlyeq\alpha}\rho_{\alpha}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq 1-\delta_{n}.

Since ρβwscρ\rho_{\beta}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\rho along a subnet β\beta\in\mathcal{E}^{\prime}, it follows from Lemma 2.2, (iv), that

ρ(εγSγ(Kn)¯)lim supβρβ(εγSγ(Kn)¯)1δn.\rho\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq\limsup_{\beta\in\mathcal{E}^{\prime}}\rho_{\beta}\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq 1-\delta_{n}.

Each Sγ(Kn)S_{\gamma}(K_{n}) is included in the compact set 𝒦n\mathcal{K}_{n}, so that the set εγSγ(Kn)¯\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})} is a closed set in 𝒦n\mathcal{K}_{n}, hence compact. Moreover, the family of sets εγSγ(Kn)¯\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}, ε\varepsilon\in\mathcal{E}, is decreasing in ε\varepsilon. Hence, it follows from Lemma 2.1 that

ρ(lim supεSε(Kn))=ρ(εεγSγ(Kn)¯)=limερ(εγSγ(Kn)¯)1δn.\rho\left(\limsup_{\varepsilon\in\mathcal{E}}S_{\varepsilon}(K_{n})\right)=\rho\left(\bigcap_{\varepsilon\in\mathcal{E}}\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)=\lim_{\varepsilon\in\mathcal{E}}\rho\left(\overline{\bigcup_{\varepsilon\preccurlyeq\gamma}S_{\gamma}(K_{n})}\right)\geq 1-\delta_{n}.

Therefore,

ρ(𝒱)=ρ(mlim supεSε(Km))1δn.\rho(\mathcal{V})=\rho\left(\bigcup_{m\in\mathbb{N}}\limsup_{\varepsilon\in\mathcal{E}}S_{\varepsilon}(K_{m})\right)\geq 1-\delta_{n}.

Finally, since this holds for any nn\in\mathbb{N} and δn0\delta_{n}\rightarrow 0, we find that

ρ(𝒱)=1.\rho(\mathcal{V})=1.

Thus, we find a tight Borel probability measure ρ\rho with ρεwscρ\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\rho along a subnet ε\varepsilon\in\mathcal{E}^{\prime}, with ρ\rho carried by the Borel subset 𝒱\mathcal{V} of 𝒰\mathcal{U}, which means that ρ\rho is a 𝒰\mathcal{U}-trajectory statistical solution.

It remains to show that Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}. We have just proved that ρεwscρ\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\rho as measures on 𝒳\mathcal{X}. Since Πt0\Pi_{t_{0}} is continuous from 𝒳\mathcal{X} to XX, this implies, as seen in the proof of Theorem 3.1, that Πt0ρεwscΠt0ρ\Pi_{t_{0}}\rho_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\Pi_{t_{0}}\rho, and Πt0ρε,Πt0ρ(X,tight)\Pi_{t_{0}}\rho_{\varepsilon},\Pi_{t_{0}}\rho\in\mathcal{M}(X,\rm{tight}).

On the other hand, from hypothesis (H1), we obtain

Πt0ρε=Πt0Sεμε=Pεμεwscμ0.\Pi_{t_{0}}\rho_{\varepsilon}=\Pi_{t_{0}}S_{\varepsilon}\mu_{\varepsilon}=P_{\varepsilon}\mu_{\varepsilon}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}.

Combining the two limits together and invoking the uniqueness of the weak-star semicontinuity limit in (X,tight)\mathcal{M}(X,\rm{tight}), it follows that Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}, which completes the proof. ∎

Remark 3.6.

Notice we do not apply Theorem 3.1 to prove Theorem 3.2. If instead we assume lim supε(𝒰ε𝒦n)𝒰\limsup_{\varepsilon\in\mathcal{E}}(\mathcal{U}_{\varepsilon}\cap\mathcal{K}_{n})\subset\mathcal{U} in place of (H5), then the hypothesis (A2) of Theorem 3.1 is satisfied. However, condition (H5) is weaker and, in fact, more natural in this context. For this reason, we prove hypothesis (A1) of Theorem 3.1 and then mimic the remaining part of the proof of Theorem 3.1 to complete the proof Theorem 3.2.

4. Applications

This section provides applications of our general framework for convergence of statistical solutions from Theorem 3.1 and Theorem 3.2 above. Section 4.2 and Section 4.3 concern the inviscid limit of the Navier-Stokes equations towards the Euler equations in two and three dimensions, respectively. And Section 4.4 deals with spectral Galerkin discretizations approximating the 3D Navier-Stokes equations. Before delving into these applications, we first recall in Section 4.1 some preliminary background regarding the Euler and Navier-Stokes equations.

4.1. Mathematical setting for 2D and 3D incompressible flows

We consider the dd-dimensional incompressible Navier-Stokes equations (NSE) for either d=2d=2 or 33, given by

t𝐮νΔ𝐮+(𝐮)𝐮+p=𝐟,𝐮=0,\displaystyle\partial_{t}\mathbf{u}-\nu\Delta\mathbf{u}+(\mathbf{u}\cdot\nabla)\mathbf{u}+\nabla p=\mathbf{f},\quad\nabla\cdot\mathbf{u}=0, (4.1)

where 𝐮=(u1,,ud)\mathbf{u}=(u_{1},\ldots,u_{d}) and pp are the unknowns and represent the velocity field and the pressure, respectively. Moreover, 𝐟=(f1,,fd)\mathbf{f}=(f_{1},\ldots,f_{d}) represents a given body force applied to the fluid and ν>0\nu>0 is the kinematic viscosity. The functions 𝐮\mathbf{u}, pp and 𝐟\mathbf{f} depend on a spatial variable xx varying in Ωd\Omega\subset\mathbb{R}^{d} and on a time variable tt varying in an interval II\subset\mathbb{R}. We will refer to (4.1) as ‘ν\nu-NSE’ whenever there is a need to emphasize the dependence on ν\nu.

In the inviscid case, i.e. when ν=0\nu=0, (4.1) becomes the dd-dimensional incompressible Euler equations

t𝐮+(𝐮)𝐮+p=𝐟,𝐮=0.\displaystyle\partial_{t}\mathbf{u}+(\mathbf{u}\cdot\nabla)\mathbf{u}+\nabla p=\mathbf{f},\quad\nabla\cdot\mathbf{u}=0. (4.2)

We assume for simplicity that (4.1) and (4.2) are subject to periodic boundary conditions, with Ω=(0,L1)××(0,Ld)d\Omega=(0,L_{1})\times\ldots\times(0,L_{d})\subset\mathbb{R}^{d} denoting a basic domain of periodicity. We say that a function 𝐯:dd\mathbf{v}:\mathbb{R}^{d}\to\mathbb{R}^{d} is Ω\Omega-periodic if 𝐯\mathbf{v} is periodic with period LiL_{i} in each spatial direction xix_{i}, i=1,,di=1,\ldots,d.

Let us fix the functional setting associated to these equations. Denote by 𝒞per(Ω)d\mathcal{C}^{\infty}_{per}(\Omega)^{d} the space of infinitely differentiable and Ω\Omega-periodic functions defined on d\mathbb{R}^{d}, and let 𝒟σ𝒞per(Ω)d\mathcal{D}_{\sigma}\subset\mathcal{C}^{\infty}_{per}(\Omega)^{d} be the set of divergence-free and periodic test functions with vanishing spatial average, namely

𝒟σ={𝐮𝒞per(Ω)d:𝐮=0 and Ω𝐮(𝐱)d𝐱=0}.\displaystyle\mathcal{D}_{\sigma}=\left\{\mathbf{u}\in\mathcal{C}^{\infty}_{per}(\Omega)^{d}\,:\,\nabla\cdot\mathbf{u}=0\mbox{ and }\int_{\Omega}\mathbf{u}(\mathbf{x})\;{\text{\rm d}}\mathbf{x}=0\right\}. (4.3)

We denote by HH, VV and Wσ1,rW_{\sigma}^{1,r}, 1r1\leq r\leq\infty, the closures of 𝒟σ\mathcal{D}_{\sigma} with respect to the norms in L2(Ω)dL^{2}(\Omega)^{d}, H1(Ω)dH^{1}(\Omega)^{d} and W1,r(Ω)dW^{1,r}(\Omega)^{d}, respectively. Note that V=Wσ1,2V=W^{1,2}_{\sigma}. The inner product and norm in HH are defined, respectively, by

(𝐮,𝐯)=Ω𝐮𝐯d𝐱 and |𝐮|=(𝐮,𝐮),(\mathbf{u},\mathbf{v})=\int_{\Omega}\mathbf{u}\cdot\mathbf{v}\;{\text{\rm d}}\mathbf{x}\quad\mbox{ and }\quad|\mathbf{u}|=\sqrt{(\mathbf{u},\mathbf{u})},

where 𝐮𝐯=i=13uivi\mathbf{u}\cdot\mathbf{v}=\sum_{i=1}^{3}u_{i}v_{i}. In the space VV, these are defined as

((𝐮,𝐯))=(𝐮,𝐯)=Ω𝐮:𝐯d𝐱 and 𝐮=((𝐮,𝐮)),(\!(\mathbf{u},\mathbf{v})\!)=(\nabla\mathbf{u},\nabla\mathbf{v})=\int_{\Omega}\nabla\mathbf{u}:\nabla\mathbf{v}\;{\text{\rm d}}\mathbf{x}\quad\mbox{ and }\quad\|\mathbf{u}\|=\sqrt{(\!(\mathbf{u},\mathbf{u})\!)},

where it is understood that 𝐮=(xjui)i,j=1d\nabla\mathbf{u}=(\partial_{x_{j}}u_{i})_{i,j=1}^{d} and that 𝐮:𝐯\nabla\mathbf{u}:\nabla\mathbf{v} is the componentwise product between 𝐮\nabla\mathbf{u} and 𝐯\nabla\mathbf{v}. In the space Wσ1,rW_{\sigma}^{1,r}, except for r=2r=2, we can only define a norm, given by

𝐮Wσ1,r={(i,j=1dxjuiLrr)1r,1r<i,j=1dxjuiLr,r=.\|\mathbf{u}\|_{W_{\sigma}^{1,r}}=\left\{\begin{array}[]{ll}\left(\displaystyle\sum_{i,j=1}^{d}\|\partial_{x_{j}}u_{i}\|_{L^{r}}^{r}\right)^{\frac{1}{r}},&1\leq r<\infty\\ \\ \displaystyle\sum_{i,j=1}^{d}\|\partial_{x_{j}}u_{i}\|_{L^{r}},&r=\infty.\end{array}\right.

The fact that \|\cdot\| and Wσ1,r\|\cdot\|_{W_{\sigma}^{1,r}}, 1r1\leq r\leq\infty, are indeed norms follows from the Poincaré inequality (4.4) and the inequality (4.6) below.

We denote by HH^{\prime}, VV^{\prime} and (Wσ1,r)(W_{\sigma}^{1,r})^{\prime} the dual spaces of HH, VV and Wσ1,rW_{\sigma}^{1,r}, respectively. The dual spaces are endowed with the classical dual norm of Banach spaces. Namely, for a given Banach space EE, the standard norm in the dual space EE^{\prime} is given by 𝐮E=sup𝐯E1𝐮,𝐯E,E\|\mathbf{u}\|_{E^{\prime}}=\sup_{\|\mathbf{v}\|_{E}\leq 1}\langle\mathbf{u},\mathbf{v}\rangle_{E^{\prime},E}, where ,E,E\langle\cdot,\cdot\rangle_{E^{\prime},E} denotes the duality product between EE and EE^{\prime}. After identifying HH with its dual HH^{\prime}, we obtain VHVV\subset H\subset V^{\prime} and Wσ1,rH(Wσ1,r)W_{\sigma}^{1,r}\subset H\subset(W_{\sigma}^{1,r})^{\prime}, with the injections being continuous, compact, and each space dense in the following one. Also, since Ω\Omega is bounded, we have that Wσ1,rVW^{1,r}_{\sigma}\subset V, with continuous injection, for all r2r\geq 2.

The negative Laplacian operator (Δ)(-\Delta) on VH2(Ω)dV\cap H^{2}(\Omega)^{d} is a positive and self-adjoint operator with compact inverse. As such, it admits a nondecreasing sequence of positive eigenvalues {λk}k\{\lambda_{k}\}_{k\in\mathbb{N}} with λk\lambda_{k}\to\infty as kk\to\infty, which is associated to a sequence of eigenfunctions {𝐰k}k\{\mathbf{w}_{k}\}_{k\in\mathbb{N}} that consists of an orthonormal basis of HH. In relation to the first eigenvalue λ1\lambda_{1} of (Δ)(-\Delta), we have the Poincaré inequality,

|𝐮|λ11/2𝐮, for all 𝐮V.\displaystyle|\mathbf{u}|\leq\lambda_{1}^{-1/2}\|\mathbf{u}\|,\quad\mbox{ for all }\mathbf{u}\in V. (4.4)

Using Hölder’s inequality, we have

𝐮|Ω|(121r)𝐮Wσ1,r, for all 𝐮Wσ1,r,  2r,\displaystyle\|\mathbf{u}\|\leq|\Omega|^{\left(\frac{1}{2}-\frac{1}{r}\right)}\|\mathbf{u}\|_{W_{\sigma}^{1,r}},\quad\mbox{ for all }\mathbf{u}\in W_{\sigma}^{1,r},\,\,2\leq r\leq\infty, (4.5)

where |Ω|=L1Ld|\Omega|=L_{1}\cdots L_{d} is the area or volume of the dd-dimensional domain. For the sake of simplicity, and with the aim of using λ1\lambda_{1} for dimensional consistency, we write (4.5) in terms of λ1,\lambda_{1}, by introducing the non-dimensional constant c=max{1,|Ω|λ1d/2}1/2,c=\max\{1,|\Omega|\lambda_{1}^{d/2}\}^{1/2}, so that

𝐮cλ1d2(121r)𝐮Wσ1,r, for all 𝐮Wσ1,r,  2r.\displaystyle\|\mathbf{u}\|\leq c\lambda_{1}^{-\frac{d}{2}\left(\frac{1}{2}-\frac{1}{r}\right)}\|\mathbf{u}\|_{W_{\sigma}^{1,r}},\quad\mbox{ for all }\mathbf{u}\in W_{\sigma}^{1,r},\,\,2\leq r\leq\infty. (4.6)

For any normed space EE, we denote by BE(R)B_{E}(R) the closed ball centered at 0 and with radius R>0R>0 in EE. Moreover, we denote by EwE_{\text{w}} and BE(R)wB_{E}(R)_{\text{w}} the spaces EE and BE(R)B_{E}(R) endowed with the weak topology, respectively.

For d=2d=2, we denote by \nabla^{\perp} the operator defined as (x2,x1)(-\partial_{x_{2}},\partial_{x_{1}}). The vorticity ω=ω(x1,x2)\omega=\omega(x_{1},x_{2}) associated with a velocity field 𝐮(x1,x2)=(𝐮1(x1,x2),𝐮2(x1,x2))\mathbf{u}(x_{1},x_{2})=(\mathbf{u}_{1}(x_{1},x_{2}),\mathbf{u}_{2}(x_{1},x_{2})) is given by

ω=𝐮=x2𝐮1+x1𝐮2.\omega=\nabla^{\perp}\cdot\mathbf{u}=-\partial_{x_{2}}\mathbf{u}_{1}+\partial_{x_{1}}\mathbf{u}_{2}.

We recall that the LrL^{r}-norm of the vorticity controls the Wσ1,rW_{\sigma}^{1,r}-norm of its associated velocity field. More precisely, let 𝐮H\mathbf{u}\in H be such that 𝐮Lr(Ω)\nabla^{\perp}\cdot\mathbf{u}\in L^{r}(\Omega), for some 1<r<1<r<\infty. Then 𝐮Wσ1,r\mathbf{u}\in W_{\sigma}^{1,r} and

𝐮Wσ1,rc𝐮Lr,\|\mathbf{u}\|_{W_{\sigma}^{1,r}}\leq c\|\nabla^{\perp}\cdot\mathbf{u}\|_{L^{r}}, (4.7)

for some other positive constant cc.

We don’t need to track the different constants that appear in the estimates, so, in what follows, we denote as cc a dimensionless positive constant whose value may change from line to line. We also occasionally use the capital letter CC to denote a positive dimensional constant.

Before proceeding to Section 4.1.1 and Section 4.1.2 with the types of solutions of NSE and Euler that suit our purposes, let us briefly provide some context for the choice of such solutions and recall some of the currently available results on existence and uniqueness. We keep the discussion restricted to the case of periodic boundary conditions, although similar results are often valid with other types of boundary conditions. We refer to the references cited below for further details.

Regarding the NSE in two dimensions, it is well known that given any forcing term 𝐟Lloc2([t0,),V)\mathbf{f}\in L^{2}_{\textrm{loc}}([t_{0},\infty),V^{\prime}), for some t00t_{0}\geq 0, and initial datum 𝐮0\mathbf{u}_{0} in H,H, there exists a unique weak solution 𝐮\mathbf{u} of (4.1) on [t0,)[t_{0},\infty) satisfying 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}, see e.g. [21, 51, 62]. Here the exact meaning of “weak solution” is recalled in Definition 4.1 below. Therefore, the initial-value problem for weak solutions of the 2D ν\nu-NSE is globally well-posed, and we may thus define a solution operator SνS_{\nu} associating to each 𝐮0H\mathbf{u}_{0}\in H the corresponding unique solution 𝐮\mathbf{u} of (4.1) on [t0,)[t_{0},\infty) satisfying 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}.

In the three-dimensional case, it is known that for any given 𝐟Lloc2([t0,),V)\mathbf{f}\in L^{2}_{\textrm{loc}}([t_{0},\infty),V^{\prime}) and 𝐮0H\mathbf{u}_{0}\in H there exists a Leray-Hopf weak solution of NSE (cf. Definition 4.3 below) on [t0,)[t_{0},\infty) satisfying the initial condition 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}. This solution is typically obtained as an appropriate limit of the unique solutions of a corresponding sequence of approximating Galerkin systems, see e.g. [21, 51, 62]. However, this Leray-Hopf weak solution is not currently known to be unique, and hence a corresponding solution operator cannot be defined as of yet. Regarding this uniqueness issue, it is worth pointing out the recent result in [1] where the authors show that for a suitably constructed non-smooth forcing function 𝐟\mathbf{f} there exist two distinct Leray-Hopf weak solutions of 3D NSE in 3×(0,T)\mathbb{R}^{3}\times(0,T) with initial data 𝐮00\mathbf{u}_{0}\equiv 0, for some T>0T>0. Additionally, non-uniqueness results for weak solutions of 3D NSE of non-Leray-Hopf type were proved in [15, 14, 19] by using convex integration techniques.

For our applications, we focus on the notions of weak solutions of 2D NSE and Leray-Hopf weak solutions of 3D NSE as recalled in Definition 4.1 and Definition 4.3 below, respectively. In view of the aforementioned results, the examples showing convergence of statistical solutions in the 2D inviscid limit (Section 4.2) and for Galerkin approximations in 3D (Section 4.4) follow as a consequence of Theorem 3.2. The 3D inviscid limit case (Section 4.3), on the other hand, requires the setting of Theorem 3.1.

Specifically, for the Galerkin application, we take each SεS_{\varepsilon} from Theorem 3.2 to be the solution operator SNS_{N} for the Galerkin system with NN\in\mathbb{N} Galerkin modes (see (4.4)), and 𝒰\mathcal{U} as the set 𝒰Iν\mathcal{U}^{\nu}_{I} of Leray-Hopf weak solutions of 3D ν\nu-NSE on a fixed time interval II\subset\mathbb{R}.

Regarding the inviscid limit examples, in the 2D case we take each SεS_{\varepsilon} from Theorem 3.2 as the solution operator SνS_{\nu} associated to the 2D Navier-Stokes equations with viscosity parameter ν>0\nu>0 (see (4.2)). In the 3D inviscid limit case, we consider each set 𝒰ε\mathcal{U}_{\varepsilon} from Theorem 3.1 to be the family 𝒰Iν\mathcal{U}^{\nu}_{I} of Leray-Hopf weak solutions of 3D ν\nu-NSE on the time interval II\subset\mathbb{R}, and ρε=ρν\rho_{\varepsilon}=\rho_{\nu} as a corresponding trajectory statistical solution in the sense of Definition 3.1. We note that existence of such trajectory statistical solution ρν\rho_{\nu} in 3D satisfying a given initial condition Πt0ρν=μ0\Pi_{t_{0}}\rho_{\nu}=\mu_{0}, for any Borel probability measure μ0\mu_{0} on HH, follows from the work [40], but is also obtained in [13, Theorem 4.2] with a more streamlined proof. Additionally, as pointed out in Remark 4.4 below, this existence result also follows via convergence of statistical solutions of corresponding Galerkin approximations, as a consequence of our application in Section 4.4.

To complete the setup for the 2D and 3D inviscid limit applications as required from Theorem 3.1 and Theorem 3.2, respectively, it remains to choose an appropriate set 𝒰\mathcal{U} of solutions of the Euler equations (4.2). In view of assumption (A2) from Theorem 3.1 or (H5) in Theorem 3.2, we must choose a set 𝒰\mathcal{U} for which it holds that any vanishing viscosity convergent sequence of individual solutions 𝐮νj\mathbf{u}_{\nu_{j}} in 𝒰νj\mathcal{U}_{\nu_{j}}, jj\in\mathbb{N}, lying in a certain compact set, has as its limit a solution in 𝒰\mathcal{U}. In the 2D periodic case, this inviscid limit result for individual solutions is known to hold with respect to the standard notions of weak solutions of NSE and Euler (cf. Definition 4.2), provided enough regularity is assumed on the initial data. More precisely, given any 𝐮0H\mathbf{u}_{0}\in H such that 𝐮0Lr\nabla^{\perp}\cdot\mathbf{u}_{0}\in L^{r}, 1<r1<r\leq\infty, a vanishing viscosity convergent sequence of weak solutions to the 2D NSE, each with initial datum 𝐮0\mathbf{u}_{0}, has as its limit a weak solution of 2D Euler with the same initial datum. This is shown in [54] (see also [25]) under the assumption of zero forcing term, but it is mentioned that more general forcing terms could also be considered. See 4.2 below for one such more general case.

Here we recall that in the case r=r=\infty, namely when 𝐮0L\nabla^{\perp}\cdot\mathbf{u}_{0}\in L^{\infty}, there is at most one weak solution 𝐮\mathbf{u} to the 2D Euler equations in vorticity formulation on [t0,)[t_{0},\infty) satisfying 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}, as originally shown in [70]. This implies that, given any Borel probability measure μ0\mu_{0} on HH that is carried by the set 𝒪0={𝐮0H:𝐮0L}\mathcal{O}_{0}=\{\mathbf{u}_{0}\in H\,:\,\nabla^{\perp}\cdot\mathbf{u}_{0}\in L^{\infty}\}, a trajectory statistical solution of the 2D Euler equations starting from this initial measure can be simply obtained as Sμ0S\mu_{0}, where SS is an associated and well-defined solution operator on 𝒪0\mathcal{O}_{0}. Moreover, together with the inviscid limit result for individual solutions, one can easily establish the convergence Sνjμ0wscSμ0S_{\nu_{j}}\mu_{0}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}S\mu_{0} for any sequence νj0\nu_{j}\to 0, where as before SνS_{\nu} denotes the solution operator associated to the 2D ν\nu-NSE. For this reason, in our results below in Section 4.2 we consider only r<r<\infty.

In the three-dimensional case, on the other hand, an analogous inviscid result for individual solutions as previously described is not currently available. Alternative, and weaker, definitions of solutions for the Euler equations were defined to circumvent the extra complications that arise in three dimensions, and consequently obtain existence of a certain type of global-in-time solution of Euler as a vanishing viscosity limit, under appropriate initial data. Two such weaker notions are the measure-valued solutions proposed in [25] and the dissipative solutions from [54]. As mentioned in Section 1, here we focus on the latter definition (cf. Definition 4.4 below), since it more directly fits our abstract framework from Section 3.

Finally, it is worth mentioning that global existence of weak solutions to the Euler equations from any given initial datum in HH and for any dimension d2d\geq 2 was recently shown in [69], but not as a vanishing viscosity limit. Specifically, [69] relies on the construction of “wild” solutions of the Euler equations developed in [24] to show the existence of an infinite number of (wild) weak solutions of (4.2) departing from any fixed initial datum in HH, and under zero forcing term.

4.1.1. 2D incompressible flows

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0}. We start by recalling the standard notions of weak solutions to the 2D Navier-Stokes and Euler equations on II. For the definitions below, we recall the space 𝒟σ\mathcal{D}_{\sigma} of test functions defined in (4.3).

Definition 4.1.

Let 𝐟Lloc2(I,V)\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,V^{\prime}). We say that 𝐮\mathbf{u} is a weak solution of the 2D Navier-Stokes equations, (4.1), on II if

  1. (i)

    𝐮𝒞loc(I,H)Lloc2(I,V)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,H)\cap L_{\textrm{\rm loc}}^{2}(I,V);

  2. (ii)

    t𝐮Lloc2(I,V)\partial_{t}\mathbf{u}\in L_{\textrm{\rm loc}}^{2}(I,V^{\prime});

  3. (iii)

    For every 𝐯𝒟σ\mathbf{v}\in\mathcal{D}_{\sigma}, the equation

    ddtΩ𝐮𝐯d𝐱+νΩ𝐮:𝐯d𝐱Ω𝐮𝐮:𝐯d𝐱=𝐟,𝐯V,V\displaystyle\frac{d}{dt}\int_{\Omega}\mathbf{u}\cdot\mathbf{v}\;{\text{\rm d}}\mathbf{x}+\nu\int_{\Omega}\nabla\mathbf{u}:\nabla\mathbf{v}\;{\text{\rm d}}\mathbf{x}-\int_{\Omega}\mathbf{u}\otimes\mathbf{u}:\nabla\mathbf{v}\;{\text{\rm d}}\mathbf{x}=\langle\mathbf{f},\mathbf{v}\rangle_{V^{\prime},V} (4.8)

    is satisfied in the sense of distributions on II.

Definition 4.2.

Let 𝐟Lloc2(I,V)\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,V^{\prime}). We say that 𝐮\mathbf{u} is a weak solution of the 2D Euler equations, (4.2), on II if

  1. (i)

    𝐮Lloc(I,H)𝒞loc(I,𝒟σ)\mathbf{u}\in L_{\textrm{\rm loc}}^{\infty}(I,H)\cap\mathcal{C}_{\textrm{\rm loc}}(I,\mathcal{D}_{\sigma}^{\prime});

  2. (ii)

    For every 𝐯𝒟σ\mathbf{v}\in\mathcal{D}_{\sigma}, the equation

    ddtΩ𝐮𝐯d𝐱Ω𝐮𝐮:𝐯d𝐱=𝐟,𝐯V,V\displaystyle\frac{d}{dt}\int_{\Omega}\mathbf{u}\cdot\mathbf{v}\;{\text{\rm d}}\mathbf{x}-\int_{\Omega}\mathbf{u}\otimes\mathbf{u}:\nabla\mathbf{v}\;{\text{\rm d}}\mathbf{x}=\langle\mathbf{f},\mathbf{v}\rangle_{V^{\prime},V} (4.9)

    is satisfied in the sense of distributions on II.

As recalled in Section 4.1, when given 𝐮0H\mathbf{u}_{0}\in H with 𝐮0Lr(Ω)\nabla^{\perp}\cdot\mathbf{u}_{0}\in L^{r}(\Omega), for some 1<r1<r\leq\infty, the existence of a weak solution 𝐮\mathbf{u} to the 2D Euler equations on II satisfying 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0} in 𝒟σ\mathcal{D}_{\sigma}^{\prime} can be shown via a vanishing viscosity limit. A proof is given in e.g. [54, Theorem 4.1] in the case of zero forcing term. However, it is not difficult to extend the proof to the case of nonzero forcing 𝐟\mathbf{f}, by assuming that 𝐟\mathbf{f} satisfies, e.g., 𝐟Lloc2(I,H)\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,H) and 𝐟Llocr(I,Lr(Ω))\nabla^{\perp}\cdot\mathbf{f}\in L^{r}_{\textrm{\rm loc}}(I,L^{r}(\Omega)). Additionally, it follows from the proof that this weak solution 𝐮\mathbf{u} also belongs to 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}).

Moreover, under these same conditions on 𝐮0\mathbf{u}_{0} and 𝐟\mathbf{f}, it is not difficult to verify that the corresponding unique weak solution 𝐮ν\mathbf{u}^{\nu} of the 2D Navier-Stokes equations on II satisfying 𝐮ν(t0)=𝐮0\mathbf{u}^{\nu}(t_{0})=\mathbf{u}_{0} in HH also belongs to 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}).

In our following results regarding the two-dimensional case, we shall maintain this assumption on 𝐟\mathbf{f}, namely 𝐟Lloc2(I,H)\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,H) and 𝐟Llocr(I,Lr(Ω))\nabla^{\perp}\cdot\mathbf{f}\in L^{r}_{\textrm{\rm loc}}(I,L^{r}(\Omega)). In this case, it thus follows that we may take the abstract space XX from Section 3 as (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, with the corresponding trajectories of Euler and Navier-Stokes lying in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}).

In the proposition below, we collect some useful inequalities valid for weak solutions of the 2D NSE, (4.1). The proof follows with similar arguments from [54, Section 4.1] under the appropriate modifications to include the forcing term 𝐟\mathbf{f}. We omit the details.

Here we point out that the upper bound in (4.11) below is uniformly bounded as ν0\nu\to 0. Clearly, this uniformity is crucial for the sake of our inviscid limit result, specifically for satisfying condition (H5) from Theorem 3.2.

Proposition 4.1.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H) with 𝐟Llocr(I,Lr(Ω))\nabla^{\perp}\cdot\mathbf{f}\in L^{r}_{\textrm{\rm loc}}(I,L^{r}(\Omega)), 2r<+2\leq r<+\infty. Then, for every weak solution 𝐮ν𝒞loc(I,(Wσ1,r)w)\mathbf{u}^{\nu}\in\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) of the 2D NSE (4.1) on II with forcing term 𝐟\mathbf{f} and for any ν0>0\nu_{0}>0, the following inequalities hold for all ν>0\nu>0 and tIt\in I:

|𝐮ν(t)|2(|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2)eν0λ1(tt0),\displaystyle|\mathbf{u}^{\nu}(t)|^{2}\leq\left(|\mathbf{u}^{\nu}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right)e^{\nu_{0}\lambda_{1}(t-t_{0})}, (4.10)
𝐮ν(t)Lrr(𝐮ν(t0)Lrr+(ν0λ1)1r𝐟Lr(t0,t;Lr)r)e(r1)ν0λ1(tt0),\displaystyle\|\nabla^{\perp}\cdot\mathbf{u}^{\nu}(t)\|_{L^{r}}^{r}\leq\left(\|\nabla^{\perp}\cdot\mathbf{u}^{\nu}(t_{0})\|_{L^{r}}^{r}+(\nu_{0}\lambda_{1})^{1-r}\|\nabla^{\perp}\cdot\mathbf{f}\|_{L^{r}(t_{0},t;L^{r})}^{r}\right)e^{(r-1)\nu_{0}\lambda_{1}(t-t_{0})}, (4.11)
t𝐮νL2(t0,t;V)cλ11/2𝐟L2(t0,t;H)\displaystyle\|\partial_{t}\mathbf{u}^{\nu}\|_{L^{2}(t_{0},t;V^{\prime})}\leq c\lambda_{1}^{-1/2}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}
+cλ11/2+1/r[(ν+(|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2)eν0λ1(tt0))\displaystyle\qquad+c\lambda_{1}^{-1/2+1/r}\Bigg{[}\left(\nu+\left(|\mathbf{u}^{\nu}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right)e^{\nu_{0}\lambda_{1}(t-t_{0})}\right)
(𝐮ν(t0)Lrr+(ν0λ1)1r𝐟Lr(t0,t;Lr)r)]e(r1)ν0λ1(tt0),\displaystyle\qquad\qquad\qquad\qquad\left(\|\nabla^{\perp}\cdot\mathbf{u}^{\nu}(t_{0})\|_{L^{r}}^{r}+(\nu_{0}\lambda_{1})^{1-r}\|\nabla^{\perp}\cdot\mathbf{f}\|_{L^{r}(t_{0},t;L^{r})}^{r}\right)\Bigg{]}e^{(r-1)\nu_{0}\lambda_{1}(t-t_{0})}, (4.12)

where c>0c>0 is a universal constant.

We present below the inviscid limit result for individual solutions that will be needed to verify some of the conditions from Theorem 3.2. The proof follows from standard arguments as in e.g. [54, Chapter 4], but we include the details here for completeness.

Proposition 4.2.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0}, and let 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{loc}(I,H) with 𝐟Llocr(I,Lr(Ω))\nabla^{\perp}\cdot\mathbf{f}\in L^{r}_{loc}(I,L^{r}(\Omega)), r2r\geq 2. Let {𝐮ν}ν>0𝒞loc(I,(Wσ1,r)w)\{\mathbf{u}^{\nu}\}_{\nu>0}\subset\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) be a vanishing viscosity net of weak solutions of the 2D Navier-Stokes equations (4.1) on II with external force 𝐟\mathbf{f}, in the sense of Definition 4.1. Then, for every convergent subnet {𝐮ν}ν\{\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} with 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) as ν0\nu^{\prime}\to 0, we have that the limit 𝐮\mathbf{u} is a weak solution of the 2D Euler equations on II with external force 𝐟\mathbf{f}, in the sense of Definition 4.2.

Proof.

Suppose {𝐮ν}ν\{\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} is a subnet converging to some 𝐮\mathbf{u} in 𝒞loc(I,(Wσ1,r)w)\mathcal{\mathcal{C}_{\textrm{\rm loc}}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) as ν0\nu^{\prime}\to 0. Let us show that 𝐮\mathbf{u} is a weak solution of the 2D Euler equations on II.

Fix any compact subinterval JIJ\subset I. Note that since, in particular, 𝐮ν(t0)𝐮(t0)\mathbf{u}^{\nu^{\prime}}(t_{0})\to\mathbf{u}(t_{0}) in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}} as ν0\nu^{\prime}\to 0, then {𝐮ν(t0)}ν\{\mathbf{u}^{\nu^{\prime}}(t_{0})\}_{\nu^{\prime}} is uniformly bounded in Wσ1,rW_{\sigma}^{1,r}. Then, from the a priori bounds (4.11) and (4.1), it follows that {𝐮ν}ν\{\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} is uniformly bounded in L(J,Wσ1,r)L^{\infty}(J,W_{\sigma}^{1,r}) and {t𝐮ν}ν\{\partial_{t}\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} is uniformly bounded in L2(J,V)L^{2}(J,V^{\prime}). Hence, since Wσ1,rW_{\sigma}^{1,r} is compactly embedded in HH for r>1r>1, we can apply Aubin-Lions Lemma ([38, Theorem A.11]) to obtain that, up to a subnet, 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in 𝒞(J,H)\mathcal{C}(J,H) as ν0\nu^{\prime}\to 0. In particular, 𝐮L(J,H)\mathbf{u}\in L^{\infty}(J,H) and, consequently, 𝐮Lloc(I,H)\mathbf{u}\in L^{\infty}_{\textrm{\rm loc}}(I,H). Moreover, since 𝐮𝒞loc(I,(Wσ1,r)w)𝒞loc(I,𝒟σ)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}})\subset\mathcal{C}_{\textrm{\rm loc}}(I,\mathcal{D}_{\sigma}^{\prime}), we deduce that condition (i) of Definition 4.2 is satisfied.

To verify the remaining condition, (ii), fix any test function φ𝒞c(I)\varphi\in\mathcal{C}^{\infty}_{\textrm{\rm c}}(I) and 𝐯𝒟σ\mathbf{v}\in\mathcal{D}_{\sigma}. By assumption, we have that

IΩ(𝐮ν𝐯)φd𝐱dt+νIΩ(𝐮ν:𝐯)φd𝐱dtIΩ(𝐮ν𝐮ν:𝐯)φd𝐱dt=Ω(𝐮ν(t0)𝐯)φ(t0)d𝐱+I𝐟,𝐯V,Vφdt-\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}\cdot\mathbf{v})\varphi^{\prime}\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t+\nu^{\prime}\int_{I}\int_{\Omega}(\nabla\mathbf{u}^{\nu^{\prime}}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t-\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}\otimes\mathbf{u}^{\nu^{\prime}}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\\ =\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}(t_{0})\cdot\mathbf{v})\varphi(t_{0})\;{\text{\rm d}}\mathbf{x}+\int_{I}\langle\mathbf{f},\mathbf{v}\rangle_{V^{\prime},V}\varphi\;{\text{\rm d}}t (4.13)

for every ν\nu^{\prime}.

Since φ\varphi has compact support in II, in view of the convergence 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in 𝒞loc(I,(Wσ1,r)w)\mathcal{\mathcal{C}_{\textrm{\rm loc}}}(I,(W_{\sigma}^{1,r})_{\rm{w}}), we immediately obtain

IΩ(𝐮ν𝐯)φd𝐱dtIΩ(𝐮𝐯)φd𝐱dt, as ν0,\displaystyle\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}\cdot\mathbf{v})\varphi^{\prime}\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\longrightarrow\int_{I}\int_{\Omega}(\mathbf{u}\cdot\mathbf{v})\varphi^{\prime}\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t,\quad\mbox{ as }\nu^{\prime}\to 0,
νIΩ(𝐮ν:𝐯)φd𝐱dt=νIΩ(𝐮ν:Δ𝐯)φd𝐱dt0, as ν0,\displaystyle\nu^{\prime}\int_{I}\int_{\Omega}(\nabla\mathbf{u}^{\nu^{\prime}}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t=-\nu^{\prime}\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}:\Delta\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\longrightarrow 0,\quad\mbox{ as }\nu^{\prime}\to 0,
Ω(𝐮ν(t0)𝐯)φ(t0)dtΩ(𝐮(t0)𝐯)φ(t0)dt, as ν0.\displaystyle\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}(t_{0})\cdot\mathbf{v})\varphi(t_{0})\;{\text{\rm d}}t\longrightarrow\int_{\Omega}(\mathbf{u}(t_{0})\cdot\mathbf{v})\varphi(t_{0})\;{\text{\rm d}}t,\quad\mbox{ as }\nu^{\prime}\to 0.

Regarding the nonlinear term in (4.13), we proceed as follows. Let JIJ\subset I be a compact subinterval containing the support of φ\varphi. Note that

|IΩ(𝐮ν𝐮ν:𝐯)φd𝐱dtIΩ(𝐮𝐮:𝐯)φd𝐱dt|\displaystyle\left|\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}\otimes\mathbf{u}^{\nu^{\prime}}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t-\int_{I}\int_{\Omega}(\mathbf{u}\otimes\mathbf{u}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\right|
=|JΩ[((𝐮ν𝐮)𝐮ν+𝐮(𝐮ν𝐮)):𝐯]φd𝐱dt|\displaystyle\qquad\qquad=\left|\int_{J}\int_{\Omega}\left[\left((\mathbf{u}^{\nu^{\prime}}-\mathbf{u})\otimes\mathbf{u}^{\nu^{\prime}}+\mathbf{u}\otimes(\mathbf{u}^{\nu^{\prime}}-\mathbf{u})\right):\nabla\mathbf{v}\right]\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\right|
𝐮ν𝐮L(J;L2(Ω))(𝐮νL(J;L2(Ω))+𝐮L(J;L2(Ω)))𝐯L(Ω)φL1(J).\displaystyle\qquad\qquad\leq\|\mathbf{u}^{\nu^{\prime}}-\mathbf{u}\|_{L^{\infty}(J;L^{2}(\Omega))}(\|\mathbf{u}^{\nu^{\prime}}\|_{L^{\infty}(J;L^{2}(\Omega))}+\|\mathbf{u}\|_{L^{\infty}(J;L^{2}(\Omega))})\|\nabla\mathbf{v}\|_{L^{\infty}(\Omega)}\|\varphi\|_{L^{1}(J)}.

Since 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in L(J,H)L^{\infty}(J,H) as ν0\nu^{\prime}\to 0, it follows that

IΩ(𝐮ν𝐮ν:𝐯)φd𝐱dtIΩ(𝐮𝐮:𝐯)φd𝐱dt, as ν0.\displaystyle\int_{I}\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}\otimes\mathbf{u}^{\nu^{\prime}}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t\longrightarrow\int_{I}\int_{\Omega}(\mathbf{u}\otimes\mathbf{u}:\nabla\mathbf{v})\varphi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t,\quad\mbox{ as }\nu^{\prime}\to 0.

Therefore, passing to the limit as ν0\nu^{\prime}\to 0 in (4.13), we deduce that 𝐮\mathbf{u} satisfies item (ii) of Definition 4.2. This concludes the proof. ∎

4.1.2. 3D incompressible flows

Let us again take II\subset\mathbb{R} to be an interval closed and bounded on the left with left endpoint t0t_{0}. We recall the following standard notion of weak solution to the 3D Navier-Stokes equations.

Definition 4.3.

Let 𝐟Lloc2(I,V)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,V^{\prime}). We say that 𝐮\mathbf{u} is a Leray-Hopf weak solution of the 3D Navier-Stokes equations (4.1) on II if

  1. (i)

    𝐮𝒞loc(I,Hw)Lloc2(I,V)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}})\cap L_{\textrm{\rm loc}}^{2}(I,V);

  2. (ii)

    t𝐮Lloc4/3(I,V)\partial_{t}\mathbf{u}\in L_{\textrm{\rm loc}}^{4/3}(I,V^{\prime});

  3. (iii)

    For every function Φ𝒞(I×3)3\Phi\in\mathcal{C}^{\infty}(I\times\mathbb{R}^{3})^{3} that is Ω\Omega-periodic, divergence-free and compactly supported on II, it holds

    IΩ𝐮tΦd𝐱dtνIΩ𝐮:Φd𝐱dt+IΩ𝐮𝐮:Φd𝐱dt\displaystyle\int_{I}\int_{\Omega}\mathbf{u}\cdot\partial_{t}\Phi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t-\nu\int_{I}\int_{\Omega}\nabla\mathbf{u}:\nabla\Phi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t+\int_{I}\int_{\Omega}\mathbf{u}\otimes\mathbf{u}:\nabla\Phi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t
    =Ω𝐮(t0)Φ(t0)d𝐱IΩ𝐟Φd𝐱dt.\displaystyle=-\int_{\Omega}\mathbf{u}(t_{0})\cdot\Phi(t_{0})\;{\text{\rm d}}\mathbf{x}-\int_{I}\int_{\Omega}\mathbf{f}\cdot\Phi\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}t. (4.14)
  4. (iv)

    𝐮\mathbf{u} satisfies the following energy inequality for almost all tIt^{\prime}\in I and for all tIt\in I with t>tt>t^{\prime}:

    12|𝐮(t)|2+νtt𝐮(s)2ds12|𝐮(t)|2+tt𝐟(s),𝐮(s)V,Vds.\frac{1}{2}|\mathbf{u}(t)|^{2}+\nu\int_{t^{\prime}}^{t}\|\mathbf{u}(s)\|^{2}\;{\text{\rm d}}s\leq\frac{1}{2}|\mathbf{u}(t^{\prime})|^{2}+\int_{t^{\prime}}^{t}\langle\mathbf{f}(s),\mathbf{u}(s)\rangle_{V^{\prime},V}\;{\text{\rm d}}s. (4.15)
  5. (v)

    If II is closed and bounded on the left, with left endpoint t0t_{0}, then 𝐮\mathbf{u} is strongly continuous in HH at t0t_{0} from the right, i.e., 𝐮(t)𝐮(t0)\mathbf{u}(t)\rightarrow\mathbf{u}(t_{0}) in HH as tt0+t\rightarrow t_{0}^{+}.

The set of allowed times tt^{\prime} in (4.15) are characterized as the points of strong continuity from the right of 𝐮\mathbf{u} in HH. In particular, condition (v) implies that t=t0t^{\prime}=t_{0} is allowed in that case.

We note that condition (iv) can be interchanged with the following inequality in the sense of distributions on II:

12ddt|𝐮(t)|2+ν𝐮(t)2𝐟(t),𝐮(t)V,V,\displaystyle\frac{1}{2}\frac{{\text{\rm d}}}{{\text{\rm d}}t}|\mathbf{u}(t)|^{2}+\nu\|\mathbf{u}(t)\|^{2}\leq\langle\mathbf{f}(t),\mathbf{u}(t)\rangle_{V^{\prime},V}, (4.16)

see e.g. [40].

Given any 𝐟Lloc2(I,V)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,V^{\prime}) and initial datum 𝐮0H\mathbf{u}_{0}\in H, it is well known that there exists at least one Leray-Hopf weak solution of the 3D Navier-Stokes equations, (4.1), defined on II and satisfying 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}. For a proof of this classical result, we refer to e.g. [21, 51, 53, 62, 61].

Regarding the 3D Euler equations, we consider the notion of dissipative solution introduced in [54, Section 4.4], where the forcing term was taken to be zero for simplicity. With the appropriate modifications to include an external force, we obtain the following definition.

Definition 4.4.

Let 𝐟Lloc1(I,H)\mathbf{f}\in L^{1}_{\textrm{\rm loc}}(I,H). We say that 𝐮\mathbf{u} is a dissipative solution of the 3D Euler equations (4.2) on II if

  1. (i)

    𝐮𝒞loc(I,Hw)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}});

  2. (ii)

    For every 𝐯𝒞loc(I,H)\mathbf{v}\in\mathcal{C}_{\textrm{\rm loc}}(I,H) such that d(𝐯)=12(𝐯+(𝐯)T)Lloc1(I,L)d(\mathbf{v})=\frac{1}{2}(\nabla\mathbf{v}+(\nabla\mathbf{v})^{T})\in L^{1}_{\textrm{\rm loc}}(I,L^{\infty}) and E(𝐯)=t𝐯[(𝐯)𝐯)]L1loc(I,H)E(\mathbf{v})=-\partial_{t}\mathbf{v}-\mathbb{P}[(\mathbf{v}\cdot\nabla)\mathbf{v})]\in L^{1}_{\textrm{\rm loc}}(I,H), where \mathbb{P} denotes the projection onto Ω\Omega-periodic divergence-free vector fields with zero spatial average, it holds

    Ω|𝐮(t)𝐯(t)|2d𝐱exp(2t0td(𝐯)Lds)Ω|𝐮(t0)𝐯(t0)|2d𝐱+2t0tΩexp(2std(𝐯)Ldτ)(E(𝐯)+𝐟)(𝐮𝐯)d𝐱ds,\int_{\Omega}|\mathbf{u}(t)-\mathbf{v}(t)|^{2}\;{\text{\rm d}}\mathbf{x}\leq\exp\left(2\int_{t_{0}}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\;{\text{\rm d}}s\right)\int_{\Omega}|\mathbf{u}(t_{0})-\mathbf{v}(t_{0})|^{2}\;{\text{\rm d}}\mathbf{x}\\ +2\int_{t_{0}}^{t}\int_{\Omega}\exp\left(2\int_{s}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\;{\text{\rm d}}\tau\right)(E(\mathbf{v})+\mathbf{f})\cdot(\mathbf{u}-\mathbf{v})\;{\text{\rm d}}\mathbf{x}\;{\text{\rm d}}s, (4.17)

    for all tIt\in I, where d(𝐯)=(inf{ξTd(𝐯)ξ:ξ2,|ξ|=1})d^{-}(\mathbf{v})=(\inf\{\xi^{T}d(\mathbf{v})\xi:\xi\in\mathbb{R}^{2},|\xi|=1\})^{-} is the negative part of the smallest eigenvalue of d(𝐯)d(\mathbf{v}).

As mentioned in Section 4.1, the main motivation behind this definition comes from establishing a notion of solution to the 3D Euler equations that is obtained as an appropriate limit of a vanishing viscosity sequence of Leray-Hopf weak solutions of the 3D NSE. This is indeed how the existence of a dissipative solution is shown in [54, Proposition 4.2], under an initial condition 𝐮(t0)=𝐮0\mathbf{u}(t_{0})=\mathbf{u}_{0}, for any 𝐮0H\mathbf{u}_{0}\in H, and in the absence of external forcing term. With a simple adaptation, one can show the same holds for any given forcing 𝐟Lloc1(I,H)\mathbf{f}\in L^{1}_{\textrm{\rm loc}}(I,H).

Following analogous steps from this proof, we obtain the inviscid limit result in 4.4 below, which we later apply for verifying condition (A2) of Theorem 3.1. We present the details of its proof here for completeness. Before proceeding, we show in the following proposition a few useful a priori estimates regarding weak solutions of the 3D NSE. For this formulation and the subsequent results, we require the forcing term to be in Lloc2(I,H),L^{2}_{\textrm{\rm loc}}(I,H), so it fits both Definitions 4.3 and 4.4.

Proposition 4.3.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Let ν0>0\nu_{0}>0. Then, for every Leray-Hopf weak solution 𝐮ν\mathbf{u}^{\nu} of the 3D NSE (4.1) on II with forcing term 𝐟\mathbf{f} and for all ν>0\nu>0, the following inequalities hold:

|𝐮ν(t)|2+2νt0teν0λ1(tτ)𝐮ν(τ)2dτeν0λ1(tt0)[|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2],\displaystyle|\mathbf{u}^{\nu}(t)|^{2}+2\nu\int_{t_{0}}^{t}e^{\nu_{0}\lambda_{1}(t-\tau)}\|\mathbf{u}^{\nu}(\tau)\|^{2}{\text{\rm d}}\tau\leq e^{\nu_{0}\lambda_{1}(t-t_{0})}\left[|\mathbf{u}^{\nu}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right], (4.18)

for all tIt\in I, and

𝐮ν(t)𝐮ν(s)(Wσ1,)c|ts|1/2(ν1/2+ν01/2)λ13/4eν0λ1(tt0)2[|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2]1/2+|ts|eν0λ1(tt0)[|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2],\|\mathbf{u}^{\nu}(t)-\mathbf{u}^{\nu}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\\ \leq c|t-s|^{1/2}(\nu^{1/2}+\nu_{0}^{1/2})\lambda_{1}^{-3/4}e^{\frac{\nu_{0}\lambda_{1}(t-t_{0})}{2}}\left[|\mathbf{u}^{\nu}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right]^{1/2}\\ +|t-s|e^{\nu_{0}\lambda_{1}(t-t_{0})}\left[|\mathbf{u}^{\nu}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right], (4.19)

for all s,tIs,t\in I with sts\leq t, and for some positive constant cc which is independent of ν\nu.

Proof.

From (4.16), it follows that, for every non-negative test function φ𝒞c(I)\varphi\in\mathcal{C}^{\infty}_{\textrm{\rm c}}(I),

12t0t|𝐮ν(τ)|2φ(τ)dτ+νt0t𝐮ν(τ)2φ(τ)dτt0t(𝐟(τ),𝐮ν(τ))φ(τ)dτ,\displaystyle-\frac{1}{2}\int_{t_{0}}^{t}|\mathbf{u}^{\nu}(\tau)|^{2}\varphi^{\prime}(\tau){\text{\rm d}}\tau+\nu\int_{t_{0}}^{t}\|\mathbf{u}^{\nu}(\tau)\|^{2}\varphi(\tau){\text{\rm d}}\tau\leq\int_{t_{0}}^{t}(\mathbf{f}(\tau),\mathbf{u}^{\nu}(\tau))\varphi(\tau){\text{\rm d}}\tau,

for all tIt\in I. Then, by choosing an appropriate sequence of test functions on II and invoking the Lebesgue differentiation theorem, together with the fact that 𝐮ν𝒞loc(I,Hw)\mathbf{u}^{\nu}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) and 𝐮ν\mathbf{u}^{\nu} is strongly continuous at t0t_{0} from the right, we deduce that

|𝐮ν(t)|2ψ(t)+2νt0t𝐮ν(τ)2ψ(τ)dτ\displaystyle|\mathbf{u}^{\nu}(t)|^{2}\psi(t)+2\nu\int_{t_{0}}^{t}\|\mathbf{u}^{\nu}(\tau)\|^{2}\psi(\tau){\text{\rm d}}\tau
|𝐮ν(t0)|2ψ(t0)+t0t|𝐮ν(τ)|2ψ(τ)dτ+2t0t(𝐟(τ),𝐮ν(τ))ψ(τ)dτ,\displaystyle\qquad\leq|\mathbf{u}^{\nu}(t_{0})|^{2}\psi(t_{0})+\int_{t_{0}}^{t}|\mathbf{u}^{\nu}(\tau)|^{2}\psi^{\prime}(\tau){\text{\rm d}}\tau+2\int_{t_{0}}^{t}(\mathbf{f}(\tau),\mathbf{u}^{\nu}(\tau))\psi(\tau){\text{\rm d}}\tau, (4.20)

for all tIt\in I and for every non-negative function ψ𝒞1(I)\psi\in\mathcal{C}^{1}(I); see e.g. [38, Chapter II, Appendix B.1] for a similar argument.

In particular, choosing ψ(t)=eν0λ1t\psi(t)=e^{-\nu_{0}\lambda_{1}t}, and estimating the integrand in the last term of (4.20) as

(𝐟(τ),𝐮ν(τ))|𝐟(τ)||𝐮ν(τ)|12ν0λ1|𝐟(τ)|2+ν0λ12|𝐮ν(τ)|2,\displaystyle(\mathbf{f}(\tau),\mathbf{u}^{\nu}(\tau))\leq|\mathbf{f}(\tau)||\mathbf{u}^{\nu}(\tau)|\leq\frac{1}{2\nu_{0}\lambda_{1}}|\mathbf{f}(\tau)|^{2}+\frac{\nu_{0}\lambda_{1}}{2}|\mathbf{u}^{\nu}(\tau)|^{2},

it follows that

|𝐮ν(t)|2eν0λ1t+2νt0t𝐮ν(τ)2eν0λ1τdτ\displaystyle|\mathbf{u}^{\nu}(t)|^{2}e^{-\nu_{0}\lambda_{1}t}+2\nu\int_{t_{0}}^{t}\|\mathbf{u}^{\nu}(\tau)\|^{2}e^{-\nu_{0}\lambda_{1}\tau}{\text{\rm d}}\tau |𝐮ν(t0)|2eν0λ1t0+1ν0λ1t0t|𝐟(τ)|2eν0λ1τdτ\displaystyle\leq|\mathbf{u}^{\nu}(t_{0})|^{2}e^{-\nu_{0}\lambda_{1}t_{0}}+\frac{1}{\nu_{0}\lambda_{1}}\int_{t_{0}}^{t}|\mathbf{f}(\tau)|^{2}e^{-\nu_{0}\lambda_{1}\tau}{\text{\rm d}}\tau
|𝐮ν(t0)|2eν0λ1t0+eν0λ1t0ν0λ1𝐟L2(t0,t;H)2,\displaystyle\leq|\mathbf{u}^{\nu}(t_{0})|^{2}e^{-\nu_{0}\lambda_{1}t_{0}}+\frac{e^{-\nu_{0}\lambda_{1}t_{0}}}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2},

which immediately yields (4.18).

Regarding (4.19), let s,tIs,t\in I with sts\leq t. Since t𝐮νLloc4/3(I,V)Lloc1(I,(Wσ1,)),\partial_{t}\mathbf{u}^{\nu}\in L^{4/3}_{\textrm{\rm loc}}(I,V^{\prime})\subset L^{1}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,\infty})^{\prime}), then, for all 𝐯Wσ1,,\mathbf{v}\in W_{\sigma}^{1,\infty},

|𝐮ν(t)𝐮ν(s),𝐯(Wσ1,),Wσ1,|=|stt𝐮ν(τ)dτ,𝐯(Wσ1,),Wσ1,|(stt𝐮ν(τ)(Wσ1,)dτ)𝐯Wσ1,|ts|1/2t𝐮νL2(s,t;(Wσ1,))𝐯Wσ1,.|\langle\mathbf{u}^{\nu}(t)-\mathbf{u}^{\nu}(s),\mathbf{v}\rangle_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}|=\left|{\left\langle\int_{s}^{t}\partial_{t}\mathbf{u}^{\nu}(\tau){\text{\rm d}}\tau,\mathbf{v}\right\rangle}_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}\right|\\ \leq\left(\int_{s}^{t}\|\partial_{t}\mathbf{u}^{\nu}(\tau)\|_{(W_{\sigma}^{1,\infty})^{\prime}}{\text{\rm d}}\tau\right)\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}}\leq|t-s|^{1/2}\|\partial_{t}\mathbf{u}^{\nu}\|_{L^{2}(s,t;(W_{\sigma}^{1,\infty})^{\prime})}\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}}.

Hence,

𝐮ν(t)𝐮ν(s)(Wσ1,)|ts|1/2t𝐮νL2(s,t;(Wσ1,)).\displaystyle\|\mathbf{u}^{\nu}(t)-\mathbf{u}^{\nu}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\leq|t-s|^{1/2}\|\partial_{t}\mathbf{u}^{\nu}\|_{L^{2}(s,t;(W_{\sigma}^{1,\infty})^{\prime})}.

We proceed to obtain an estimate of t𝐮νL2(s,t;(Wσ1,))\|\partial_{t}\mathbf{u}^{\nu}\|_{L^{2}(s,t;(W_{\sigma}^{1,\infty})^{\prime})}. From ((iii)), it follows that, for all 𝐯V,\mathbf{v}\in V,

ddt(𝐮ν,𝐯)+ν(𝐮ν,𝐯)(𝐮ν𝐮ν,𝐯)=(𝐟,𝐯)\displaystyle\frac{{\text{\rm d}}}{{\text{\rm d}}t}(\mathbf{u}^{\nu},\mathbf{v})+\nu(\nabla\mathbf{u}^{\nu},\nabla\mathbf{v})-(\mathbf{u}^{\nu}\otimes\mathbf{u}^{\nu},\nabla\mathbf{v})=(\mathbf{f},\mathbf{v}) (4.21)

in the sense of distributions on II. In particular, again since t𝐮νLloc1(I,(Wσ1,))\partial_{t}\mathbf{u}^{\nu}\in L^{1}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,\infty})^{\prime}) then for all 𝐯Wσ1,\mathbf{v}\in W_{\sigma}^{1,\infty}

t𝐮ν,𝐯(Wσ1,),Wσ1,=ddt𝐮ν,𝐯(Wσ1,),Wσ1,=ddt(𝐮ν,𝐯).\displaystyle{\left\langle\partial_{t}\mathbf{u}^{\nu},\mathbf{v}\right\rangle}_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}=\frac{{\text{\rm d}}}{{\text{\rm d}}t}\langle\mathbf{u}^{\nu},\mathbf{v}\rangle_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}=\frac{{\text{\rm d}}}{{\text{\rm d}}t}(\mathbf{u}^{\nu},\mathbf{v}).

From (4.21), along with Cauchy-Schwarz, Hölder’s inequality, Poincaré inequality (4.4), and the inequality (4.6) for r=,r=\infty, we thus obtain

t𝐮ν(τ),𝐯(Wσ1,),Wσ1,\displaystyle{\left\langle\partial_{t}\mathbf{u}^{\nu}(\tau),\mathbf{v}\right\rangle}_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}} ν𝐮ν(τ)𝐯+|𝐮ν(τ)|2𝐯Wσ1,+|𝐟(τ)||𝐯|\displaystyle\leq\nu\|\mathbf{u}^{\nu}(\tau)\|\|\mathbf{v}\|+|\mathbf{u}^{\nu}(\tau)|^{2}\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}}+|\mathbf{f}(\tau)||\mathbf{v}|
cνλ13/4𝐮ν(τ)𝐯Wσ1,+|𝐮ν(τ)|2𝐯Wσ1,+cλ15/4|𝐟(τ)|𝐯Wσ1,,\displaystyle\leq c\nu\lambda_{1}^{-3/4}\|\mathbf{u}^{\nu}(\tau)\|\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}}+|\mathbf{u}^{\nu}(\tau)|^{2}\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}}+c\lambda_{1}^{-5/4}|\mathbf{f}(\tau)|\|\mathbf{v}\|_{W_{\sigma}^{1,\infty}},

for a.e. τI\tau\in I. Consequently,

t𝐮ν(τ)(Wσ1,)cλ13/4(ν𝐮ν(τ)+λ11/2|𝐟(τ)|)+|𝐮ν(τ)|2,for a.e. τI.\displaystyle\|\partial_{t}\mathbf{u}^{\nu}(\tau)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\leq c\lambda_{1}^{-3/4}(\nu\|\mathbf{u}^{\nu}(\tau)\|+\lambda_{1}^{-1/2}|\mathbf{f}(\tau)|)+|\mathbf{u}^{\nu}(\tau)|^{2},\quad\mbox{for a.e. }\tau\in I.

Hence,

t𝐮νL2(s,t;(Wσ1,))cλ13/4(ν𝐮νL2(s,t;V)+λ11/2𝐟L2(s,t;H))+𝐮νL4(s,t;H)2cλ13/4(ν𝐮νL2(t0,t;V)+λ11/2𝐟L2(t0,t;H))+|ts|1/2𝐮νL(t0,t;H)2.\|\partial_{t}\mathbf{u}^{\nu}\|_{L^{2}(s,t;(W_{\sigma}^{1,\infty})^{\prime})}\leq c\lambda_{1}^{-3/4}\left(\nu\|\mathbf{u}^{\nu}\|_{L^{2}(s,t;V)}+\lambda_{1}^{-1/2}\|\mathbf{f}\|_{L^{2}(s,t;H)}\right)+\|\mathbf{u}^{\nu}\|^{2}_{L^{4}(s,t;H)}\\ \leq c\lambda_{1}^{-3/4}\left(\nu\|\mathbf{u}^{\nu}\|_{L^{2}(t_{0},t;V)}+\lambda_{1}^{-1/2}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}\right)+|t-s|^{1/2}\|\mathbf{u}^{\nu}\|_{L^{\infty}(t_{0},t;H)}^{2}. (4.22)

Thus, (4.19) follows by invoking (4.18) to further estimate the right-hand side of (4.22). ∎

Proposition 4.4.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0}, and let 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Consider also a (strongly) compact set KK in HH, and let {𝐮ν}ν>0\{\mathbf{u}^{\nu}\}_{\nu>0} be a vanishing viscosity net of Leray-Hopf weak solutions to the 3D Navier-Stokes equations (4.1) on II with external force 𝐟\mathbf{f} and with initial data 𝐮ν(t0)K\mathbf{u}^{\nu}(t_{0})\in K, in the sense of Definition 4.3. Then, for every convergent subnet {𝐮ν}ν\{\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} with 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) as ν0\nu^{\prime}\to 0, we have that the limit 𝐮\mathbf{u} is a dissipative solution of the 3D Euler equations on II with external force 𝐟\mathbf{f}, in the sense of Definition 4.4.

Proof.

Let {𝐮ν}ν\{\mathbf{u}^{\nu^{\prime}}\}_{\nu^{\prime}} be a subnet converging to some 𝐮\mathbf{u} in 𝒞loc(I,Hw)\mathcal{\mathcal{C}_{\textrm{\rm loc}}}(I,H_{\textrm{\rm w}}) as ν0\nu^{\prime}\to 0. Thus, 𝐮\mathbf{u} satisfies condition (i) of Definition 4.4.

Now let us prove that 𝐮\mathbf{u} satisfies condition (ii). By a simple density argument, it suffices to show that (4.17) holds for any test function 𝐯𝒞(I×3)3\mathbf{v}\in\mathcal{C}^{\infty}(I\times\mathbb{R}^{3})^{3} that is Ω\Omega-periodic, divergence-free and compactly supported on II (see [54, Section 4.4]). Let 𝐯\mathbf{v} be such a test function. From ((iii)), it follows that, for every ν,\nu^{\prime},

ddt(𝐮ν,𝐯)(𝐮ν,t𝐯)+ν(𝐮ν,𝐯)(𝐮ν𝐮ν,𝐯)=(𝐟,𝐯)\displaystyle\frac{{\text{\rm d}}}{{\text{\rm d}}t}(\mathbf{u}^{\nu^{\prime}},\mathbf{v})-(\mathbf{u}^{\nu^{\prime}},\partial_{t}\mathbf{v})+\nu^{\prime}(\nabla\mathbf{u}^{\nu^{\prime}},\nabla\mathbf{v})-(\mathbf{u}^{\nu^{\prime}}\otimes\mathbf{u}^{\nu^{\prime}},\nabla\mathbf{v})=(\mathbf{f},\mathbf{v})

in the sense of distributions on II.

Since t𝐯=E(𝐯)+[(𝐯)𝐯]-\partial_{t}\mathbf{v}=E(\mathbf{v})+\mathbb{P}[(\mathbf{v}\cdot\nabla)\mathbf{v}] and

(𝐮ν,[(𝐯)𝐯])(𝐮ν𝐮ν,𝐯)=(𝐮ν,(𝐯)𝐯)((𝐮ν)𝐯,𝐮ν)=([(𝐮ν𝐯)]𝐯,𝐮ν)=([(𝐮ν𝐯)]𝐯,𝐮ν𝐯),(\mathbf{u}^{\nu^{\prime}},\mathbb{P}[(\mathbf{v}\cdot\nabla)\mathbf{v}])-(\mathbf{u}^{\nu^{\prime}}\otimes\mathbf{u}^{\nu^{\prime}},\nabla\mathbf{v})=(\mathbf{u}^{\nu^{\prime}},(\mathbf{v}\cdot\nabla)\mathbf{v})-((\mathbf{u}^{\nu^{\prime}}\cdot\nabla)\mathbf{v},\mathbf{u}^{\nu^{\prime}})\\ =-([(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla]\mathbf{v},\mathbf{u}^{\nu^{\prime}})=-([(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla]\mathbf{v},\mathbf{u}^{\nu^{\prime}}-\mathbf{v}),

then

ddt(𝐮ν,𝐯)+(𝐮ν,E(𝐯))([(𝐮ν𝐯)]𝐯,𝐮ν𝐯)+ν(𝐮ν,𝐯)=(𝐟,𝐯).\displaystyle\frac{d}{dt}(\mathbf{u}^{\nu^{\prime}},\mathbf{v})+(\mathbf{u}^{\nu^{\prime}},E(\mathbf{v}))-([(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla]\mathbf{v},\mathbf{u}^{\nu^{\prime}}-\mathbf{v})+\nu^{\prime}(\nabla\mathbf{u}^{\nu^{\prime}},\nabla\mathbf{v})=(\mathbf{f},\mathbf{v}).

Note also that

ddt|𝐯|2=2(𝐯,t𝐯)=2(𝐯,E(𝐯)).\displaystyle\frac{d}{dt}|\mathbf{v}|^{2}=2(\mathbf{v},\partial_{t}\mathbf{v})=-2(\mathbf{v},E(\mathbf{v})).

Combining the last two equations with the energy inequality ddt|𝐮ν|22(𝐟,𝐮ν)\frac{d}{dt}|\mathbf{u}^{\nu^{\prime}}|^{2}\leq 2(\mathbf{f},\mathbf{u}^{\nu^{\prime}}), which follows from (4.16), we obtain

ddt|𝐮ν𝐯|22(𝐮ν𝐯,E(𝐯)+𝐟)2([(𝐮ν𝐯)]𝐯,𝐮ν𝐯)+2ν(𝐮ν,𝐯).\frac{d}{dt}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}\leq 2(\mathbf{u}^{\nu^{\prime}}-\mathbf{v},E(\mathbf{v})+\mathbf{f})-2([(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla]\mathbf{v},\mathbf{u}^{\nu^{\prime}}-\mathbf{v})+2\nu^{\prime}(\nabla\mathbf{u}^{\nu^{\prime}},\nabla\mathbf{v}).

Observe that

2([(𝐮ν𝐯)]𝐯,𝐮ν𝐯)=2Ω[((𝐮ν𝐯))𝐯](𝐮ν𝐯)d𝐱\displaystyle-2([(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla]\mathbf{v},\mathbf{u}^{\nu^{\prime}}-\mathbf{v})=-2\int_{\Omega}[((\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\cdot\nabla)\mathbf{v}]\cdot(\mathbf{u}^{\nu^{\prime}}-\mathbf{v}){\text{\rm d}}\mathbf{x}
=2Ω(𝐮ν𝐯)ii𝐯j(𝐮ν𝐯)jd𝐱\displaystyle\qquad\qquad=-2\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})_{i}\partial_{i}\mathbf{v}_{j}(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})_{j}\,{\text{\rm d}}\mathbf{x}
=2Ω(𝐮ν𝐯)ii𝐯j+j𝐯i2(𝐮ν𝐯)jd𝐱=2Ω(d(𝐯)(𝐮ν𝐯))(𝐮ν𝐯)d𝐱\displaystyle\qquad\qquad=-2\int_{\Omega}(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})_{i}\frac{\partial_{i}\mathbf{v}_{j}+\partial_{j}\mathbf{v}_{i}}{2}(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})_{j}\,{\text{\rm d}}\mathbf{x}=-2\int_{\Omega}(d(\mathbf{v})(\mathbf{u}^{\nu^{\prime}}-\mathbf{v}))\cdot(\mathbf{u}^{\nu^{\prime}}-\mathbf{v})\,{\text{\rm d}}\mathbf{x}
2Ωsup{(d(𝐯)ξ)ξ:|ξ|=1}|𝐮ν𝐯|2d𝐱\displaystyle\qquad\qquad\leq 2\int_{\Omega}\sup\{-(d(\mathbf{v})\xi)\cdot\xi:|\xi|=1\}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}\,{\text{\rm d}}\mathbf{x}
=2Ωinf{ξTd(𝐯)ξ:|ξ|=1}|𝐮ν𝐯|2d𝐱\displaystyle\qquad\qquad=-2\int_{\Omega}\inf\{\xi^{T}d(\mathbf{v})\xi:|\xi|=1\}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}\,{\text{\rm d}}\mathbf{x}
2Ωd(𝐯)|𝐮ν𝐯|2d𝐱2d(𝐯)L|𝐮ν𝐯|2.\displaystyle\qquad\qquad\leq 2\int_{\Omega}d^{-}(\mathbf{v})|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}\,{\text{\rm d}}\mathbf{x}\leq 2\|d^{-}(\mathbf{v})\|_{L^{\infty}}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}.

Thus,

ddt|𝐮ν𝐯|22(𝐮ν𝐯,E(𝐯)+𝐟)+2d(𝐯)L|𝐮ν𝐯|2+2ν𝐯𝐮ν\displaystyle\frac{{\text{\rm d}}}{{\text{\rm d}}t}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}\leq 2(\mathbf{u}^{\nu^{\prime}}-\mathbf{v},E(\mathbf{v})+\mathbf{f})+2\|d^{-}(\mathbf{v})\|_{L^{\infty}}|\mathbf{u}^{\nu^{\prime}}-\mathbf{v}|^{2}+2\nu^{\prime}\|\mathbf{v}\|\|\mathbf{u}^{\nu^{\prime}}\|

in the sense of distributions on II. Choosing an appropriate sequence of test functions on II and invoking the Lebesgue differentiation theorem, similarly as in [38, Chapter II, Appendix B.1], we obtain the following Gronwall-type inequality

|𝐮ν(t)𝐯(t)|2\displaystyle|\mathbf{u}^{\nu^{\prime}}(t)-\mathbf{v}(t)|^{2} exp(2t0td(𝐯)Lds)|𝐮ν(t0)𝐯(t0)|2\displaystyle\leq\exp\left(2\int_{t_{0}}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}s\right)|\mathbf{u}^{\nu^{\prime}}(t_{0})-\mathbf{v}(t_{0})|^{2}
+2t0texp(2std(𝐯)Ldτ)(𝐮ν𝐯,E(𝐯)+𝐟)ds\displaystyle\qquad+2\int_{t_{0}}^{t}\exp\left(2\int_{s}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)(\mathbf{u}^{\nu^{\prime}}-\mathbf{v},E(\mathbf{v})+\mathbf{f})\,{\text{\rm d}}s
+2νt0texp(2std(𝐯)Ldτ)𝐯𝐮νds,\displaystyle\qquad+2\nu^{\prime}\int_{t_{0}}^{t}\exp\left(2\int_{s}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)\|\mathbf{v}\|\|\mathbf{u}^{\nu^{\prime}}\|\,{\text{\rm d}}s, (4.23)

for all tIt\in I.

Fix ν0>0\nu_{0}>0 such that all parameters ν\nu^{\prime} satisfy νν0\nu^{\prime}\leq\nu_{0}. From (4.18), we can bound the last term in the right-hand side of (4.1.2) by

2ν1/2exp(2t0td(𝐯)Ldτ)𝐯L2(t0,t;V)(νt0t𝐮ν2ds)1/2cν1/2exp(2t0td(𝐯)Ldτ)𝐯L2(t0,t;V)eν0λ12(tt0)[|𝐮ν(t0)|2+1ν0λ1𝐟L2(t0,t;H)2]1/2.2\nu^{\prime 1/2}\exp\left(2\int_{t_{0}}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)\|\mathbf{v}\|_{L^{2}(t_{0},t;V)}\left(\nu^{\prime}\int_{t_{0}}^{t}\|\mathbf{u}^{\nu^{\prime}}\|^{2}\,{\text{\rm d}}s\right)^{1/2}\\ \leq c\nu^{\prime 1/2}\exp\left(2\int_{t_{0}}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)\|\mathbf{v}\|_{L^{2}(t_{0},t;V)}e^{\frac{\nu_{0}\lambda_{1}}{2}(t-t_{0})}\left[|\mathbf{u}^{\nu^{\prime}}(t_{0})|^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right]^{1/2}. (4.24)

Since the net {𝐮ν(t0)}ν\{\mathbf{u}^{\nu^{\prime}}(t_{0})\}_{\nu^{\prime}} is in the compact set KK, and hence is bounded in HH, then the expression in the right-hand side of (4.24) vanishes as ν0\nu^{\prime}\to 0.

Moreover, since 𝐮ν𝐮\mathbf{u}^{\nu^{\prime}}\to\mathbf{u} in 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}), then together with the bound (4.18) it follows that the second term in the right-hand side of (4.1.2) converges as ν0\nu^{\prime}\to 0 to

2t0texp(2std(𝐯)Ldτ)(𝐮𝐯,E(𝐯)+𝐟)ds.\displaystyle 2\int_{t_{0}}^{t}\exp\left(2\int_{s}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)(\mathbf{u}-\mathbf{v},E(\mathbf{v})+\mathbf{f})\,{\text{\rm d}}s.

Additionally, again since {𝐮ν(t0)}νK\{\mathbf{u}^{\nu^{\prime}}(t_{0})\}_{\nu^{\prime}}\subset K then, modulo a subnet, we have the strong convergence: 𝐮ν(t0)𝐮(t0)\mathbf{u}^{\nu^{\prime}}(t_{0})\to\mathbf{u}(t_{0}) in HH. Combining these facts, we obtain by taking the lim inf\liminf as ν0\nu^{\prime}\to 0 in (4.1.2) that

|𝐮(t)𝐯(t)|2lim infν0|𝐮ν(t)𝐯(t)|2exp(2t0td(𝐯)Lds)|𝐮(t0)𝐯(t0)|2+2t0texp(2std(𝐯)Ldτ)(𝐮𝐯,E(𝐯)+𝐟)ds,|\mathbf{u}(t)-\mathbf{v}(t)|^{2}\leq\liminf_{\nu^{\prime}\to 0}|\mathbf{u}^{\nu^{\prime}}(t)-\mathbf{v}(t)|^{2}\leq\exp\left(2\int_{t_{0}}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}s\right)|\mathbf{u}(t_{0})-\mathbf{v}(t_{0})|^{2}\\ +2\int_{t_{0}}^{t}\exp\left(2\int_{s}^{t}\|d^{-}(\mathbf{v})\|_{L^{\infty}}\,{\text{\rm d}}\tau\right)(\mathbf{u}-\mathbf{v},E(\mathbf{v})+\mathbf{f})\,{\text{\rm d}}s,

which shows (4.17), and concludes the proof. ∎

4.2. Convergence of statistical solutions of 2D Navier-Stokes to 2D Euler

In this section, we verify the assumptions of Theorem 3.2 to deduce the convergence of a net of trajectory statistical solutions of the 2D NSE towards a trajectory statistical solution of the 2D Euler equations, as stated in Theorem 4.1 below.

We start by fixing the required setting from Theorem 3.2. Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0}. Take X=(Wσ1,r)wX=(W_{\sigma}^{1,r})_{\rm{w}}, for any given 1<r<1<r<\infty, and define, for each fixed ν>0\nu>0,

Sν:(Wσ1,r)w\displaystyle S_{\nu}:(W_{\sigma}^{1,r})_{\rm{w}} \displaystyle\rightarrow 𝒞loc(I,(Wσ1,r)w)\displaystyle\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}})
𝐮0\displaystyle\mathbf{u}_{0} \displaystyle\mapsto 𝐮ν,\displaystyle\mathbf{u}^{\nu}, (4.25)

where 𝐮ν\mathbf{u}^{\nu} is the unique weak solution of (4.1) on II in the sense of Definition 4.1 satisfying 𝐮ν(t0)=𝐮0\mathbf{u}^{\nu}(t_{0})=\mathbf{u}_{0}. Thus, the operator Pν=Πt0SνP_{\nu}=\Pi_{t_{0}}S_{\nu}, as defined in Theorem 3.2, is the identity operator.

Note that since Wσ1,rW_{\sigma}^{1,r} is a separable Banach space then every Borel probability measure on (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}} is also a Borel probability measure on Wσ1,rW_{\sigma}^{1,r} (and vice-versa), and hence tight in Wσ1,rW_{\sigma}^{1,r}, i.e. inner regular with respect to the family of compact subsets of Wσ1,rW_{\sigma}^{1,r} (see Section 2.2), which are also compact sets in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}. In summary, every Borel probability measure on (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}} (or, equivalently, Wσ1,rW_{\sigma}^{1,r}) is tight in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}. For this reason, we consider μ0\mu_{0} as any Borel probability measure in Wσ1,rW_{\sigma}^{1,r} in the statements of this section.

Then, given a Borel probability measure μ0\mu_{0} on Wσ1,r,W_{\sigma}^{1,r}, we set, for simplicity, μν=μ0\mu_{\nu}=\mu_{0} for all ν>0\nu>0. Thus, Pνμν=μν=μ0P_{\nu}\mu_{\nu}=\mu_{\nu}=\mu_{0} and assumption (H1) is immediately satisfied. Also, from the tightness of μ0\mu_{0} we obtain the existence of a sequence {Kn}n\{K_{n}\}_{n\in\mathbb{N}} of compact sets in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}} satisfying (H3).

As we shall see, assumptions (H2), (H4) and (H5) actually hold for any compact set KK of (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, with 𝒰\mathcal{U} defined as

𝒰I={𝐮𝒞loc(I,(Wσ1,r)w):𝐮 is a weak solution of the 2D Euler equations(4.2) on I in the sense of Definition 4.2}.\displaystyle\mathcal{U}_{I}=\left\{\begin{array}[]{cc}\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}):&\mathbf{u}\mbox{ is a weak solution of the 2D Euler equations}\\ &\mbox{\eqref{ee} on $I$ in the sense of \lx@cref{creftype~refnum}{Eweak}}\end{array}\right\}. (4.28)

For simplicity, we assume throughout this section that rr is restricted to the range 2r<2\leq r<\infty. In particular, this allows us to obtain the bound (4.11) below for the LrL^{r}-norm of the vorticity associated with a weak solution of the 2D NSE. We note, however, that the case 1<r<21<r<2 can also be treated, by appealing to the notion of renormalized solutions, see [54, Section 4.1]. This case is indeed considered in the work [67], where an analogous convergence result for trajectory statistical solutions of the 2D NSE towards a trajectory statistical solution of 2D Euler is obtained, albeit under the assumption of zero forcing term and with a slightly different setting than ours, particularly concerning the definition of XX and the fact that the spatial domain is taken as 2\mathbb{R}^{2}.

The next proposition shows that condition (H2) from Theorem 3.2 holds true in this context.

Proposition 4.5.

Let KK be a compact set in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, 2r<2\leq r<\infty. Then, for each ν>0\nu>0, the operator Sν|K:K𝒞loc(I,(Wσ1,r)w)S_{\nu}|_{K}:K\rightarrow\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) is continuous.

Proof.

Since the weak topology is metrizable on bounded subsets of Wσ1,rW_{\sigma}^{1,r}, it suffices to show that, for any given 𝐮0K\mathbf{u}_{0}\in K and any sequence {𝐮0,n}n\{\mathbf{u}_{0,n}\}_{n\in\mathbb{N}} in KK converging weakly to 𝐮0,\mathbf{u}_{0}, it follows that Sν(𝐮0,n)S_{\nu}(\mathbf{u}_{0,n}) converges to Sν(𝐮0)S_{\nu}(\mathbf{u}_{0}) in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}).

Consider any compact subinterval JIJ\subset I with left endpoint t0t_{0}. It is sufficient to show that Sν(𝐮0,n)S_{\nu}(\mathbf{u}_{0,n}) converges to Sν(𝐮0)S_{\nu}(\mathbf{u}_{0}) in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). We first show that {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n})\}_{n} is relatively compact in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}).

Since {𝐮0,n}n\{\mathbf{u}_{0,n}\}_{n} is contained in the compact set KK, then {𝐮0,n}n\{\mathbf{u}_{0,n}\}_{n} is a bounded sequence in Wσ1,rW_{\sigma}^{1,r}. Thus, from (4.7) and (4.11) it follows that {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n})\}_{n} is uniformly bounded in 𝒞(J,Wσ1,r)\mathcal{C}(J,W_{\sigma}^{1,r}). We may thus consider a ball BWσ1,r(R)B_{W_{\sigma}^{1,r}}(R) in Wσ1,rW_{\sigma}^{1,r}, R>0R>0, such that Sν(𝐮0,n)(t)BWσ1,r(R)S_{\nu}(\mathbf{u}_{0,n})(t)\subset B_{W_{\sigma}^{1,r}}(R) for all nn\in\mathbb{N} and tJt\in J. Note also that, from (4.1), it follows that {tSν(𝐮0,n)}n\{\partial_{t}S_{\nu}(\mathbf{u}_{0,n})\}_{n} is uniformly bounded in L2(J,V)L^{2}(J,V^{\prime}).

Let 𝐰(Wσ1,r)\mathbf{w}\in(W_{\sigma}^{1,r})^{\prime} and ε>0\varepsilon>0. Since VV is dense in (Wσ1,r)(W_{\sigma}^{1,r})^{\prime}, we may take 𝐯V\mathbf{v}\in V such that 𝐯𝐰(Wσ1,r)<ε/(4R)\|\mathbf{v}-\mathbf{w}\|_{(W_{\sigma}^{1,r})^{\prime}}<\varepsilon/(4R). Hence, for all nn\in\mathbb{N} and s<ts<t, we have

|𝐰,Sν(𝐮0,n)(t)Sν(𝐮0,n)(s)(Wσ1,r),Wσ1,r||𝐰𝐯,Sν(𝐮0,n)(t)Sν(𝐮0,n)(s)(Wσ1,r),Wσ1,r|+|Sν(𝐮0,n)(t)Sν(𝐮0,n)(s),𝐯V,V|.|\langle\mathbf{w},S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s)\rangle_{(W_{\sigma}^{1,r})^{\prime},W_{\sigma}^{1,r}}|\\ \qquad\leq|\langle\mathbf{w}-\mathbf{v},S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s)\rangle_{(W_{\sigma}^{1,r})^{\prime},W_{\sigma}^{1,r}}|+|\langle S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s),\mathbf{v}\rangle_{V^{\prime},V}|. (4.29)

The first term is estimated as

|𝐰𝐯,Sν(𝐮0,n)(t)Sν(𝐮0,n)(s)(Wσ1,r),Wσ1,r|2R𝐰𝐯(Wσ1,r)<ε2.\displaystyle|\langle\mathbf{w}-\mathbf{v},S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s)\rangle_{(W_{\sigma}^{1,r})^{\prime},W_{\sigma}^{1,r}}|\leq 2R\|\mathbf{w}-\mathbf{v}\|_{(W_{\sigma}^{1,r})^{\prime}}<\frac{\varepsilon}{2}.

Regarding the second term in (4.29), we have

|Sν(𝐮0,n)(t)Sν(𝐮0,n)(s),𝐯V,V|=|sttSν(𝐮0,n)(τ)dτ,𝐯V,V|(sttSν(𝐮0,n)(τ)Vdτ)𝐯|ts|1/2tSν(𝐮0,n)L2(J,V)𝐯C|ts|1/2𝐯.|\langle S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s),\mathbf{v}\rangle_{V^{\prime},V}|=\left|{\left\langle\int_{s}^{t}\partial_{t}S_{\nu}(\mathbf{u}_{0,n})(\tau){\text{\rm d}}\tau,\mathbf{v}\right\rangle}_{V^{\prime},V}\right|\\ \leq\left(\int_{s}^{t}\|\partial_{t}S_{\nu}(\mathbf{u}_{0,n})(\tau)\|_{V^{\prime}}{\text{\rm d}}\tau\right)\|\mathbf{v}\|\leq|t-s|^{1/2}\|\partial_{t}S_{\nu}(\mathbf{u}_{0,n})\|_{L^{2}(J,V^{\prime})}\|\mathbf{v}\|\\ \leq C|t-s|^{1/2}\|\mathbf{v}\|. (4.30)

Hence, it follows from (4.29)-(4.30) that, for all s,tJs,t\in J with |ts|<ε2/(2C𝐯)2,|t-s|<\varepsilon^{2}/(2C\|\mathbf{v}\|)^{2},

|𝐰,Sν(𝐮0,n)(t)Sν(𝐮0,n)(s)(Wσ1,r),Wσ1,r|<ε2+C|ts|1/2𝐯<ε.\displaystyle|\langle\mathbf{w},S_{\nu}(\mathbf{u}_{0,n})(t)-S_{\nu}(\mathbf{u}_{0,n})(s)\rangle_{(W_{\sigma}^{1,r})^{\prime},W_{\sigma}^{1,r}}|<\frac{\varepsilon}{2}+C|t-s|^{1/2}\|\mathbf{v}\|<\varepsilon.

Since 𝐰\mathbf{w} and ε\varepsilon are arbitrary, this implies that {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n})\}_{n} is equicontinuous in 𝒞(J,BWσ1,r(R)w)\mathcal{C}(J,B_{W_{\sigma}^{1,r}}(R)_{\textrm{\rm w}}). Moreover, since BWσ1,r(R)B_{W_{\sigma}^{1,r}}(R) is a compact set in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, we also have that, for each fixed tJt\in J, {Sν(𝐮0,n)(t)}n\{S_{\nu}(\mathbf{u}_{0,n})(t)\}_{n} is relatively compact in BWσ1,r(R)wB_{W_{\sigma}^{1,r}}(R)_{\textrm{\rm w}}. Therefore, by the Arzelà-Ascoli theorem, it follows that {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n})\}_{n} is relatively compact in 𝒞(J,BWσ1,r(R)w)\mathcal{C}(J,B_{W_{\sigma}^{1,r}}(R)_{\textrm{\rm w}}), and hence in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}).

Thus, there exists a subsequence {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n^{\prime}})\}_{n^{\prime}} and 𝐮~𝒞(J,(Wσ1,r)w)\tilde{\mathbf{u}}\in\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}) such that Sν(𝐮0,n)𝐮~S_{\nu}(\mathbf{u}_{0,n^{\prime}})\to\tilde{\mathbf{u}} in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). In particular, Sν(𝐮0,n)(t0)=𝐮0,nS_{\nu}(\mathbf{u}_{0,n^{\prime}})(t_{0})=\mathbf{u}_{0,n^{\prime}} converges weakly to 𝐮~(t0)\tilde{\mathbf{u}}(t_{0}) in Wσ1,rW_{\sigma}^{1,r} and, by uniqueness of the limit, 𝐮~(t0)=𝐮0\tilde{\mathbf{u}}(t_{0})=\mathbf{u}_{0}. Also, by 4.2, we have that 𝐮~\tilde{\mathbf{u}} is a weak solution of the 2D NSE (4.1) on II. By uniqueness of solutions, it follows that 𝐮~=Sν(𝐮0)\tilde{\mathbf{u}}=S_{\nu}(\mathbf{u}_{0}). Then, by a contradiction argument, we obtain that in fact the entire sequence {Sν(𝐮0,n)}n\{S_{\nu}(\mathbf{u}_{0,n})\}_{n} converges to Sν(𝐮0)S_{\nu}(\mathbf{u}_{0}) in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). This concludes the proof. ∎

To verify assumptions (H4) and (H5), we fix ν0>0\nu_{0}>0 and introduce the following auxiliary space. Let R>0R>0 and JIJ\subseteq I be an interval closed and bounded on the left with left endpoint t0t_{0}, and consider the following inequalities for 𝐮𝒞(J,(Wσ1,r)w)\mathbf{u}\in\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}):

𝐮(t)Lr(Rr+(ν0λ1)1r𝐟Lr(t0,t;Lr)r)1/re(r1)ν0λ1(tt0)/r,\displaystyle\|\nabla^{\perp}\cdot\mathbf{u}(t)\|_{L^{r}}\leq\left(R^{r}+(\nu_{0}\lambda_{1})^{1-r}\|\nabla^{\perp}\cdot\mathbf{f}\|^{r}_{L^{r}(t_{0},t;L^{r})}\right)^{1/r}e^{(r-1)\nu_{0}\lambda_{1}(t-t_{0})/r}, (4.31)

and

t𝐮L2(t0,t;V)cλ11/2𝐟L2(t0,t;H)+cλ11/2+1/r[ν0+(R2+1ν0λ1𝐟L2(t0,t;H)2)eν0λ1(tt0)]×(Rr+(ν0λ1)1r𝐟Lr(t0,t;Lr)r)e(r1)ν0λ1(tt0),\|\partial_{t}\mathbf{u}\|_{L^{2}(t_{0},t;V^{\prime})}\leq c\lambda_{1}^{-1/2}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}+\\ c\lambda_{1}^{-1/2+1/r}\left[\nu_{0}+\left(R^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right)e^{\nu_{0}\lambda_{1}(t-t_{0})}\right]\times\\ \left(R^{r}+(\nu_{0}\lambda_{1})^{1-r}\|\nabla^{\perp}\cdot\mathbf{f}\|_{L^{r}(t_{0},t;L^{r})}^{r}\right)e^{(r-1)\nu_{0}\lambda_{1}(t-t_{0})}, (4.32)

for tJt\in J, where c>0c>0 is a universal constant. Then, we define

𝒴J(R)={𝐮𝒞(J,(Wσ1,r)w):𝐮 satisfies (4.31) and (4.32) for all tJ}.\displaystyle\mathcal{Y}_{J}(R)=\left\{\mathbf{u}\in\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}})\,:\,\mathbf{u}\mbox{ satisfies }\eqref{ineq:sup:Wrw}\mbox{ and }\eqref{ineq:Dt:L2Vp}\mbox{ for all }t\in J\right\}. (4.33)

Note that, for all 0<νν00<\nu\leq\nu_{0} and for every initial datum in BWσ1,r(R)B_{W_{\sigma}^{1,r}}(R), the restriction to JJ of the corresponding weak solution of the 2D ν\nu-NSE belongs to 𝒴J(R)\mathcal{Y}_{J}(R).

We observe that given any sequence of compact subintervals JnIJ_{n}\subset I, nn\in\mathbb{N}, each with left endpoint t0t_{0} and such that I=nJnI=\bigcup_{n}J_{n}, then

𝒴I(R)=n=1ΠJn1𝒴Jn(R),\displaystyle\mathcal{Y}_{I}(R)=\bigcap_{n=1}^{\infty}\Pi_{J_{n}}^{-1}\mathcal{Y}_{J_{n}}(R), (4.34)

where ΠJn:𝒞(I,(Wσ1,r)w)𝒞(Jn,(Wσ1,r)w)\Pi_{J_{n}}:\mathcal{C}(I,(W_{\sigma}^{1,r})_{\rm{w}})\rightarrow\mathcal{C}(J_{n},(W_{\sigma}^{1,r})_{\rm{w}}) denotes the restriction operator on JnJ_{n} defined in (2.2). We now show that this auxiliary space is compact.

Lemma 4.1.

Let R>0R>0 and let JIJ\subset I be a compact subinterval with left endpoint t0t_{0}. Then, 𝒴J(R)\mathcal{Y}_{J}(R) is a compact subset of 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). Consequently, 𝒴I(R)\mathcal{Y}_{I}(R) is a compact subset of 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}).

Proof.

First, from the definition of 𝒴J(R)\mathcal{Y}_{J}(R) in (4.33) it follows that, for all 𝐮𝒴J(R)\mathbf{u}\in\mathcal{Y}_{J}(R), it holds

𝐮(t)LrRJ and t𝐮L2(t0,t;V)R~J,for all tJ,\displaystyle\|\nabla^{\perp}\cdot\mathbf{u}(t)\|_{L^{r}}\leq R_{J}\quad\mbox{ and }\quad\|\partial_{t}\mathbf{u}\|_{L^{2}(t_{0},t;V^{\prime})}\leq\tilde{R}_{J},\quad\mbox{for all }t\in J, (4.35)

where RJ,R~JR_{J},\tilde{R}_{J} are positive constants which depend on JJ, but are independent of tt. In particular, the first inequality in (4.35) implies that 𝒴J(R)𝒞(J,BWσ1,r(RJ)w)\mathcal{Y}_{J}(R)\subset\mathcal{C}(J,B_{W_{\sigma}^{1,r}}(R_{J})_{\textrm{\rm w}}), so that 𝒴J(R)\mathcal{Y}_{J}(R) is metrizable, and it suffices to show that it is sequentially compact.

Let {𝐮n}n\{\mathbf{u}_{n}\}_{n} be a sequence in 𝒴J(R)\mathcal{Y}_{J}(R). Then, {𝐮n}n\{\mathbf{u}_{n}\}_{n} is uniformly bounded in 𝒞(J,Wσ1,r)\mathcal{C}(J,W_{\sigma}^{1,r}) and {t𝐮n}n\{\partial_{t}\mathbf{u}_{n}\}_{n} is uniformly bounded in L2(t0,t,V)L^{2}(t_{0},t,V^{\prime}) for all tJt\in J. With a similar argument as in the proof of 4.5, it follows that {𝐮n}n\{\mathbf{u}_{n}\}_{n} is relatively compact in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). Then, we can show that there exists a subsequence {𝐮n}n\{\mathbf{u}_{n^{\prime}}\}_{n^{\prime}} and 𝐮𝒞(J,(Wσ1,r)w)\mathbf{u}\in\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}) with t𝐮L2(t0,t,V)\partial_{t}\mathbf{u}\in L^{2}(t_{0},t,V^{\prime}) such that

𝐮n𝐮 in 𝒞(J,(Wσ1,r)w),\displaystyle\mathbf{u}_{n^{\prime}}\to\mathbf{u}\mbox{ in }\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}), (4.36)
t𝐮nt𝐮 in L2(t0,t;V), for all tJ.\displaystyle\partial_{t}\mathbf{u}_{n^{\prime}}\rightharpoonup\partial_{t}\mathbf{u}\mbox{ in }L^{2}(t_{0},t;V^{\prime}),\quad\mbox{ for all }t\in J. (4.37)

To see this, first let {𝐮n}n\{\mathbf{u}_{n^{\prime}}\}_{n^{\prime}} be a subsequence for which 𝐮n𝐮\mathbf{u}_{n^{\prime}}\to\mathbf{u} in 𝒞(J,(Wσ1,r)w)\mathcal{C}(J,(W_{\sigma}^{1,r})_{\rm{w}}). Then, consider a sequence of points {tk}kJ\{t_{k}\}_{k\in\mathbb{N}}\subset J that is dense in JJ. We have that {t𝐮n}\{\partial_{t}\mathbf{u}_{n^{\prime}}\} is uniformly bounded in L2(t0,tk;V)L^{2}(t_{0},t_{k};V^{\prime}) for all kk. Then, by a diagonalization argument, we may construct a further subsequence of {𝐮n}n\{\mathbf{u}_{n^{\prime}}\}_{n^{\prime}}, which we still denote as {𝐮n}n\{\mathbf{u}_{n^{\prime}}\}_{n^{\prime}} for simplicity, such that t𝐮nt𝐮\partial_{t}\mathbf{u}_{n^{\prime}}\rightharpoonup\partial_{t}\mathbf{u} in L2(t0,tk;V)L^{2}(t_{0},t_{k};V^{\prime}) for all kk. Due to the continuity of the functional tt0tt𝐮nt𝐮,𝐯V,V𝑑st\mapsto\int_{t_{0}}^{t}\langle\partial_{t}\mathbf{u}_{n^{\prime}}-\partial_{t}\mathbf{u},\mathbf{v}\rangle_{V^{\prime},V}ds for any 𝐯L2(J;V)\mathbf{v}\in L^{2}(J;V^{\prime}), and the density of {tk}\{t_{k}\} in JJ, we thus obtain that in fact t𝐮nt𝐮\partial_{t}\mathbf{u}_{n^{\prime}}\rightharpoonup\partial_{t}\mathbf{u} in L2(t0,t;V)L^{2}(t_{0},t;V^{\prime}) for all tJt\in J.

With the convergences in (4.36) and (4.37), we can pass to the limit in the inequalities from the definition of 𝒴J(R)\mathcal{Y}_{J}(R) and conclude that 𝐮𝒴J(R)\mathbf{u}\in\mathcal{Y}_{J}(R). This concludes the proof of the compactness of 𝒴J(R)\mathcal{Y}_{J}(R).

Consequently, in view of the characterization (4.34), it follows by employing again a standard diagonalization argument that 𝒴I(R)\mathcal{Y}_{I}(R) is compact in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}). ∎

Next, we invoke Lemma 4.1 to verify assumption (H4).

Proposition 4.6.

Let KK be a compact set in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, 2r<2\leq r<\infty. Then, there exists a compact set 𝒦𝒞loc(I,(Wσ1,r)w)\mathcal{K}\subset\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) such that

Sν(K)𝒦, for all ν(0,ν0].\displaystyle S_{\nu}(K)\subset\mathcal{K},\quad\mbox{ for all }\nu\in(0,\nu_{0}].
Proof.

Let R>0R>0 be such that KBWσ1,r(R)K\subset B_{W_{\sigma}^{1,r}}(R) and let {Jn}n\{J_{n}\}_{n} be any sequence of compact subintervals of II, each with left endpoint t0t_{0}, such that I=n=1JnI=\bigcup_{n=1}^{\infty}J_{n}. From the estimates (4.11) and (4.1) in 4.1, it follows that ΠJnSν(K)𝒴Jn(R)\Pi_{J_{n}}S_{\nu}(K)\subset\mathcal{Y}_{J_{n}}(R) for all nn\in\mathbb{N} and for every ν(0,ν0]\nu\in(0,\nu_{0}]. Thus, from the characterization (4.34), we deduce that Sν(K)S_{\nu}(K) is contained in the compact set 𝒦=𝒴I(R)\mathcal{K}=\mathcal{Y}_{I}(R) for all ν(0,ν0]\nu\in(0,\nu_{0}]. ∎

Finally, we verify that assumption (H5) from Theorem 3.2 is satisfied.

Proposition 4.7.

Let KK be a compact set in (Wσ1,r)w(W_{\sigma}^{1,r})_{\rm{w}}, 2r<2\leq r<\infty. Then,

lim supν0Sν(K)𝒰I,\limsup_{\nu\to 0}S_{\nu}(K)\subset\mathcal{U}_{I},

with 𝒰I\mathcal{U}_{I} as defined in (4.28).

Proof.

From the proof of 4.6, we know that

lim supν0Sν(K)𝒴I(R),\displaystyle\limsup_{\nu\to 0}S_{\nu}(K)\subset\mathcal{Y}_{I}(R),

for any R>0R>0 such that KBWσ1,r(R)K\subset B_{W_{\sigma}^{1,r}}(R). Since 𝒴I(R)\mathcal{Y}_{I}(R) is a metrizable space, given 𝐮lim supν0Sν(K),\mathbf{u}\in\limsup_{\nu\to 0}S_{\nu}(K), there exists a sequence {𝐮νj}j\{\mathbf{u}^{\nu_{j}}\}_{j} such that νj(0,ν0]\nu_{j}\in(0,\nu_{0}] and 𝐮νjSνj(K)\mathbf{u}^{\nu_{j}}\in S_{\nu_{j}}(K) for all jj\in\mathbb{N}, with νj0\nu_{j}\to 0 and 𝐮νj𝐮\mathbf{u}^{\nu_{j}}\to\mathbf{u} in 𝒴I(R)\mathcal{Y}_{I}(R) as jj\to\infty. Moreover, since the metric in 𝒴I(R)\mathcal{Y}_{I}(R) is compatible with the topology in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}), we also have that 𝐮νj𝐮\mathbf{u}^{\nu_{j}}\to\mathbf{u} in 𝒞loc(I,(Wσ1,r)w)\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) as jj\to\infty. By 4.2, this implies that 𝐮𝒰I\mathbf{u}\in\mathcal{U}_{I}, as desired. ∎

Having verified all the required assumptions, we may now apply Theorem 3.2 with the choices of XX, 𝒰\mathcal{U}, {Sε}ε\{S_{\varepsilon}\}_{\varepsilon\in\mathcal{E}}, and {με}ε\{\mu_{\varepsilon}\}_{\varepsilon\in\mathcal{E}} fixed in the beginning of this section and obtain the following result on the convergence of trajectory statistical solutions of the 2D Navier-Stokes equations to a trajectory statistical solution of the 2D Euler equations in the inviscid limit.

Theorem 4.1.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and assume 𝐟Lloc2(I,V).\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,V^{\prime}). Fix r[2,)r\in[2,\infty), and let Sν:(Wσ1,r)w𝒞loc(I,(Wσ1,r)w)S_{\nu}:(W_{\sigma}^{1,r})_{\rm{w}}\to\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}), ν>0\nu>0, and 𝒰I𝒞loc(I,(Wσ1,r)w)\mathcal{U}_{I}\subset\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) be defined as in (4.2) and (4.28), respectively. Then, given a Borel probability measure μ0\mu_{0} on Wσ1,rW_{\sigma}^{1,r}, the net {Sνμ0}ν>0\{S_{\nu}\mu_{0}\}_{\nu>0} has a subnet that converges as ν0\nu\to 0, with respect to the weak-star semicontinuity topology, to a 𝒰I\mathcal{U}_{I}-trajectory statistical solution ρ\rho of the 2D Euler equations that satisfies Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

Clearly, Theorem 4.1 thus yields the existence of a 𝒰I\mathcal{U}_{I}-trajectory statistical solution of the 2D Euler equations satisfying a given initial datum. To emphasize this fact, we state it as the following corollary.

Corollary 4.1.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and assume 𝐟Lloc2(I,V).\mathbf{f}\in L_{\textrm{\rm loc}}^{2}(I,V^{\prime}). Fix r[2,)r\in[2,\infty), and let 𝒰I𝒞loc(I,(Wσ1,r)w)\mathcal{U}_{I}\subset\mathcal{C}_{\textrm{\rm loc}}(I,(W_{\sigma}^{1,r})_{\rm{w}}) be as defined in (4.28). Then, given a Borel probability measure μ0\mu_{0} on Wσ1,rW_{\sigma}^{1,r}, there exists a 𝒰I\mathcal{U}_{I}-trajectory statistical solution ρ\rho of the 2D Euler equations satisfying the initial condition Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

4.3. Convergence of statistical solutions of 3D Navier-Stokes to 3D Euler

In this section, we show the existence of a trajectory statistical solution of the 3D Euler equations starting from any given initial measure μ0\mu_{0} on HH. This is done by considering, for each ν>0\nu>0, a trajectory statistical solution ρν\rho_{\nu} of the 3D NSE starting from μ0\mu_{0}, and constructing a suitable family of compact sets for which the assumptions of Theorem 3.1 above are verified.

Here we consider the setting of Theorem 3.1 with the following choices. Let X=HwX=H_{\textrm{\rm w}}, and let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0}. Moreover, for a fixed forcing term 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H), we define the following corresponding sets of solutions of the 3D ν\nu-Navier-Stokes equations, with ν>0\nu>0, and 3D Euler equations, in 𝒳=𝒞loc(I,Hw)\mathcal{X}=\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}):

𝒰Iν={𝐮ν𝒞loc(I,Hw):𝐮ν is a weak solution of the 3D ν-Navier-Stokes equations (4.1) on I in the sense of Definition 4.3},\displaystyle\mathcal{U}^{\nu}_{I}=\left\{\begin{array}[]{cc}\mathbf{u}^{\nu}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}):&\mathbf{u}^{\nu}\mbox{ is a weak solution of the 3D $\nu$-Navier-Stokes}\\ &\mbox{ equations \eqref{nse} on $I$ in the sense of \lx@cref{creftype~refnum}{3NSweak}}\end{array}\right\}, (4.40)

and

𝒰Idiss={𝐮𝒞loc(I,Hw):𝐮 is a dissipative solution of the 3D Euler equations(4.2) on I in the sense of Definition 4.4}.\displaystyle\mathcal{U}^{\textrm{\rm diss}}_{I}=\left\{\begin{array}[]{cc}\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}):&\mathbf{u}\mbox{ is a dissipative solution of the 3D Euler equations}\\ &\eqref{ee}\mbox{ on $I$ in the sense of \lx@cref{creftype~refnum}{Edissipative}}\end{array}\right\}. (4.43)

As in Section 4.2, we also define an auxiliary space in view of the a priori estimates from 4.3. Namely, for fixed ν0>0\nu_{0}>0, R>0R>0, and any subinterval JIJ\subseteq I that is closed and bounded on the left with left endpoint t0t_{0}, consider the following inequalities for 𝐮𝒞loc(J,Hw)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(J,H_{\textrm{\rm w}}):

|𝐮(t)|2(R2+1ν0λ1𝐟L2(t0,t;H)2)eν0λ1(tt0),\displaystyle|\mathbf{u}(t)|^{2}\leq\left(R^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right)e^{\nu_{0}\lambda_{1}(t-t_{0})}, (4.44)

for tJt\in J, and

𝐮(t)𝐮(s)(Wσ1,)c|ts|1/2ν01/2λ13/4eν0λ1(tt0)2[R2+1ν0λ1𝐟L2(t0,t;H)2]1/2\displaystyle\|\mathbf{u}(t)-\mathbf{u}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\leq c|t-s|^{1/2}\nu_{0}^{1/2}\lambda_{1}^{-3/4}e^{\frac{\nu_{0}\lambda_{1}(t-t_{0})}{2}}\left[R^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right]^{1/2}
+|ts|eν0λ1(tt0)[R2+1ν0λ1𝐟L2(t0,t;H)2],\displaystyle+|t-s|e^{\nu_{0}\lambda_{1}(t-t_{0})}\left[R^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(t_{0},t;H)}^{2}\right], (4.45)

for s,tJs,t\in J with s<ts<t, where c>0c>0 is a fixed universal constant. Then, we define

𝒴J(R)={𝐮𝒞loc(J,Hw):𝐮 satisfies (4.44) for all tJ and (4.3) for all s<t in J}.\displaystyle\mathcal{Y}_{J}(R)=\left\{\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(J,H_{\textrm{\rm w}}):\mathbf{u}\mbox{ satisfies }\eqref{ineq:YI:1}\mbox{ for all $t\in J$ and }\eqref{ineq:YI:2}\mbox{ for all $s<t$ in $J$}\right\}. (4.46)

As in the two-dimensional case, note that for all 0<νν00<\nu\leq\nu_{0} and for every initial datum in BH(R)B_{H}(R), the restriction to JJ of the corresponding weak solution of the 3D ν\nu-NSE belongs to 𝒴J(R)\mathcal{Y}_{J}(R).

From the definition (4.46), it follows that for any nondecreasing sequence of compact subintervals JnIJ_{n}\subset I, nn\in\mathbb{N}, each with left endpoint t0t_{0} and such that I=nJnI=\bigcup_{n}J_{n}, we may write

𝒴I(R)=n=1ΠJn1𝒴Jn(R),\displaystyle\mathcal{Y}_{I}(R)=\bigcap_{n=1}^{\infty}\Pi_{J_{n}}^{-1}\mathcal{Y}_{J_{n}}(R), (4.47)

where we recall from (2.2) that ΠJn\Pi_{J_{n}} denotes the operator that takes any function 𝐮𝒞loc(I,Hw)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) to its restriction to JnJ_{n}, namely ΠJn𝐮=𝐮|Jn𝒞(Jn,Hw)\Pi_{J_{n}}\mathbf{u}=\mathbf{u}|_{J_{n}}\in\mathcal{C}(J_{n},H_{\textrm{\rm w}}).

We have the following compactness result.

Lemma 4.2.

Let R>0R>0 and ν0>0\nu_{0}>0. Then, for every compact subinterval JIJ\subset I with left endpoint t0t_{0}, 𝒴J(R)\mathcal{Y}_{J}(R) is a compact subset of 𝒞(J,Hw)\mathcal{C}(J,H_{\textrm{\rm w}}). Consequently, 𝒴I(R)\mathcal{Y}_{I}(R) is a compact subset of 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}).

Proof.

Let JIJ\subset I be a compact subinterval with left endpoint t0t_{0}. Denote

RJ=(R2+1ν0λ1𝐟L2(J;H)2)1/2eν0λ12|J|.\displaystyle R_{J}=\left(R^{2}+\frac{1}{\nu_{0}\lambda_{1}}\|\mathbf{f}\|_{L^{2}(J;H)}^{2}\right)^{1/2}e^{\frac{\nu_{0}\lambda_{1}}{2}|J|}.

According to the definition in (4.46), it follows that every 𝐮𝒴J(R)\mathbf{u}\in\mathcal{Y}_{J}(R) satisfies:

|𝐮(t)|RJ, for all tJ,\displaystyle|\mathbf{u}(t)|\leq R_{J},\quad\mbox{ for all }t\in J, (4.48)

and

𝐮(t)𝐮(s)(Wσ1,)\displaystyle\|\mathbf{u}(t)-\mathbf{u}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}} c|ts|1/2ν01/2λ13/4RJ+|ts|RJ2\displaystyle\leq c|t-s|^{1/2}\nu_{0}^{1/2}\lambda_{1}^{-3/4}R_{J}+|t-s|R_{J}^{2}
C|ts|1/2 for all s,tJ.\displaystyle\leq C|t-s|^{1/2}\quad\mbox{ for all }s,t\in J. (4.49)

In particular, (4.48) implies that 𝒴J(R)𝒞(J,BH(RJ)w)\mathcal{Y}_{J}(R)\subset\mathcal{C}(J,B_{H}(R_{J})_{\textrm{\rm w}}). Thus, 𝒴J(R)\mathcal{Y}_{J}(R) is metrizable, and it suffices to show that 𝒴J(R)\mathcal{Y}_{J}(R) is sequentially compact.

Let {𝐮n}n\{\mathbf{u}_{n}\}_{n} be a sequence in 𝒴J(R)\mathcal{Y}_{J}(R). We first show that {𝐮n}n\{\mathbf{u}_{n}\}_{n} is equicontinuous in 𝒞(J,BH(RJ)w)\mathcal{C}(J,B_{H}(R_{J})_{\textrm{\rm w}}). Let 𝐯H\mathbf{v}\in H and ε>0\varepsilon>0 be arbitrary. Since Wσ1,W_{\sigma}^{1,\infty} is dense in HH, we may take 𝐰Wσ1,\mathbf{w}\in W_{\sigma}^{1,\infty} such that |𝐯𝐰|<ε/(4RJ)|\mathbf{v}-\mathbf{w}|<\varepsilon/(4R_{J}). Thus, together with (4.48) and (4.3), we obtain that, for any s,tJ,s,t\in J,

|(𝐮n(t)𝐮n(s),𝐯)|\displaystyle|(\mathbf{u}_{n}(t)-\mathbf{u}_{n}(s),\mathbf{v})| |(𝐮n(t)𝐮n(s),𝐯𝐰)|+|𝐮n(t)𝐮n(s),𝐰(Wσ1,),Wσ1,|\displaystyle\leq|(\mathbf{u}_{n}(t)-\mathbf{u}_{n}(s),\mathbf{v}-\mathbf{w})|+|\langle\mathbf{u}_{n}(t)-\mathbf{u}_{n}(s),\mathbf{w}\rangle_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}|
2RJ|𝐯𝐰|+𝐮n(t)𝐮n(s)(Wσ1,)𝐰Wσ1,\displaystyle\leq 2R_{J}|\mathbf{v}-\mathbf{w}|+\|\mathbf{u}_{n}(t)-\mathbf{u}_{n}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\|\mathbf{w}\|_{W_{\sigma}^{1,\infty}}
<ε2+C|ts|1/2𝐰Wσ1,,\displaystyle<\frac{\varepsilon}{2}+C|t-s|^{1/2}\|\mathbf{w}\|_{W_{\sigma}^{1,\infty}},

so that, if |ts|<ε2/(2C𝐰Wσ1,)2,|t-s|<\varepsilon^{2}/(2C\|\mathbf{w}\|_{W_{\sigma}^{1,\infty}})^{2}, then

|(𝐮n(t)𝐮n(s),𝐯)|<ε, for all n.\displaystyle|(\mathbf{u}_{n}(t)-\mathbf{u}_{n}(s),\mathbf{v})|<\varepsilon,\quad\mbox{ for all }n.

Since 𝐯\mathbf{v} and ε\varepsilon are arbitrary, we deduce that {𝐮n}n\{\mathbf{u}_{n}\}_{n} is equicontinuous in 𝒞(J,BH(RJ)w)\mathcal{C}(J,B_{H}(R_{J})_{\textrm{\rm w}}). Moreover, for each fixed tJt\in J, {𝐮n(t)}nBH(RJ)\{\mathbf{u}_{n}(t)\}_{n}\subset B_{H}(R_{J}), and thus {𝐮n(t)}n\{\mathbf{u}_{n}(t)\}_{n} is relatively compact in BH(RJ)wB_{H}(R_{J})_{\textrm{\rm w}}. Therefore, by the Arzelà-Ascoli theorem, it follows that {𝐮n}n\{\mathbf{u}_{n}\}_{n} is relatively compact in 𝒞(J,BH(RJ)w)\mathcal{C}(J,B_{H}(R_{J})_{\textrm{\rm w}}), and hence in 𝒞(J,Hw)\mathcal{C}(J,H_{\textrm{\rm w}}).

Thus, there exists a subsequence {𝐮n}n\{\mathbf{u}_{n^{\prime}}\}_{n^{\prime}} of {𝐮n}n\{\mathbf{u}_{n}\}_{n} and 𝐮𝒞(J,Hw)\mathbf{u}\in\mathcal{C}(J,H_{\textrm{\rm w}}) such that 𝐮n𝐮\mathbf{u}_{n^{\prime}}\to\mathbf{u} in 𝒞(J,Hw)\mathcal{C}(J,H_{\textrm{\rm w}}). It only remains to show that 𝐮𝒴J(R)\mathbf{u}\in\mathcal{Y}_{J}(R). The inequality (4.44) follows immediately from the weak convergence 𝐮n(t)𝐮(t)\mathbf{u}_{n^{\prime}}(t)\to\mathbf{u}(t) in HwH_{\textrm{\rm w}}. Moreover, denoting the right-hand side of (4.3) by K(s,t)K(s,t), it follows that for every 𝐰Wσ1,\mathbf{w}\in W_{\sigma}^{1,\infty} such that 𝐰Wσ1,1\|\mathbf{w}\|_{W_{\sigma}^{1,\infty}}\leq 1 and for all s,tJs,t\in J with s<ts<t, we have

|𝐮(t)𝐮(s),𝐰(Wσ1,),Wσ1,|=|(𝐮(t)𝐮(s),𝐰)|=limn|(𝐮n(t)𝐮n(s),𝐰)|=limn|𝐮n(t)𝐮n(s),𝐰(Wσ1,),Wσ1,|K(s,t).|\langle\mathbf{u}(t)-\mathbf{u}(s),\mathbf{w}\rangle_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}|=|(\mathbf{u}(t)-\mathbf{u}(s),\mathbf{w})|\\ =\lim_{n^{\prime}\to\infty}|(\mathbf{u}_{n^{\prime}}(t)-\mathbf{u}_{n^{\prime}}(s),\mathbf{w})|=\lim_{n^{\prime}\to\infty}|\langle\mathbf{u}_{n^{\prime}}(t)-\mathbf{u}_{n^{\prime}}(s),\mathbf{w}\rangle_{(W_{\sigma}^{1,\infty})^{\prime},W_{\sigma}^{1,\infty}}|\leq K(s,t).

This implies that

𝐮(t)𝐮(s)(Wσ1,)K(s,t),\displaystyle\|\mathbf{u}(t)-\mathbf{u}(s)\|_{(W_{\sigma}^{1,\infty})^{\prime}}\leq K(s,t),

and hence 𝐮\mathbf{u} satisfies (4.3). Consequently, 𝐮𝒴J(R)\mathbf{u}\in\mathcal{Y}_{J}(R), as desired.

The second part of the statement follows from the characterization in (4.47), and the fact that each 𝒴Jn(R)\mathcal{Y}_{J_{n}}(R) is a compact set in 𝒞(Jn,Hw)\mathcal{C}(J_{n},H_{\textrm{\rm w}}). By a diagonalization argument, we deduce that 𝒴I(R)\mathcal{Y}_{I}(R) is compact in 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}). This concludes the proof. ∎

We now obtain the following result as an application of Theorem 3.1. It shows convergence of a vanishing viscosity net of 𝒰Iν\mathcal{U}_{I}^{\nu}-trajectory statistical solutions of the 3D NSE towards a 𝒰Idiss\mathcal{U}^{\textrm{\rm diss}}_{I}-trajectory statistical solution of the 3D Euler equations.

Theorem 4.2.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and assume that 𝐟Lloc2(I,H).\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Consider 𝒰Iν,𝒰Idiss𝒞loc(I,Hw)\mathcal{U}^{\nu}_{I},\mathcal{U}^{\textrm{\rm diss}}_{I}\subset\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}), ν>0\nu>0, as defined in (4.40) and (4.43), respectively. Also, let μ0\mu_{0} be a Borel probability measure on HH, and, for each ν>0\nu>0, let ρν\rho_{\nu} be a 𝒰Iν\mathcal{U}^{\nu}_{I}-trajectory statistical solution of the 3D ν\nu-NSE such that Πt0ρν=μ0\Pi_{t_{0}}\rho_{\nu}=\mu_{0}. Then, there exists a subnet of {ρν}ν>0\{\rho_{\nu}\}_{\nu>0} that converges as ν0\nu\to 0, with respect to the weak-star semicontinuity topology, to a 𝒰Idiss\mathcal{U}^{\textrm{\rm diss}}_{I}-trajectory statistical solution ρ\rho of the 3D Euler equations satisfying Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

Proof.

As in the beginning of this subsection, let us denote X=HwX=H_{\textrm{\rm w}} and 𝒳=𝒞loc(I,Hw)\mathcal{X}=\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}). We proceed by constructing a sequence of compact sets {𝒦n}n\{\mathcal{K}_{n}\}_{n\in\mathbb{N}} in 𝒳\mathcal{X} which satisfies the assumptions of Theorem 3.1. First, since HH is a separable Banach space then, as recalled in Section 2.2, it follows that μ0\mu_{0} is tight. Moreover, the Borel σ\sigma-algebras in HH and HwH_{\textrm{\rm w}} coincide. Thus, given any sequence of positive real numbers {δn}n\{\delta_{n}\}_{n\in\mathbb{N}} with δn0\delta_{n}\to 0, there exists a corresponding sequence of (strongly) compact sets {Kn}n\{K_{n}\}_{n\in\mathbb{N}} in HH such that

μ0(X\Kn)<δn,for all n.\displaystyle\mu_{0}(X\backslash K_{n})<\delta_{n},\quad\mbox{for all }n. (4.50)

For each nn, let Rn>0R_{n}>0 such that KnBH(Rn)K_{n}\subset B_{H}(R_{n}), and define

𝒦n=𝒴I(Rn)Πt01Kn,\displaystyle\mathcal{K}_{n}=\mathcal{Y}_{I}(R_{n})\cap\Pi_{t_{0}}^{-1}K_{n},

with 𝒴I(Rn)\mathcal{Y}_{I}(R_{n}) as defined in (4.46), for fixed ν0>0\nu_{0}>0. From Lemma 4.2, 𝒴I(Rn)\mathcal{Y}_{I}(R_{n}) is a compact set in 𝒳\mathcal{X}. Moreover, since KnK_{n} is compact in HH, hence also compact (and closed) in X=HwX=H_{\textrm{\rm w}}, and Πt0:𝒳X\Pi_{t_{0}}:\mathcal{X}\to X is a continuous operator, then Πt01Kn\Pi_{t_{0}}^{-1}K_{n} is a closed set in 𝒳\mathcal{X}. This implies that each 𝒦n\mathcal{K}_{n} is a compact set in 𝒳\mathcal{X}.

Let us verify that the sequence {𝒦n}n\{\mathcal{K}_{n}\}_{n} satisfies condition (A1) of Theorem 3.1. From Definition 3.1, for each ν>0\nu>0 there exists a Borel set 𝒱ν\mathcal{V}_{\nu} in 𝒳\mathcal{X} such that 𝒱ν𝒰Iν\mathcal{V}_{\nu}\subset\mathcal{U}_{I}^{\nu} and ρν(𝒳\𝒱ν)=0\rho_{\nu}(\mathcal{X}\backslash\mathcal{V}_{\nu})=0111In fact, since, as shown in [40, Proposition 2.12], 𝒰Iν\mathcal{U}_{I}^{\nu} is itself a Borel set in 𝒳\mathcal{X}, then we could take 𝒱ν=𝒰Iν\mathcal{V}_{\nu}=\mathcal{U}_{I}^{\nu}.. Moreover, from 4.3 and the definition of 𝒴I(Rn)\mathcal{Y}_{I}(R_{n}) in (4.46) and the fact that 𝒰Iν𝒴I(Rn)\mathcal{U}_{I}^{\nu}\subset\mathcal{Y}_{I}(R_{n}) when νν0\nu\leq\nu_{0}, it follows that

𝒱νΠt01Kn𝒰IνΠt01Kn𝒴I(Rn)Πt01Kn=𝒦n,for all 0<νν0.\displaystyle\mathcal{V}_{\nu}\cap\Pi_{t_{0}}^{-1}K_{n}\subset\mathcal{U}_{I}^{\nu}\cap\Pi_{t_{0}}^{-1}K_{n}\subset\mathcal{Y}_{I}(R_{n})\cap\Pi_{t_{0}}^{-1}K_{n}=\mathcal{K}_{n},\quad\mbox{for all }0<\nu\leq\nu_{0}.

With these two facts, we obtain that, for all 0<νν0,0<\nu\leq\nu_{0},

ρν(𝒳\𝒦n)\displaystyle\rho_{\nu}(\mathcal{X}\backslash\mathcal{K}_{n}) ρν(𝒳\(𝒱νΠt01Kn))\displaystyle\leq\rho_{\nu}(\mathcal{X}\backslash(\mathcal{V}_{\nu}\cap\Pi_{t_{0}}^{-1}K_{n}))
=ρν(𝒳\Πt01Kn)=ρν(Πt01(X\Kn))=Πt0ρν(X\Kn)=μ0(X\Kn).\displaystyle\quad=\rho_{\nu}(\mathcal{X}\backslash\Pi_{t_{0}}^{-1}K_{n})=\rho_{\nu}(\Pi_{t_{0}}^{-1}(X\backslash K_{n}))=\Pi_{t_{0}}\rho_{\nu}(X\backslash K_{n})=\mu_{0}(X\backslash K_{n}).

Thus, together with (4.50), we deduce that

ρν(𝒳\𝒦n)<δn,for all n and  0<νν0,\displaystyle\rho_{\nu}(\mathcal{X}\backslash\mathcal{K}_{n})<\delta_{n},\quad\mbox{for all }n\,\mbox{ and }\,0<\nu\leq\nu_{0},

as desired.

Regarding condition (A2) of Theorem 3.1, first note that

𝒱ν𝒦n𝒰Iν𝒦n𝒞loc(I,BH(Rn)w)\displaystyle\mathcal{V}_{\nu}\cap\mathcal{K}_{n}\subset\mathcal{U}_{I}^{\nu}\cap\mathcal{K}_{n}\subset\mathcal{C}_{\textrm{\rm loc}}(I,B_{H}(R_{n})_{\textrm{\rm w}})

for all ν>0\nu>0 and nn\in\mathbb{N}. Thus,

lim supν0(𝒱ν𝒦n)𝒞loc(I,BH(Rn)w),for all n.\displaystyle\limsup_{\nu\to 0}(\mathcal{V}_{\nu}\cap\mathcal{K}_{n})\subset\mathcal{C}_{\textrm{\rm loc}}(I,B_{H}(R_{n})_{\textrm{\rm w}}),\quad\mbox{for all }n.

Since 𝒞loc(I,BH(Rn)w)\mathcal{C}_{\textrm{\rm loc}}(I,B_{H}(R_{n})_{\textrm{\rm w}}) is metrizable, given 𝐮lim supν0(𝒰Iν𝒦n)\mathbf{u}\in\limsup_{\nu\to 0}(\mathcal{U}_{I}^{\nu}\cap\mathcal{K}_{n}) there exists a sequence {𝐮νj}j\{\mathbf{u}_{\nu_{j}}\}_{j\in\mathbb{N}} such that 𝐮νj𝒱ν𝒦n𝒰IνΠt01Kn\mathbf{u}_{\nu_{j}}\in\mathcal{V}_{\nu}\cap\mathcal{K}_{n}\subset\mathcal{U}_{I}^{\nu}\cap\Pi_{t_{0}}^{-1}K_{n} for all jj\in\mathbb{N}, with νj0\nu_{j}\to 0 and 𝐮νj𝐮\mathbf{u}_{\nu_{j}}\to\mathbf{u} in 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) as jj\to\infty. By 4.4, this implies that 𝐮𝒰Idiss\mathbf{u}\in\mathcal{U}^{\textrm{\rm diss}}_{I}. Hence, condition (A2) of Theorem 3.1 is satisfied. The conclusion now follows as an application of Theorem 3.1. ∎

Given any Borel probability measure μ0\mu_{0} on HH and ν>0\nu>0, existence of a 𝒰Iν\mathcal{U}_{I}^{\nu}-trajectory statistical solution ρν\rho_{\nu} of the 3D NSE in the sense of Definition 3.1 satisfying the initial condition Πt0ρν=μ0\Pi_{t_{0}}\rho_{\nu}=\mu_{0} is shown in [13, Theorem 4.2] (see also [40]). This fact together with Theorem 4.2 thus yields the following corollary on the existence of 𝒰Idiss\mathcal{U}^{\textrm{\rm diss}}_{I}-trajectory statistical solutions of the 3D Euler equations satisfying a given initial datum.

Corollary 4.2.

Let II\subset\mathbb{R} be an interval closed and bounded on the left with left endpoint t0t_{0} and assume that 𝐟Lloc2(I,H).\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Let 𝒰Idiss𝒞loc(I,Hw)\mathcal{U}^{\textrm{\rm diss}}_{I}\subset\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) be as defined in (4.43). Then, given a Borel probability measure μ0\mu_{0} on HH, there exists a 𝒰Idiss\mathcal{U}^{\textrm{\rm diss}}_{I}-trajectory statistical solution ρ\rho of the 3D Euler equations satisfying the initial condition Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

4.4. Convergence of statistical solutions of the Galerkin approximations of the 3D NSE

We now address the Galerkin approximation of the three-dimensional Navier-Stokes equations (4.1) on a periodic spatial domain Ω\Omega, on a time interval II\subset\mathbb{R} closed and bounded on the left, with left endpoint t0t_{0}\in\mathbb{R}, and with 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Our aim is to apply Theorem 3.2 to show that trajectory statistical solutions generated by the well-defined solution operator of the Galerkin approximations converge to a trajectory statistical solution of the 3D NSE. For the framework of Section 4.1.2, the phase space is taken to be X=Hw,X=H_{\textrm{\rm w}}, so the trajectory space is 𝒳=𝒞loc(I,Hw)\mathcal{X}=\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}). The set 𝒰Iν\mathcal{U}^{\nu}_{I} is that of the weak solutions of the 3D ν\nu-NSE on II defined in (4.40).

For each mm\in\mathbb{N}, let PmP_{m} denote the projection of HH onto the finite-dimensional subspace of HH spanned by the first mm eigenfunctions 𝐰1,,𝐰m\mathbf{w}_{1},\ldots,\mathbf{w}_{m} of the Stokes operator which, on the periodic case, coincides with the negative Laplacian operator Δ-\Delta on VH2(Ω)3V\cap H^{2}(\Omega)^{3}. The Galerkin approximation in PmHP_{m}H of the 3D NSE (4.1) is defined as

t𝐮mνΔ𝐮m+Pm[(𝐮m)𝐮m]+pm=Pm𝐟,𝐮m=0,\displaystyle\partial_{t}\mathbf{u}_{m}-\nu\Delta\mathbf{u}_{m}+P_{m}[(\mathbf{u}_{m}\cdot\nabla)\mathbf{u}_{m}]+\nabla p_{m}=P_{m}\mathbf{f},\quad\nabla\cdot\mathbf{u}_{m}=0, (4.51)

see e.g. [61] for more details.

By taking the inner product of the first equation in (4.51) with each of the eigenfunctions 𝐰1,,𝐰m\mathbf{w}_{1},\ldots,\mathbf{w}_{m} and writing 𝐮m(t)=i=1mαim(t)𝐰i\mathbf{u}_{m}(t)=\sum_{i=1}^{m}\alpha_{im}(t)\mathbf{w}_{i}, it follows that the Galerkin approximation is equivalent to a system of mm ordinary differential equations on II of the form 𝜶t=𝐅(t,𝜶)\boldsymbol{\alpha}_{t}=\mathbf{F}(t,\boldsymbol{\alpha}), where 𝜶m=(α1m,,αmm)\boldsymbol{\alpha}_{m}=(\alpha_{1m},\ldots,\alpha_{mm}). The right-hand side 𝐅\mathbf{F} is quadratic (hence locally Lipschitz) in 𝜶\boldsymbol{\alpha} and, due to 𝐟Lloc2(I,H)\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H), 𝐅\mathbf{F} is also measurable in tt and bounded by an integrable function of tt on every compact subset of I×mI\times\mathbb{R}^{m}. As such, we obtain, from the classical Carathéodory theory of existence and uniqueness of solutions for ODEs [20, 45], an absolutely continuous function 𝐮m\mathbf{u}_{m} on [t0,tm][t_{0},t_{m}], for some tmIt_{m}\in I, which satisfies (4.51) a.e. in tt and a given initial condition 𝐮m(t0)=𝐮0,mPmH\mathbf{u}_{m}(t_{0})=\mathbf{u}_{0,m}\in P_{m}H. Moreover, from standard energy estimates, it follows that 𝐮m\mathbf{u}_{m} satisfies the following inequality for any compact subinterval JIJ\subset I with left endpoint t0t_{0} and containing [t0,tm][t_{0},t_{m}]:

|𝐮m(t)|2|𝐮0,m|2eνλ1(tt0)+1νλ1𝐟L2(J,H)2\displaystyle|\mathbf{u}_{m}(t)|^{2}\leq|\mathbf{u}_{0,m}|^{2}e^{-\nu\lambda_{1}(t-t_{0})}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(J,H)}

for every t[t0,tm]t\in[t_{0},t_{m}]. This implies that 𝐮m\mathbf{u}_{m} is uniformly bounded in tt, and consequently 𝐮m,\mathbf{u}_{m}, in fact, exists and is unique for all tIt\in I, with 𝐮m𝒞loc(I,PmH)\mathbf{u}_{m}\in\mathcal{C}_{\textrm{\rm loc}}(I,P_{m}H).

Therefore, we can define a solution operator associated to (4.51), given as

Sm:Hw\displaystyle S_{m}:H_{\textrm{\rm w}} \displaystyle\rightarrow 𝒞loc(I,PmH)𝒞loc(I,Hw)\displaystyle\mathcal{C}_{\textrm{\rm loc}}(I,P_{m}H)\subset\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}})
𝐮0\displaystyle\mathbf{u}_{0} \displaystyle\mapsto 𝐮m,\displaystyle\mathbf{u}_{m}, (4.52)

where 𝐮m=𝐮m(t)\mathbf{u}_{m}=\mathbf{u}_{m}(t) is the unique trajectory solving (4.51) on II subject to the initial condition 𝐮m(t0)=Pm𝐮0\mathbf{u}_{m}(t_{0})=P_{m}\mathbf{u}_{0}.

For the statistical solutions, we consider a sequence of initial measures {μ0,m}m\{\mu_{0,m}\}_{m\in\mathbb{N}} on HwH_{\textrm{\rm w}} associated with the Galerkin approximations. Since the initial conditions associated with the Galerkin operator SmS_{m} belong to PmH,P_{m}H, it is natural to assume that each μ0,m\mu_{0,m} is a Borel probability measure on HwH_{\textrm{\rm w}} (or, equivalently, on HH) which is carried by PmH.P_{m}H. For the sake of convergence, we also assume that these measures converge to a Borel probability measure on H,H, in the sense of weak-star semicontinuity topology.

With this setting, we have the following convergence result.

Theorem 4.3.

Let ν>0\nu>0 and let II\subset\mathbb{R} be an interval that is closed and bounded on the left with left endpoint t0t_{0}. Assume that 𝐟Lloc2(I,H).\mathbf{f}\in L^{2}_{\textrm{\rm loc}}(I,H). Let Sm:Hw𝒞loc(I,PmH)S_{m}:H_{\textrm{\rm w}}\to\mathcal{C}_{\textrm{\rm loc}}(I,P_{m}H), mm\in\mathbb{N}, and 𝒰Iν𝒞loc(I,Hw)\mathcal{U}^{\nu}_{I}\subset\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) be defined as in (4.4) and (4.40), respectively. Let μ0\mu_{0} be a (tight) Borel probability measure on Hw.H_{\textrm{\rm w}}. Suppose {μ0,m}m\{\mu_{0,m}\}_{m\in\mathbb{N}} is a sequence of (tight) Borel probability measures on HwH_{\textrm{\rm w}} carried by PmHP_{m}H which is uniformly tight on HwH_{\textrm{\rm w}} and converges to μ0\mu_{0} in the sense of weak-star semicontinuity topology, i.e. μ0,mwscμ0\mu_{0,m}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} in Hw.H_{\textrm{\rm w}}. Then, the sequence of measures {Smμ0,m}m\{S_{m}\mu_{0,m}\}_{m\in\mathbb{N}} has a subsequence that converges, as mm\to\infty, with respect to the weak-star semicontinuity topology, to a 𝒰Iν\mathcal{U}^{\nu}_{I}-trajectory statistical solution ρ\rho of the 3D Navier-Stokes equations satisfying Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}.

Proof.

We proceed to verify that assumptions (H1)-(H5) of Theorem 3.2 hold under this setting.

First, from the definition (4.4) of the Galerkin semigroup as the weak solution of the Galerkin approximation (4.51) with the initial condition 𝐮m(t0)=Pm𝐮0,\mathbf{u}_{m}(t_{0})=P_{m}\mathbf{u}_{0}, it follows that the operator Πt0Sm\Pi_{t_{0}}S_{m} considered in the statement of Theorem 3.2 is such that Πt0Sm𝐮0=𝐮m(t0)=Pm𝐮0,\Pi_{t_{0}}S_{m}\mathbf{u}_{0}=\mathbf{u}_{m}(t_{0})=P_{m}\mathbf{u}_{0}, so that this operator is precisely the Galerkin projector, i.e. Πt0Sm=Pm.\Pi_{t_{0}}S_{m}=P_{m}. Thus, condition (H1) reads Pmμ0,mwscμ0.P_{m}\mu_{0,m}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}. Since μ0,m\mu_{0,m} is assumed to be carried by PmH,P_{m}H, we thus have Pmμ0,m=μ0,m.P_{m}\mu_{0,m}=\mu_{0,m}. Indeed, for any Borel set AA in HH, we have

Pmμ0,m(A)=μ0,m(Pm1A)=μ0,m(Pm1APmH)=μ0,m(APmH)=μ0,m(A),\displaystyle P_{m}\mu_{0,m}(A)=\mu_{0,m}(P_{m}^{-1}A)=\mu_{0,m}(P_{m}^{-1}A\cap P_{m}H)=\mu_{0,m}(A\cap P_{m}H)=\mu_{0,m}(A),

where, in the second and fourth equalities, we used the fact that μ0,m\mu_{0,m} is carried by PmH,P_{m}H, while, in the third equality, we used that PmP_{m} is a projection operator, so that Pm1APmH=APmH.P_{m}^{-1}A\cap P_{m}H=A\cap P_{m}H.

Thus, since Pmμ0,m=μ0,m,P_{m}\mu_{0,m}=\mu_{0,m}, condition (H1) is precisely the assumption that we have, i.e. that μ0,mwscμ0.\mu_{0,m}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0}. Hence, condition (H1) is satisfied.

From the equivalence between (4.51) and the system of mm ordinary differential equations 𝜶t=𝐅(t,𝜶)\boldsymbol{\alpha}_{t}=\mathbf{F}(t,\boldsymbol{\alpha}), together with the properties of 𝐅\mathbf{F} recalled above, it follows again from classical ODE theory that any solution 𝐮m\mathbf{u}_{m} of (4.51) depends continuously on initial data. This implies that the solution operator Sm:Hw𝒞loc(I,PmH)𝒞loc(I,Hw)S_{m}:H_{\textrm{\rm w}}\to\mathcal{C}_{\textrm{\rm loc}}(I,P_{m}H)\subset\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) is continuous, and hence assumption (H2) is verified.

The validity of assumption (H3) follows immediately from the condition that the sequence of initial measures {μ0,m}m\{\mu_{0,m}\}_{m} is uniformly tight in HwH_{\textrm{\rm w}}. Indeed, given any sequence δn0,\delta_{n}\rightarrow 0, there is a corresponding sequence of compact sets KnK_{n} in HwH_{\textrm{\rm w}}, nn\in\mathbb{N}, such that (H3) holds. Since, as we show next, the remaining assumptions hold for any compact set KK in HwH_{\textrm{\rm w}}, they hold in particular for such sequence.

To establish (H4) and (H5), we first define an auxiliary space analogous to (4.46). Since, in the current setting, ν\nu is a fixed parameter, we may invoke a different set of inequalities than (4.44)-(4.3) to define this auxiliary space, which yield estimates in stronger norms. These alternative inequalities are indeed necessary for guaranteeing compactness of such auxiliary space in the topology of Lloc2(I,H)L^{2}_{\textrm{\rm loc}}(I,H). This in turn allows us to obtain a result analogous to 4.4, showing that individual solutions of the Galerkin approximations converge to a Leray-Hopf weak solution of 3D NSE.

More precisely, under the present framework, we have that for any mm\in\mathbb{N} and any solution 𝐮m\mathbf{u}_{m} of (4.51) on II, the following inequalities hold:

|𝐮m(t)|2+νt0t𝐮m(τ)2dτ|𝐮m(t0)|2+1νλ1𝐟L2(t0,t;H)2,\displaystyle|\mathbf{u}_{m}(t)|^{2}+\nu\int_{t_{0}}^{t}\|\mathbf{u}_{m}(\tau)\|^{2}{\text{\rm d}}\tau\leq|\mathbf{u}_{m}(t_{0})|^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}, (4.53)

for all tIt\in I, and

𝐮m(t)𝐮m(s)Vcν1/2|ts|1/2(|𝐮m(t0)|2+1νλ1𝐟L2(t0,t;H)2)1/2+cν3/4|ts|1/4(|𝐮m(t0)|2+1νλ1𝐟L2(t0,t;H)2),\|\mathbf{u}_{m}(t)-\mathbf{u}_{m}(s)\|_{V^{\prime}}\leq c\nu^{1/2}|t-s|^{1/2}\left(|\mathbf{u}_{m}(t_{0})|^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}\right)^{1/2}\\ +c\nu^{-3/4}|t-s|^{1/4}\left(|\mathbf{u}_{m}(t_{0})|^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}\right), (4.54)

for all s,tIs,t\in I with sts\leq t, and for some positive constant cc which is independent of mm.

The proof of (4.53) follows from typical energy estimates, see e.g. [61, Chapter 3]. Inequality (4.54) then follows by proceeding similarly as in the proof of (4.19) in 4.3 and invoking (4.53). We omit further details.

Now, given an arbitrary R>0R>0 and an arbitrary subinterval JIJ\subseteq I that is closed and bounded on the left with left endpoint t0t_{0}, consider the following inequalities for 𝐮𝒞loc(J,Hw)\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(J,H_{\textrm{\rm w}}):

|𝐮(t)|2+νt0t𝐮(τ)2dτR2+1νλ1𝐟L2(t0,t;H)2,\displaystyle|\mathbf{u}(t)|^{2}+\nu\int_{t_{0}}^{t}\|\mathbf{u}(\tau)\|^{2}{\text{\rm d}}\tau\leq R^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}, (4.55)

for tJt\in J, and

𝐮(t)𝐮(s)Vcν1/2|ts|1/2(R2+1νλ1𝐟L2(t0,t;H)2)1/2+cν3/4|ts|1/4(R2+1νλ1𝐟L2(t0,t;H)2),\|\mathbf{u}(t)-\mathbf{u}(s)\|_{V^{\prime}}\leq c\nu^{1/2}|t-s|^{1/2}\left(R^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}\right)^{1/2}\\ +c\nu^{-3/4}|t-s|^{1/4}\left(R^{2}+\frac{1}{\nu\lambda_{1}}\|\mathbf{f}\|^{2}_{L^{2}(t_{0},t;H)}\right), (4.56)

for s,tJs,t\in J with s<ts<t, where cc is the same constant from (4.54). Based on these, we define the set

𝒴J(R)={𝐮𝒞loc(J,Hw):𝐮 satisfies (4.55) for all tJ and (4.56) for all s<t in J}.\displaystyle\mathcal{Y}_{J}(R)=\left\{\mathbf{u}\in\mathcal{C}_{\textrm{\rm loc}}(J,H_{\textrm{\rm w}}):\mathbf{u}\mbox{ satisfies }\eqref{ineq:YI:Gal:1}\mbox{ for all $t\in J$ and }\eqref{ineq:YI:Gal:2}\mbox{ for all $s<t$ in $J$}\right\}. (4.57)

Then, the same characterization as in (4.47) holds for 𝒴I(R)\mathcal{Y}_{I}(R) in this case, and by analogous arguments as in Lemma 4.2 we deduce that 𝒴I(R)\mathcal{Y}_{I}(R) is a compact subset of 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}). Moreover, analogously as in 4.6, we can invoke the inequalities (4.53) and (4.54) to show that for any compact set KK in HwH_{\textrm{\rm w}} and any R>0R>0 such that KBH(R)K\subset B_{H}(R), it holds that Sm(K)𝒴I(R)S_{m}(K)\subset\mathcal{Y}_{I}(R) for all mm\in\mathbb{N}. This shows that assumption (H4) is satisfied.

Finally, to verify assumption (H5), we argue similarly to the the proof of 4.7. Specifically, given a compact set KK in HwH_{\textrm{\rm w}}, let R>0R>0 such that KBH(R)K\subset B_{H}(R). As we showed in the verification of (H4), this implies that Sm(K)𝒴I(R)S_{m}(K)\subset\mathcal{Y}_{I}(R) for all mm, and hence lim supmSm(K)𝒴I(R)\limsup_{m}S_{m}(K)\subset\mathcal{Y}_{I}(R). Since 𝒴I(R)\mathcal{Y}_{I}(R) is metrizable, given 𝐮lim supmSm(K)\mathbf{u}\in\limsup_{m}S_{m}(K) there exists a sequence {𝐮mj}j\{\mathbf{u}_{m_{j}}\}_{j\in\mathbb{N}} such that 𝐮mjSmj(K)\mathbf{u}_{m_{j}}\in S_{m_{j}}(K) for all jj, and 𝐮mj𝐮\mathbf{u}_{m_{j}}\to\mathbf{u} in 𝒴I(R)\mathcal{Y}_{I}(R) as jj\to\infty. Consequently, 𝐮mj𝐮\mathbf{u}_{m_{j}}\to\mathbf{u} in 𝒞loc(I,Hw)\mathcal{C}_{\textrm{\rm loc}}(I,H_{\textrm{\rm w}}) as jj\to\infty.

Moreover, from the fact that 𝐮mjSmj(K)𝒴I(R)\mathbf{u}_{m_{j}}\in S_{m_{j}}(K)\subset\mathcal{Y}_{I}(R) for all jj and from the definition of 𝒴I(R)\mathcal{Y}_{I}(R) in (4.57), it follows that the sequence {𝐮mj}j\{\mathbf{u}_{m_{j}}\}_{j} has the uniform upper bounds implied by (4.55) and (4.56) on every compact subinterval JIJ\subset I with left endpoint t0t_{0}. Then, standard compactness arguments yield that 𝐮Lloc2(I,V)\mathbf{u}\in L^{2}_{\textrm{\rm loc}}(I,V) and, modulo a subsequence, 𝐮mj𝐮\mathbf{u}_{m_{j}}\to\mathbf{u} in Lloc2(I,H)L^{2}_{\textrm{\rm loc}}(I,H). Combining all these facts, we may thus pass to the limit mm\to\infty in the weak formulation of (4.51) and deduce that 𝐮\mathbf{u} is a Leray-Hopf weak solution of the 3D NSE, as defined in Definition 4.3. See e.g. [61, Chapter 3, Section 3] for similar arguments. Therefore, 𝐮𝒰Iν\mathbf{u}\in\mathcal{U}^{\nu}_{I}, and we conclude that lim supmSm(K)𝒰Iν\limsup_{m}S_{m}(K)\subset\mathcal{U}^{\nu}_{I}. This shows that (H5) holds.

We have thus verified all the assumptions from Theorem 3.2, which then yields the desired result. ∎

Remark 4.1.

Note that if {μ0,m}m\{\mu_{0,m}\}_{m} is a sequence of Borel probability measures on HwH_{\textrm{\rm w}} which converges to a Borel probability measure μ0\mu_{0} with respect to the weak-star semicontinuity topology in HH (with the strong topology), then it also converges in the weak-star topology in HH, see Lemma 2.2. Since HH is a Polish space, then by Prohorov’s theorem [56] the relatively compact subsets of 𝒫(H)\mathcal{P}(H) coincide with the uniformly tight ones. In particular, it follows that {μ0,m}m\{\mu_{0,m}\}_{m} is uniformly tight in HH, and consequently in HwH_{\textrm{\rm w}}. Therefore, if μ0,mwscμ0\mu_{0,m}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} in HH for some μ0𝒫(H)\mu_{0}\in\mathcal{P}(H) then the condition from Theorem 4.3 that {μ0,m}m\{\mu_{0,m}\}_{m} is uniformly tight on HwH_{\textrm{\rm w}} is immediately satisfied.

Remark 4.2.

In the statement of Theorem 4.3, the conditions imposed on the initial approximating measures μ0,m\mu_{0,m} are, in a sense, generic. In practice, one would want to start with something more specific. For example, given an initial tight Borel probability measure μ0\mu_{0} of interest for the limit problem, we may consider the Galerkin projections μ0,m=Pmμ0\mu_{0,m}=P_{m}\mu_{0} of that measure. Note that Pmμ0P_{m}\mu_{0} is carried by PmHw.P_{m}H_{\textrm{\rm w}}. Let us verify that such approximating measures satisfy the remaining conditions of Theorem 4.3.

In order to see that {Pmμ0}m\{P_{m}\mu_{0}\}_{m} is uniformly tight in HwH_{\textrm{\rm w}}, fix δ>0\delta>0 and let KK be a compact set in HwH_{\textrm{\rm w}} such that μ0(H\K)<δ\mu_{0}(H\backslash K)<\delta. Let also R>0R>0 be such that KBH(R)K\subset B_{H}(R). The subset BH(R)B_{H}(R) is a compact set in HwH_{\textrm{\rm w}}, and

Pmμ0(H\BH(R))=μ0(Pm1(H\BH(R)))=μ0(H\Pm1BH(R))μ0(H\BH(R))\displaystyle P_{m}\mu_{0}(H\backslash B_{H}(R))=\mu_{0}(P_{m}^{-1}(H\backslash B_{H}(R)))=\mu_{0}(H\backslash P_{m}^{-1}B_{H}(R))\leq\mu_{0}(H\backslash B_{H}(R))
μ0(H\K)<δ\displaystyle\leq\mu_{0}(H\backslash K)<\delta

for all mm, where in the first inequality we used that PmBH(R)BH(R)P_{m}B_{H}(R)\subset B_{H}(R). This shows that {Pmμ0}m\{P_{m}\mu_{0}\}_{m} is uniformly tight in HwH_{\textrm{\rm w}}.

The second condition is that Pmμ0wscμ0P_{m}\mu_{0}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} in Hw.H_{\textrm{\rm w}}. This can be seen by noting that Pm𝐮0𝐮0P_{m}\mathbf{u}_{0}\to\mathbf{u}_{0} in H,H, for all 𝐮0H\mathbf{u}_{0}\in H, and hence, by the Dominated Convergence Theorem, we have that for every bounded and continuous real-valued function φ\varphi on HH

Hφ(𝐮)d(Pmμ0)(𝐮)=Hφ(Pm𝐮)dμ0(𝐮)Hφ(𝐮)dμ0(𝐮),\int_{H}\varphi(\mathbf{u}){\text{\rm d}}(P_{m}\mu_{0})(\mathbf{u})=\int_{H}\varphi(P_{m}\mathbf{u}){\text{\rm d}}\mu_{0}(\mathbf{u})\to\int_{H}\varphi(\mathbf{u}){\text{\rm d}}\mu_{0}(\mathbf{u}),

so that Pmμ0wμ0P_{m}\mu_{0}\stackrel{{\scriptstyle w*}}{{\rightharpoonup}}\mu_{0} in HH. Since HH is completely regular, it follows from Lemma 2.2 that Pmμ0wscμ0P_{m}\mu_{0}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} in H.H. Moreover, since any open set in HwH_{\textrm{\rm w}} is open in H,H, we obtain from the equivalence between conditions (i) and (v) in Lemma 2.2 that Pmμ0wscμ0P_{m}\mu_{0}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} in Hw.H_{\textrm{\rm w}}.

Remark 4.3.

Another useful practical example is with a Monte-Carlo approximation μ0N=(1/N)k=1Nδ𝐮k\mu_{0}^{N}=(1/N)\sum_{k=1}^{N}\delta_{\mathbf{u}_{k}}, 𝐮kH,\mathbf{u}_{k}\in H, of a desired initial (tight) Borel probability measure μ0\mu_{0} on Hw.H_{\textrm{\rm w}}. The convergence Pmμ0,Nmwscμ0P_{m}\mu_{0,N_{m}}\stackrel{{\scriptstyle wsc*}}{{\rightharpoonup}}\mu_{0} for a suitable subsequence {Nm}m\{N_{m}\}_{m} is a delicate issue, though, but it can be proved in some cases. For a related result for the two-dimensional Navier-Stokes equations and a Gaussian initial measure with the eigenvalues of the covariance operator decaying sufficiently fast, see [4]. This will be further discussed in subsequent works.

Remark 4.4.

As a byproduct of Theorem 4.3, we obtain, for any given initial measure μ0\mu_{0}, the existence of a 𝒰Iν\mathcal{U}^{\nu}_{I}-trajectory statistical solution ρ\rho of the 3D Navier-Stokes equations satisfying the initial condition Πt0ρ=μ0\Pi_{t_{0}}\rho=\mu_{0}. Theorem 4.3 thus provides an alternative proof of this fact to the one previously given in [13, Theorem 4.2] (see also [40]), where existence was shown via an approximation by convex combinations of Dirac measures, by invoking the Krein-Milman theorem together with a tightness argument. Here, existence is derived instead via convergence of standard Galerkin approximations.

Acknowledgements

ACB received support under the grants #2019/02512-5, São Paulo Research Foundation (FAPESP), #312119/2016-0, Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and FAEPEX/
UNICAMP. CFM was supported by the grants NSF-DMS 2009859 and NSF-DMS 2239325. RMSR received support under the grants Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), Brasil, #001, and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Brasília, Brasil, #408751/2023-1.

References

  • [1] D. Albritton, E. Brué, and M. Colombo. Non-uniqueness of Leray solutions of the forced Navier-Stokes equations. Annals of Mathematics, 196(1):415–455, 2022.
  • [2] C.D. Aliprantis and K.C. Border. Infinite Dimensional Analysis: A Hitchhiker’s Guide. Springer-Verlag, 3rd edition, 2006.
  • [3] A. Barbaszewska-Wisniowska. On statistical solutions of the Navier-Stokes type equation. Univ. Iagel. Acta Math, (32):291–304, 1995.
  • [4] A. Barth, C. Schwab, and J. Sukys. Multilevel monte carlo simulation of statistical solutions to the navier-stokes equations. In R. Cools and D. Nuyens, editors, Monte Carlo and Quasi-Monte Carlo Methods, MCQMC 2014, Leuven, Belgium, April 2014, volume 163 of Springer Proceedings in Mathematics and Statistics, pages 209–227. Springer, 2014.
  • [5] A. Basson. Homogeneous statistical solutions and local energy inequality for 3D Navier-Stokes equations. Communications in mathematical physics, 266:17–35, 2006.
  • [6] G. Beer. Topologies on closed and closed convex sets, volume 268. Springer Science & Business Media, 1993.
  • [7] H. Bercovici, P. Constantin, C. Foias, and O.P. Manley. Exponential decay of the power spectrum of turbulence. Journal of Statistical Physics, 80:579–602, 1995.
  • [8] A. Biswas, C. Foias, C.F. Mondaini, and E.S. Titi. Downscaling data assimilation algorithm with applications to statistical solutions of the Navier–Stokes equations. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 36(2):295–326, 2019.
  • [9] V.I. Bogachev. Measure theory, volume 2. Springer, 2007.
  • [10] D. Breit, E. Feireisl, and M. Hofmanová. Markov selection for the stochastic compressible Navier–Stokes system. Annals of Applied Probability, 30(6):2547–2572, 2020.
  • [11] A.C. Bronzi, C.F. Mondaini, and R.M.S. Rosa. On the convergence of statistical solutions of the 3D Navier-Stokes-α\alpha model as α\alpha vanishes. Discrete and Continuous Dynamical Systems, 34(1):19–49, 2014.
  • [12] A.C. Bronzi, C.F. Mondaini, and R.M.S. Rosa. Trajectory Statistical Solutions for Three-Dimensional Navier–Stokes-Like Systems. SIAM Journal on Mathematical Analysis, 46(3):1893–1921, 2014.
  • [13] A.C. Bronzi, C.F. Mondaini, and R.M.S. Rosa. Abstract framework for the theory of statistical solutions. Journal of Differential Equations, 260(12):8428–8484, 2016.
  • [14] T. Buckmaster, M. Colombo, and V. Vicol. Wild solutions of the Navier–Stokes equations whose singular sets in time have Hausdorff dimension strictly less than 1. Journal of the European Mathematical Society, 24(9):3333–3378, 2021.
  • [15] T. Buckmaster and V. Vicol. Nonuniqueness of weak solutions to the Navier-Stokes equation. Annals of Mathematics, 189(1):101–144, 2019.
  • [16] R. Burachik and A. Iusem. Set-valued mappings and enlargements of monotone operators, volume 8 of Springer Optimization and Its Applications. Springer, 2008.
  • [17] J.E. Cardona and L. Kapitanski. Semiflow selection and Markov selection theorems. Topological Methods in Nonlinear Analysis, 56(1):197 – 227, 2020.
  • [18] D. Chae. The vanishing viscosity limit of statistical solutions of the Navier-Stokes equations. II. The general case. J. Math. Anal. Appl., 155(2):460–484, 1991.
  • [19] A. Cheskidov and X. Luo. Sharp nonuniqueness for the Navier–Stokes equations. Inventiones mathematicae, 229(3):987–1054, 2022.
  • [20] E.A. Coddington and N. Levinson. Theory of Ordinary Differential Equations. Tata McGraw-Hill Publishing Co. Ltd., 1987.
  • [21] P. Constantin and C. Foias. Navier-Stokes equations. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1988.
  • [22] P. Constantin, C. Foias, and O.P. Manley. Effects of the forcing function spectrum on the energy spectrum in 2-D turbulence. Physics of Fluids, 6(1):427–429, 1994.
  • [23] P. Constantin and F. Ramos. Inviscid limit for damped and driven incompressible Navier-Stokes equations in 2\mathbb{R}^{2}. Communications in Mathematical Physics, 275(2):529–551, 2007.
  • [24] C. De Lellis and L. Székelyhidi Jr. On admissibility criteria for weak solutions of the Euler equations. Archive for Rational Mechanics and Analysis, 195:225–260, 2010.
  • [25] R.J. DiPerna and A.J. Majda. Concentrations in regularizations for 2-D incompressible flow. Communications on Pure and Applied Mathematics, 40(3):301–345, 1987.
  • [26] C.R. Doering and E.S. Titi. Exponential decay rate of the power spectrum for solutions of the Navier–Stokes equations. Physics of Fluids, 7(6):1384–1390, 1995.
  • [27] R.E. Edwards. Functional analysis: theory and applications. Holt, Rinehart and Winston, Inc., New York, 1965.
  • [28] F. Fanelli and E. Feireisl. Statistical solutions to the barotropic Navier–Stokes system. Journal of Statistical Physics, 181:212–245, 2020.
  • [29] U.S. Fjordholm, S. Lanthaler, and S. Mishra. Statistical solutions of hyperbolic conservation laws: foundations. Archive for Rational Mechanics and Analysis, 226:809–849, 2017.
  • [30] U.S. Fjordholm, K. Lye, S. Mishra, and F. Weber. Statistical solutions of hyperbolic systems of conservation laws: numerical approximation. Mathematical Models and Methods in Applied Sciences, 30(3):539–609, 2020.
  • [31] U.S. Fjordholm, S. Mishra, and F. Weber. On the vanishing viscosity limit of statistical solutions of the incompressible Navier–Stokes equations. SIAM Journal on Mathematical Analysis, 56(4):5099–5143, 2024.
  • [32] F. Flandoli and M. Romito. Markov selections for the 3D stochastic Navier-Stokes equations. Probab. Theory Related Fields, 140:407–458, 2008.
  • [33] C. Foias. Statistical study of Navier-Stokes equations, I. Rendiconti del Seminario matematico della Università di Padova, 48:219–348, 1972.
  • [34] C. Foias. Statistical study of Navier-Stokes equations, II. Rendiconti del Seminario Matematico della Universita di Padova, 49:9–123, 1973.
  • [35] C. Foias. A functional approach to turbulence. Russian Mathematical Surveys, 29(2):293–326, 1974.
  • [36] C. Foias. What do the Navier-Stokes equations tell us about turbulence? Contemporary Mathematics, 208:151–180, 1997.
  • [37] C. Foias, O. Manley, R. Rosa, and R. Temam. Estimates for the energy cascade in three-dimensional turbulent flows. Comptes Rendus Acad. Sci. Paris, Série I, 333:499–504, 2001.
  • [38] C. Foias, O. Manley, R. Rosa, and R. Temam. Navier–Stokes Equations and Turbulence, volume 83 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 2001.
  • [39] C. Foias and G. Prodi. Sur les solutions statistiques des équations de Navier-Stokes. Ann. Mat. Pura Appl., 111(4):307–330, 1976.
  • [40] C. Foias, R. Rosa, and R. Temam. Properties of time-dependent statistical solutions of the three-dimensional Navier-Stokes equations. Annales de l’Institut Fourier, 63(6):2515–2573, 2013.
  • [41] G.B. Folland. Real analysis: modern techniques and their applications, volume 40 of Pure and Applied Mathematics. John Wiley & Sons, Inc., New York, 2nd edition, 1999.
  • [42] Z. Frolík. Concerning topological convergence of sets. Czechoslovak Mathematical Journal, 10(2):168–180, 1960.
  • [43] A.V. Fursikov. The closure problem for Friedman-Keller infinite chain of moment equations, corresponding to the Navier-Stokes system. In Fundamental problematic issues in turbulence, Trends in Mathematics, pages 17–24. Springer, 1999.
  • [44] J. Golec. On arsenev-type statistical solutions of differential equations with non-uniquely solvable cauchy problems. Univ. Iagel. Acta Math., 30:33–44, 1993.
  • [45] J.K. Hale. Ordinary Differential Equations. Robert E. Krieger Publishing Company, Inc., 1980.
  • [46] D. Hărăguş. Statistical solutions for an abstract equation of Navier-Stokes type. An. Univ. Timişoara Ser. Mat.-Inform., 37(1):73–84, 1999.
  • [47] R. Illner and J. Wick. Statistical solutions of differential equations with non-uniquely solvable Cauchy problems. Journal of Differential Equations, 41(3):289–300, 1981.
  • [48] J.L. Kelley. General topology, volume 27 of Graduate Texts in Mathematics. Springer-Verlag, 1975.
  • [49] J.L. Kelley and T.P. Srinivasan. Measure and integral, Vol. 1, volume 116 of Graduate Texts in Mathematics. Springer-Verlag, 1988.
  • [50] J.P. Kelliher. Infinite-energy 2D statistical solutions to the equations of incompressible fluids. Journal of Dynamics and Differential Equations, 21:631–661, 2009.
  • [51] O.A. Ladyzhenskaya. The mathematical theory of viscous incompressible flow. Revised English edition, Translated from the Russian by Richard A. Silverman Gordon and Breach Science Publishers, New York-London, 1963.
  • [52] O.A. Ladyzhenskaya and A.M. Vershik. Sur l’évolution des mesures déterminées par les équations de Navier-Stokes et la résolution du problème de Cauchy pour l’équation statistique de E. Hopf. Ann. Scuola Norm. Sup. Pisa Cl. Sci., 4(2):209–230, 1977.
  • [53] J.L. Lions. Quelques méthodes de résolution des problemes aux limites non linéaires. Dunod, Gauthier-Villars, Paris, 1969.
  • [54] P.L. Lions. Mathematical topics in fluid mechanics, Volume 1: Incompressible models. Number 3 in Oxford Lecture Series in Mathematics and its Applications. Clarendon Press, Oxford, 1996.
  • [55] C.F. Mondaini. Abstract framework for the theory of statistical solutions. PhD thesis, Federal University of Rio de Janeiro, Brazil, jun 2014.
  • [56] Y.V. Prokhorov. Convergence of random processes and limit theorems in probability theory. Theory of Probability & Its Applications, 1(2):157–214, 1956.
  • [57] F. Ramos, R. Rosa, and R. Temam. Statistical estimates for channel flows driven by a pressure gradient. Physica D: Nonlinear Phenomena, 237(10–12):1368–1387, 2008.
  • [58] F. Ramos and E.S. Titi. Invariant measures for the 33D Navier-Stokes-Voigt equations and their Navier-Stokes limit. Discrete and Continuous Dynamical Systems, 28(1):375–403, 2010.
  • [59] R. Rosa. Theory and applications of statistical solutions of the Navier–Stokes equations, volume 364 of London Mathematical Society Lecture Note Series, pages 228–257. Cambridge University Press, 2009.
  • [60] L. Sławik. Note on some variational problem related to statistical solutions of differential equations in Banach spaces. Annales Polonici Mathematici, 93(2):171–176, 2008.
  • [61] R. Temam. Navier–Stokes Equations and Nonlinear Functional Analysis, volume 66 of CBMS-NSF Regional Conference Series in Applied Mathematics. Society for Industrial and Applied Mathematics, Philadelphia, PA, 1995.
  • [62] R. Temam. Navier-Stokes equations: theory and numerical analysis. AMS Chelsea Publishing, Providence, Rhode Island, 2001.
  • [63] F. Topsøe. Topology and Measure, volume 133 of Lecture Notes in Mathematics. Springer-Verlag, 1970.
  • [64] M.I. Vishik and A.V. Fursikov. L’équation de Hopf, les solutions statistiques, les moments correspondant aux systémes des équations paraboliques quasilinéaires. J. Math. Pures Appl., 59(9):85–122, 1977.
  • [65] M.I. Vishik and A.V. Fursikov. Translationally homogeneous statistical solutions and individual solutions with infinite energy of a system of Navier–Stokes equations. Sibirskii Matematicheskii Zhurnal, 19(5):1005–1031, 1978.
  • [66] M.I. Vishik, E.S. Titi, and V.V. Chepyzhov. On convergence of trajectory attractors of the 3D Navier-Stokes-α\alpha model as α\alpha approaches 0. Sbornik: Mathematics, 198(12):1703–1736, 2007.
  • [67] R. Wagner and E. Wiedemann. Statistical solutions of the two-dimensional incompressible Euler equations in spaces of unbounded vorticity. Journal of Functional Analysis, 284(4):109777, 2023.
  • [68] X. Wang. Stationary statistical properties of Rayleigh-Bénard convection at large Prandtl number. Communications on Pure and Applied Mathematics, 61(6):789–815, 2008.
  • [69] E. Wiedemann. Existence of weak solutions for the incompressible Euler equations. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 28(5):727–730, 2011.
  • [70] V.I. Yudovich. Non-stationary flow of an ideal incompressible fluid. Zhurnal Vychislitel’noi Matematiki i Matematicheskoi Fiziki, 3(6):1032–1066, 1963.