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

On the vanishing viscosity limit of statistical solutions of the incompressible Navier–Stokes equations

Ulrik Skre Fjordholm
Department of Mathematics
University of Oslo
Postboks 1053 Blindern, 0316 Oslo, Norway
[email protected] https://www.mn.uio.no/math/english/people/aca/ulriksf/
Siddhartha Mishra
Seminar for Applied Mathematics (SAM)
ETH Zürich
HG G 57.2, Rämistrasse 101, Zürich, Switzerland.
[email protected] http://www.sam.math.ethz.ch/ smishra
 and  Franziska Weber
Department of Mathematical Sciences
Carnegie Mellon University
5000 Forbes Avenue, Pittsburgh, PA 15213, USA.
[email protected]
Abstract.

We study statistical solutions of the incompressible Navier–Stokes equation and their vanishing viscosity limit. We show that a formulation using correlation measures, which are probability measures accounting for spatial correlations, and moment equations is equivalent to statistical solutions in the Foiaş–Prodi sense. Under the assumption of weak scaling, a weaker version of Kolmogorov’s self-similarity at small scales hypothesis that allows for intermittency corrections, we show that the limit is a statistical solution of the incompressible Euler equations. To pass to the limit, we derive a Kármán–Howarth–Monin relation for statistical solutions and combine it with the weak scaling assumption and a compactness theorem for correlation measures.

1. Introduction

The motion of an incompressible viscous fluid can be described by the Navier–Stokes equations

(1.1) tu+div(uu)+p=εΔudivu=0u|t=0=u0,\begin{split}\partial_{t}u+\operatorname{div}\bigl{(}u\otimes u\bigr{)}+\nabla p&=\varepsilon\Delta u\\ \operatorname{div}u&=0\\ u\big{|}_{t=0}&=u_{0},\end{split}

where u:DUdu\colon D\to U\coloneqq\mathbb{R}^{d} is the fluid velocity and p:Dp\colon D\to\mathbb{R}, the pressure, acting as a Lagrange multiplier to enforce the divergence constraint divu=0\operatorname{div}u=0, and u0u_{0} is the initial condition. Here, we take the spatial domain DD to be the dd-dimensional torus D𝕋dD\coloneqq\mathbb{T}^{d}, and we denote the phase space by UdU\coloneqq\mathbb{R}^{d}. The divergence is defined as divu:=i=1dxiui\operatorname{div}u:=\sum_{i=1}^{d}\partial_{x_{i}}u^{i} and p:=(x1p,,xdp)\nabla p:=(\partial_{x_{1}}p,\dots,\partial_{x_{d}}p)^{\top} is the spatial gradient. The parameter ε>0\varepsilon>0 denotes the viscosity and is proportional to the reciprocal of the Reynolds number. It is well-known that many flows of interest are characterized by high to very-high Reynolds numbers. Hence, one is interested in studying what happens when viscosity ε\varepsilon equal to zero. In this formal limit ε0\varepsilon\to 0, one obtains the incompressible Euler equations, which are the prototypical models for an ideal fluid.

The question of whether the ε0\varepsilon\to 0 limit is a good approximation of (1.1) is of great practical relevance and has received considerable attention, both from a physical as well as mathematical point of view. Furthermore, it plays an essential role in computational fluid dynamics, as many numerical methods for the Euler equations, as well as large eddy simulations (LES) for Navier-Stokes equations, can be viewed as discretizations of (1.1) with ε\varepsilon of the order of the discretization parameter.

While in two spatial dimension, the convergence of a sequence of solutions {uε}ε>0\{u^{\varepsilon}\}_{\varepsilon>0} of (1.1) to a solution of the Euler equations has been proved rigorously for many settings, see e.g., [14, 52, 13, 15], it turns out to be very challenging in 3D. Leray [47] proved in 1934 the existence of “Leray–Hopf solutions” of (1.1), which are weak solutions of (1.1), i.e., they satisfy (1.1) in the sense of distributions and in addition an energy inequality of the form

(1.2) D|u(t,x)|2𝑑x+ε0tD|u(s,x)|2𝑑x𝑑sD|u0(x)|2𝑑x,t0.\int_{D}|u(t,x)|^{2}dx+\varepsilon\int_{0}^{t}\int_{D}|\nabla u(s,x)|^{2}dxds\leqslant\int_{D}|u_{0}(x)|^{2}dx,\quad t\geqslant 0.

(Here and in the remainder, we suppress the dependence of uεu^{\varepsilon} on ε\varepsilon for convenience and write just uu.) Hence, if the initial data u0u_{0} lies in Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U), such solutions satisfy uL([0,T];Ldiv2(D;d))L2([0,T];Hdiv1(D;d))u\in L^{\infty}\big{(}[0,T];L^{2}_{\operatorname{div}}(D;\mathbb{R}^{d})\big{)}\cap L^{2}\big{(}[0,T];H^{1}_{\operatorname{div}}(D;\mathbb{R}^{d})\big{)}. (Here, Ldiv2L^{2}_{\operatorname{div}} and Hdiv1H^{1}_{\operatorname{div}} are the weakly divergence free functions in L2L^{2} and H1H^{1}, respectively.) However, as ε0\varepsilon\to 0, the L2L^{2}-bound on the gradient of uu, that stems from the energy inequality (1.2), no longer suffices for deriving sufficient compactness of the sequence {u}ε>0\{u\}_{\varepsilon>0} in L2L^{2}—which would be needed to pass to the limit in the nonlinear terms. It appears that, at least as far as global solutions are concerned, there is currently no means of gaining sufficient compactness through other conserved quantities or bootstrapping; in fact, it is unclear if global solutions of higher regularity than the one given by (1.2) exist in 3D. Closely related to this issue is the lack of stability estimates, i.e., well-posedness of Leray-Hopf solutions [22, 45]. The main obstruction to better regularity or stability estimates is caused by the nonlinear convective term (u)u(u\cdot\nabla)u. The role of the nonlinear term and possible instabilities in the Leray-Hopf solutions are often related to the issue of turbulence in fluid flows.

The theory of mathematical turbulence was initiated in the 1930s and 1940s by Taylor, Richardson, Kolmogorov and others, see [32] and references therein, and has since influenced fluid mechanics, as well as atmospheric sciences and plasma physics heavily. In his sequence of three papers [42, 41, 43], nowadays referred to as K41, Kolmogorov took a probabilistic approach to turbulence and formulated basic hypotheses about fluid flow at high Reynolds numbers and derived predictions based on these. Many of these have later been confirmed by experiments. The idea of studying equations (1.1) in a probabilistic setting has since been taken up again in many works, in different frameworks, by adding stochastic forcing terms to (1.1), see e.g. [26, 49, 10], or taking uncertain or measure-valued initial data, e.g. [17]. In the latter case, the solution of (1.1) may not be a function any more but instead a time-parametrized probability measure on the phase space. Global existence of such measure-valued solutions for incompressible flows has been shown in 3D, and even the passage to the limit ε0\varepsilon\to 0 can be made rigorous in this case [17]. However, measure-valued solutions are generally not unique, which can be shown by counterexample even in the case of Burgers’ equation [23]. Hence, measure-valued solutions are too broad a solution concept to resolve the problem of non-uniqueness, and more information or constraints need to be added.

To overcome this, in [24], it was suggested to take into account the (time) evolution of all possible multi-point spatial correlations. Instead of a single probability measure on the phase space UU, such a statistical solution is a family of probability measures on the phase space and products of the phase space UkU^{k}, for kk\in\mathbb{N}, corresponding to the multi-point correlations. Hence, one can interpret the solution as a measure-valued solution augmented with information about higher order spatial correlations. From a practical point of view, this approach is very natural, as often only averaged quantities of interest of the fluid flow can be observed. Moreover, it is also in line with Kolmogorov’s turbulence theory, as this theory studies statistical properties of the fluid and makes predictions about these. The system of equations that arises for the higher order correlations is also known as the Friedman–Keller infinite chain of moment equations [39, 58] and finite closure relation for this infinite family of equations have been studied for small and large Reynolds numbers in [34, 36, 35, 37].

An alternative point of view in this context, is to consider instead probability measures on a space of suitable initial conditions; in the case of (1.1) this would be Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U). Equation (1.1) is then interpreted as a Liouville equation on an infinite dimensional function space and the solution is a mapping assigning to each time tt a probability measure on Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U). This setting was first considered by Prodi [53] and later on extensively studied by Foiaş and collaborators [27, 28, 30, 29], see also [38]. A closely related notion of statistical solutions was studied by Vishik and Fursikov [58]. Foiaş and his collaborators proved existence of such solutions in 2D and 3D, uniqueness in 2D, and further properties related to turbulence [29]. The relations between the Foiaş and Prodi notion of statistical solutions and the Vishik-Fursikov version were explored in [31, 8, 9]. The latter work etends the notion of statistical solutions to other relevant PDEs in fluid mechanics.

Given this plurality of definitions of statistical solutions, it is natural to examine, if and under what conditions, these solution concepts are equivalent. The first goal of this paper is to prove that both these concepts of statistical solutions of the incompressible Navier-Stokes equations (1.1) are equivalent as long as a statistical version of the of the energy inequality (1.2) holds.

The second and main goal of this paper is to investigate the vanishing viscosity limit of the statistical solutions of incompressible Navier-Stokes equations. Under an weak scaling assumption on the Navier-Stokes statistical solutions, we will use compactness criteria, presented recently in [25], to prove that vanishing viscosity limits of the statistical solutions of Navier-Stokes equations are statistical solutions of the incompressible Euler equations.

Our weak scaling assumption is a significantly weaker version of the scaling hypothesis of Kolmogorov’s 1941 theory and allows for intermittent corrections. Our main technical tool is a statisical version of the well-known Kármán–Howarth–Monin relation [32, 16, 48], that relates the evolution of 2-point correlations to the longitudinal structure function S3S_{\|}^{3}, which is, roughly speaking, defined as

S3(||)((u(x+)u(x))^)3,^||.S_{\|}^{3}(|\ell|)\coloneqq\Bigl{\langle}\Bigl{(}(u(x+\ell)-u(x))\cdot\widehat{\ell}\Bigr{)}^{3}\Bigr{\rangle},\quad\widehat{\ell}\coloneqq\frac{\ell}{|\ell|}.

Here \langle\cdot\rangle is a suitable average of the flow.

Thus, by characterizing this vanishing viscosity limit, we establish a rigorous relationship between the incompressible Navier-Stokes and Euler equations, while accommodating physically observed facts about turbulent flows in this description.

The remainder of this article is organized as follows: In Section 2, we introduce the concept of correlation measures and in Section 3 we show the equivalence of statistical solutions as introduced by Foiaş and Prodi with families of correlation measures satisfying the Friedman–Keller chain of moment equations. Then in Section 4, we consider the passage to the limit ε0\varepsilon\to 0 and conclude with an appendix with technical results.

2. Correlation measures

In this section, we recall the definition of correlation measures and some important properties of them from [24, 25]. We start by introducing the necessary notation.

2.1. Notation

For kk\in\mathbb{N}, k1k\geqslant 1, we denote the tensor products

Dk=D××Dktimes,Uk=UUktimes.D^{k}=\underbrace{D\times\dots\times D}_{k\,\textrm{times}},\quad U^{k}=\underbrace{U\otimes\dots\otimes U}_{k\,\textrm{times}}.

If XX is a topological space then we let (X)\mathscr{B}(X) denote the Borel σ\sigma-algebra on XX, we let (X)\mathscr{M}(X) denote the set of signed Radon measures on (X,(X))(X,\mathscr{B}(X)), and we let 𝒫(X)(X)\mathscr{P}(X)\subset\mathscr{M}(X) denote the set of all probability measures on (X,(X))(X,\mathscr{B}(X)), i.e. all 0μ(X)0\leqslant\mu\in\mathscr{M}(X) with μ(X)=1\mu(X)=1 (see e.g. [3, 7, 40]). For kk\in\mathbb{N} and a multiindex α{0,1}k\alpha\in\{0,1\}^{k} we write |α|=α1++αk|\alpha|=\alpha_{1}+\dots+\alpha_{k} and α¯=𝟙α=(1α1,,1αk)\bar{\alpha}=\mathbbm{1}-\alpha=(1-\alpha_{1},\dots,1-\alpha_{k}), and we let xαx_{\alpha} be the vector of length |α||\alpha| consisting of the elements xix_{i} of xx for which αi\alpha_{i} is non-zero. For a vector x=(x1,,xk)x=(x_{1},\dots,x_{k}) we write x^i=(x1,,xi1,xi+1,,xk)\widehat{x}_{i}=(x_{1},\dots,x_{i-1},x_{i+1},\dots,x_{k}). For a vector ξ=(ξ1,,ξk)\xi=(\xi_{1},\dots,\xi_{k}) we write |ξα|=|ξ1|α1|ξk|αk|\xi^{\alpha}|=|\xi_{1}|^{\alpha_{1}}\cdots|\xi_{k}|^{\alpha_{k}} with the convention 00=10^{0}=1.

2.1.1. Carathéodory functions

If EE and VV are Euclidean spaces then a measurable function g:E×Vg\colon E\times V\to\mathbb{R} is called a Carathéodory function if ξg(x,ξ)\xi\mapsto g(x,\xi) is continuous for a.e. yEy\in E and yg(y,ξ)y\mapsto g(y,\xi) is measurable for every ξV\xi\in V (see e.g. [1, Section 4.10]). Given kk\in\mathbb{N} and a Carathéodory function g=g(x,ξ):Dk×Ukg=g(x,\xi)\colon D^{k}\times U^{k}\to\mathbb{R} we define the functional Lg:Lp(D;U)L_{g}\colon L^{p}(D;U)\to\mathbb{R} by

(2.1) Lg(u)Dkg(x1,,xk,u(x1),,u(xk))𝑑x.L_{g}(u)\coloneqq\int_{D^{k}}g(x_{1},\dots,x_{k},u(x_{1}),\dots,u(x_{k}))\,dx.

(It is not obvious that LgL_{g} is continuous, or even well-defined; see [24].) We denote the set of Carathéodory functions depending on space and time by 0k([0,T),D;U)L1([0,T)×Dk;C0(Uk))\mathcal{H}^{k}_{0}([0,T),D;U)\coloneqq L^{1}([0,T)\times D^{k};C_{0}(U^{k})) and its dual space by 0k([0,T),D;U)Lw([0,T)×Dk;(Uk))\mathcal{H}^{k*}_{0}([0,T),D;U)\coloneqq L^{\infty}_{w}([0,T)\times D^{k};\mathscr{M}(U^{k})) (see e.g. [5]).

In the following, we will focus on a specific type of Carathéodory functions. In particular, for p1p\geqslant 1 we let k,p([0,T],D;U)\mathcal{H}^{k,p}([0,T],D;U) denote the space of Carathéodory functions g:[0,T]×Dk×Ukg\colon[0,T]\times D^{k}\times U^{k}\to\mathbb{R} satisfying

(2.2) |g(t,x,ξ)|α{0,1}kφ|α¯|(t,xα¯)|ξα|pxDk,ξUk|g(t,x,\xi)|\leqslant\sum_{\alpha\in\{0,1\}^{k}}\varphi_{|\bar{\alpha}|}(t,x_{\bar{\alpha}})|\xi^{\alpha}|^{p}\qquad\forall\ x\in D^{k},\ \xi\in U^{k}

for nonnegative functions φiL([0,T];L1(Di))\varphi_{i}\in L^{\infty}([0,T];L^{1}(D^{i})), i=0,1,,ki=0,1,\dots,k. We let 1k,p([0,T],D;U)k,p([0,T],D;U)\mathcal{H}^{k,p}_{1}([0,T],D;U)\subset\mathcal{H}^{k,p}([0,T],D;U) denote the subspace of functions gg satisfying the local Lipschitz condition

(2.3) |g(t,x,ζ)g(t,y,ξ)|ψ(t)i=1k|ζiξi|max(|ξi|,|ζi|)p1h(t,x^i,ξ^i)+O(|xy|)h~(t,x,ξ)\begin{split}\big{|}g(t,x,\zeta)-g(t,y,\xi)\big{|}\leqslant&\ \psi(t)\sum_{i=1}^{k}|\zeta_{i}-\xi_{i}|\max\big{(}|\xi_{i}|,|\zeta_{i}|\big{)}^{p-1}h(t,\widehat{x}_{i},\widehat{\xi}_{i})\\ &+O(|x-y|)\widetilde{h}(t,x,\xi)\end{split}

for every xDkx\in D^{k}, yBr(x)y\in B_{r}(x) for some r>0r>0, for some nonnegative hk1,p([0,T],D;U)h\in\mathcal{H}^{k-1,p}([0,T],D;U) and 0ψ(t)L([0,T])0\leqslant\psi(t)\in L^{\infty}([0,T]) and some h~k,p([0,T],D;U)\widetilde{h}\in\mathcal{H}^{k,p}([0,T],D;U). (Note that the term h~\widetilde{h} was not present in [25, Definition 2.2], but one can generalize the results of that paper to include such a term.)

We also denote for a parametrized probability measure νkLw([0,T)×Dk;(Uk))\nu^{k}\in L^{\infty}_{w}([0,T)\times D^{k};\mathscr{M}(U^{k})) and a Carathéodory function gg the pairing

νk,gk=Dkνt,xk,g(t,x)𝑑x\bigl{\langle}\nu^{k},g\bigr{\rangle}_{\mathcal{H}^{k}}=\int_{D^{k}}\bigl{\langle}\nu^{k}_{t,x},g(t,x)\bigr{\rangle}\,dx

(where νxk,g(t,x)=Ukg(t,x,ξ)𝑑νt,xk(ξ)\bigl{\langle}\nu^{k}_{x},g(t,x)\bigr{\rangle}=\int_{U^{k}}g(t,x,\xi)\,d\nu^{k}_{t,x}(\xi) is the usual duality pairing between Radon measures (Uk)\mathscr{M}(U^{k}) and continuous functions C0(Uk)C_{0}(U^{k})).

2.2. Definitions

We are now in a position to define time-dependent correlation measures.

Definition 2.1.

A time-dependent correlation measure is a collection 𝝂=(ν1,ν2,)\bm{\nu}=(\nu^{1},\nu^{2},\dots) of functions νk0k([0,T),D;U)\nu^{k}\in\mathcal{H}^{k*}_{0}([0,T),D;U) such that

  1. (i)

    νt,xk𝒫(Uk)\nu^{k}_{t,x}\in\mathscr{P}(U^{k}) for a.e. (t,x)[0,T]×Dk(t,x)\in[0,T]\times D^{k}, and the map xνt,xk,fx\mapsto\bigl{\langle}\nu^{k}_{t,x},f\bigr{\rangle} is measurable for every fCb(Uk)f\in C_{b}(U^{k}) and almost every t[0,T]t\in[0,T]. (In other words, νtk\nu^{k}_{t} is a Young measure from DkD^{k} to UkU^{k}.)

  2. (ii)

    LpL^{p} integrability:

    (2.4) esssupt[0,T)(Dνt,x1,|ξ|pdx)1/pc<+\operatorname*{ess\,sup}_{t\in[0,T)}\left(\int_{D}\bigl{\langle}\nu^{1}_{t,x},|\xi|^{p}\bigr{\rangle}\,dx\right)^{1/p}\leqslant c<+\infty
  3. (iii)

    Diagonal continuity (DC):

    (2.5) 0Tωrp(νt2)𝑑t0 as r0 for all T(0,T),\int_{0}^{T^{\prime}}\omega_{r}^{p}\big{(}\nu^{2}_{t}\big{)}\,dt\to 0\qquad\text{ as }r\to 0\text{ for all }T^{\prime}\in(0,T),

    where

    ωrp(νt2)DBr(x)νt,x,y2,|ξ1ξ2|p𝑑y𝑑x\omega_{r}^{p}(\nu^{2}_{t})\coloneqq\int_{D}\mathchoice{{\vbox{\hbox{$\textstyle-$ }}\kern-7.31105pt}}{{\vbox{\hbox{$\scriptstyle-$ }}\kern-5.50664pt}}{{\vbox{\hbox{$\scriptscriptstyle-$ }}\kern-4.4333pt}}{{\vbox{\hbox{$\scriptscriptstyle-$ }}\kern-3.96664pt}}\!\int_{B_{r}(x)}\bigl{\langle}\nu^{2}_{t,x,y},|\xi_{1}-\xi_{2}|^{p}\bigr{\rangle}\,dy\,dx

    is called the modulus of continuity of 𝝂\bm{\nu}.

We denote the set of all time-dependent correlation measures by p([0,T),D;U)\mathcal{L}^{p}([0,T),D;U).

In [25] (and see [24] for a time-independent version), the following equivalence between time-dependent correlation measures and parametrized probability measures on Lp(D)L^{p}(D) was proved:

Theorem 2.2.

For every time-dependent correlation measure 𝛎p([0,T),D;U)\bm{\nu}\in\mathcal{L}^{p}([0,T),D;U) there is a unique (up to subsets of [0,T)[0,T) of Lebesgue measure 0) map μ:[0,T)𝒫(Lp(D;U))\mu\colon[0,T)\to\mathscr{P}(L^{p}(D;U)) such that

  1. (i)

    the map

    (2.6) tμt,Lg=LpDkg(x,u(x))𝑑x𝑑μt(u)t\mapsto\bigl{\langle}\mu_{t},L_{g}\bigr{\rangle}=\int_{L^{p}}\int_{D^{k}}g(x,u(x))\,dx\,d\mu_{t}(u)

    is measurable for all g0k(D;U)g\in\mathcal{H}^{k}_{0}(D;U),

  2. (ii)

    μ\mu is LpL^{p}-bounded:

    (2.7) esssupt[0,T)LpuLpp𝑑μt(u)cp<\operatorname*{ess\,sup}_{t\in[0,T)}\int_{L^{p}}\|u\|_{L^{p}}^{p}\,d\mu_{t}(u)\leqslant c^{p}<\infty
  3. (iii)

    μ\mu is dual to 𝝂\bm{\nu}: the identity

    (2.8) Dkνtk,g(x)𝑑x=LpDkg(x,u(x))𝑑x𝑑μt(u)\int_{D^{k}}\bigl{\langle}\nu_{t}^{k},g(x)\bigr{\rangle}\,dx=\int_{L^{p}}\int_{D^{k}}g(x,u(x))\,dx\,d\mu_{t}(u)

    holds for a.e. t[0,T)t\in[0,T), every g0k(D;U)g\in\mathcal{H}^{k}_{0}(D;U) and all kk\in\mathbb{N}.

Conversely, for every μ:[0,T)𝒫(Lp(D;U))\mu\colon[0,T)\to\mathscr{P}(L^{p}(D;U)) satisfying (i) and (ii), there is a unique correlation measure 𝛎p([0,T),D;U)\bm{\nu}\in\mathcal{L}^{p}([0,T),D;U) satisfying (iii).

We also have the following “Compactness” Theorem for time-dependent correlation measures [25, Theorem 2.21]

Theorem 2.3.

Let 𝛎np([0,T),D;U)\bm{\nu}_{n}\in\mathcal{L}^{p}([0,T),D;U) for n=1,2,n=1,2,\dots be a sequence of correlation measures such that

(2.9) supnesssupt[0,T)(Dνn;t,x1,|ξ|pdx)1/pc<+\displaystyle\sup_{n\in\mathbb{N}}\operatorname*{ess\,sup}_{t\in[0,T)}\left(\int_{D}\bigl{\langle}\nu^{1}_{n;t,x},|\xi|^{p}\bigr{\rangle}\,dx\right)^{1/p}\leqslant c<+\infty
(2.10) limr0lim supn0Tωrp(νn,t2)𝑑t=0\displaystyle\lim_{r\to 0}\limsup_{n\to\infty}\int_{0}^{T^{\prime}}\omega_{r}^{p}\bigl{(}\nu^{2}_{n,t}\bigr{)}\,dt=0

for some c>0c>0 and all T[0,T)T^{\prime}\in[0,T). Then there exists a subsequence (nj)j=1(n_{j})_{j=1}^{\infty} and some 𝛎p([0,T),D;U)\bm{\nu}\in\mathcal{L}^{p}([0,T),D;U) such that

  1. (i)

    𝝂nj𝝂\bm{\nu}_{n_{j}}\overset{*}{\rightharpoonup}\bm{\nu} as jj\to\infty, that is, νnjk,gkνk,gk\bigl{\langle}\nu^{k}_{n_{j}},g\bigr{\rangle}_{\mathcal{H}^{k}}\to\bigl{\langle}\nu^{k},g\bigr{\rangle}_{\mathcal{H}^{k}} for every g0k([0,T),D;U)g\in\mathcal{H}^{k}_{0}([0,T),D;U) and every kk\in\mathbb{N}

  2. (ii)

    νt1,|ξ|p1cp\bigl{\langle}\nu^{1}_{t},|\xi|^{p}\bigr{\rangle}_{\mathcal{H}^{1}}\leqslant c^{p} for a.e. t[0,T)t\in[0,T)

  3. (iii)

    0Tωrp(νt2)𝑑tlim infn0Tωrp(νn,t2)𝑑t\int_{0}^{T^{\prime}}\omega_{r}^{p}\big{(}\nu^{2}_{t}\big{)}\,dt\leqslant\liminf_{n\to\infty}\int_{0}^{T^{\prime}}\omega_{r}^{p}\big{(}\nu^{2}_{n,t}\big{)}\,dt for every r>0r>0 and T[0,T)T^{\prime}\in[0,T)

  4. (iv)

    for kk\in\mathbb{N}, let φLloc1([0,T)×Dk)\varphi\in L^{1}_{\textrm{loc}}([0,T)\times D^{k}) and κC(Uk)\kappa\in C(U^{k}) be nonnegative, and let g(t,x,ξ)φ(t,x)κ(ξ)g(t,x,\xi)\coloneqq\varphi(t,x)\kappa(\xi). Then

    (2.11) νk,gklim infjνnjk,gk.\bigl{\langle}\nu^{k},g\bigr{\rangle}_{\mathcal{H}^{k}}\leqslant\liminf_{j\to\infty}\bigl{\langle}\nu_{n_{j}}^{k},g\bigr{\rangle}_{\mathcal{H}^{k}}.
  5. (v)

    Assume moreover that DdD\subset\mathbb{R}^{d} is compact, T<T<\infty and that 𝝂n\bm{\nu}_{n} have uniformly bounded support, in the sense that

    (2.12) uLpRfor μtn-a.e. uLp(D;U) for every n,a.et(0,T),\|u\|_{L^{p}}\leqslant R\qquad\text{for $\mu^{n}_{t}$-a.e.\ }u\in L^{p}(D;U)\text{ for every }n\in\mathbb{N},~{}a.e~{}t\in(0,T),

    with μtn𝒫T(Lp(D;U))\mu_{t}^{n}\in\mathscr{P}_{T}(L^{p}(D;U)) being dual to 𝝂n\bm{\nu}_{n}, then the following observables converge strongly:

    (2.13) limjDk|0T(νnj;t,xk,g(t,x)νt,xk,g(t,x))𝑑t|𝑑x=0\lim_{j\to\infty}\int_{D^{k}}\left|\int_{0}^{T}\left(\bigl{\langle}\nu^{k}_{n_{j};t,x},g(t,x)\bigr{\rangle}-\bigl{\langle}\nu^{k}_{t,x},g(t,x)\bigr{\rangle}\right)\,dt\right|\,dx=0

    for every g1k,p([0,T],D;U)g\in\mathcal{H}^{k,p}_{1}([0,T],D;U).

3. Statistical solutions

The goal of this section is to show that the statistical solutions of Navier–Stokes as introduced by Foiaş and Prodi [27, 28, 29, 30, 53] are equivalent to families of correlation measures as introduced in [24] that satisfy the Friedman–Keller system of moment equations. For the sake of simplicity we will assume that the support of the initial measure μ0\mu_{0} lies in a bounded set BL2(D;U)B\subset L^{2}(D;U), that is,

(3.1) supp(μ0)BL2(D;U).\text{supp}(\mu_{0})\subset B\subset L^{2}(D;U).

3.1. The Leray projector

We recall first that the Helmholtz–Leray projector, or simply Leray projector, is the linear map :L2(D;U)Ldiv2(D;U){vL2(D;U):divv=0,v𝑑x=0}\mathbb{P}\colon L^{2}(D;U)\to L^{2}_{\operatorname{div}}(D;U)\coloneqq\{v\in L^{2}(D;U):\operatorname{div}v=0,\,\int vdx=0\} that projects a vector field fL2(D;U)f\in L^{2}(D;U) to its divergence free component, that is f=f+ψff=\mathbb{P}f+\nabla\psi_{f} with div(f)=0\operatorname{div}(\mathbb{P}f)=0 and f,ψfL2(D;U)\mathbb{P}f,\nabla\psi_{f}\in L^{2}(D;U). One can show that ψf\nabla\psi_{f} is orthogonal in L2(D;U)L^{2}(D;U) to any function uLdiv2(D;U)u\in L^{2}_{\operatorname{div}}(D;U),

Duψfdx=0.\int_{D}u\cdot\nabla\psi_{f}\,dx=0.

For functions in the tensor product space L2(Dk;Uk)L^{2}(D^{k};U^{k}) we let xi\mathbb{P}_{x_{i}} denote the Leray projector in the iith component, i.e., φ=xiφ+xiψφ,i\varphi=\mathbb{P}_{x_{i}}\varphi+\nabla_{x_{i}}\psi_{\varphi,i} where divxi(xiφ)=0\operatorname{div}_{x_{i}}(\mathbb{P}_{x_{i}}\varphi)=0, and ψφ,iL2(Dk;Uk1)\psi_{\varphi,i}\in L^{2}(D^{k};U^{k-1}) with xiψφ,iL2(Dk;Uk)\nabla_{x_{i}}\psi_{\varphi,i}\in L^{2}(D^{k};U^{k}).

3.2. Definitions

We will start by recalling the different definitions of statistical solutions introduced in [29, 31, 24].

Definition 3.1 (Definition 3.2 in [24]).

Let ε0\varepsilon\geqslant 0. The Friedman–Keller system of moment equations, defined for time-dependent correlation measures 𝝂2([0,T),D;U)\bm{\nu}\in\mathcal{L}^{2}([0,T),D;U), is the hierarchy of equations

(3.2) 0TDkUk(ξ1ξk):φt(t,x)dνt,xk(ξ)dxdt+DkUk(ξ1ξk):φ(0,x)dν0,xk(ξ)dx+i=1k0TDkUk(ξ1(ξiξi)ξk):xiφ(t,x)dνt,xk(ξ)dxdt=εi=1k0TDkUk(ξ1ξk):Δxiφ(t,x)dνt,xk(ξ)dxdt\begin{split}&\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\frac{\partial\varphi}{\partial t}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\\ &+\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx\\ &+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\\ =&\ -\varepsilon\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\Delta_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\end{split}

for all kk\in\mathbb{N}, for all φCc2([0,T)×Dk;Uk)\varphi\in C^{2}_{c}([0,T)\times D^{k};U^{k}) with divxiφ=0\operatorname{div}_{x_{i}}\varphi=0 for all i=1,,ki=1,\dots,k, along with the divergence constraint

(3.3) DkUkξ1ξα+1(ξ+1)αk(ξk)𝑑νt,xk(ξ)x1,,xφ(x)𝑑x=0,\int_{D^{k}}\int_{U^{k}}\xi_{1}\otimes\dots\otimes\xi_{\ell}\otimes\alpha_{\ell+1}(\xi_{\ell+1})\otimes\dots\otimes\alpha_{k}(\xi_{k})\,d\nu_{t,x}^{k}(\xi)\cdot\nabla_{x_{1},\dots,x_{\ell}}\varphi(x)\,dx=0,

where x1,,x=(x1,,x)\nabla_{x_{1},\dots,x_{\ell}}=(\nabla_{x_{1}},\dots,\nabla_{x_{\ell}})^{\top}, 1k1\leqslant\ell\leqslant k\in\mathbb{N}, for all φH1(Dk;Uk)\varphi\in H^{1}(D^{k};U^{k-\ell}), αjC(U;U)\alpha_{j}\in C(U;U), with αj(v)C(1+|v|2)\alpha_{j}(v)\leqslant C(1+|v|^{2}) for all j=1,,kj=1,\dots,k.

If 𝝂\bm{\nu} solves the Friedman–Keller system of moments equations and in addition satisfies the energy inequality

(3.4) k=0KakDkUk|ξ1|2|ξk|2𝑑νt,xk(ξ)𝑑x\displaystyle\sum_{k=0}^{K}a_{k}\int_{D^{k}}\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{t,x}^{k}(\xi)\,dx
+2εk=0Kaki=1kj=1dlimh01h20tDkUk+1|ξ1|2|ξiξk+1|2|ξk|2𝑑νt,(x,xi+h𝐞j)k+1(ξ,ξk+1)𝑑x𝑑s\displaystyle+2\varepsilon\sum_{k=0}^{K}a_{k}\sum_{i=1}^{k}\sum_{j=1}^{d}\lim_{h\rightarrow 0}\frac{1}{h^{2}}\int_{0}^{t}\!\!\int_{D^{k}}\!\int_{U^{k+1}}\!\!\!|\xi_{1}|^{2}\dots|\xi_{i}-\xi_{k+1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{t,(x,x_{i}+h\mathbf{e}_{j})}^{k+1}(\xi,\xi_{k+1})\,dx\,ds
k=0KakDkUk|ξ1|2|ξk|2𝑑ν0,xk(ξ)𝑑x\displaystyle\leqslant\sum_{k=0}^{K}a_{k}\int_{D^{k}}\!\!\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{0,x}^{k}(\xi)\,dx

for all KK\in\mathbb{N} and aka_{k}\in\mathbb{R}, k=0,,Kk=0,\dots,K such that pK(s)=k=0Kakskp_{K}(s)=\sum_{k=0}^{K}a_{k}s^{k} is a nonnegative, nondecreasing polynomial for s[0,R]s\in[0,R] for RR sufficiently large related to the support of the correlation measure (see (2.12)), then we call 𝝂\bm{\nu} a Friedman–Keller statistical solution of the Navier–Stokes (when ε>0\varepsilon>0) or Euler (when ε=0\varepsilon=0) equations.

Remark 3.2.

Friedman–Keller statistical solutions are analogous to the definition of statistical solutions for hyperbolic systems of conservation laws as introduced in [24, 25] for compressible flows.

Remark 3.3.

By a standard argument for weak solutions to continuity equations, the map tνt,k,ξ1ξkt\mapsto\langle\nu^{k}_{t,\cdot},\,\xi_{1}\otimes\cdots\otimes\xi_{k}\rangle is weakly continuous for every kk\in\mathbb{N}; see e.g. [2, Remark 2.2].

Remark 3.4 (Formulation of (3.2) with non-divergence free test functions).

Denoting kx1xk\mathbb{P}_{k}\coloneqq\mathbb{P}_{x_{1}}\dots\mathbb{P}_{x_{k}} (cf. Section 3.1), we can replace the divergence-free test function φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}) in (3.2) by kφ\mathbb{P}_{k}\varphi for an arbitrary φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}). Using the fact that the Leray projection is self-adjoint and that, by (3.3),

divxiUk(ξ1ξk)𝑑νt,xk(ξ)=0i=1,,k,a.e. (t,x)[0,T]×Dk\operatorname{div}_{x_{i}}\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k})\,d\nu_{t,x}^{k}(\xi)=0\qquad\forall\ i=1,\dots,k,\quad\text{a.e. }\,(t,x)\in[0,T]\times D^{k}

we observe that

kUk(ξ1ξk)𝑑νt,xk(ξ)=Uk(ξ1ξk)𝑑νt,xk(ξ),a.e. t[0,T],xDk.\mathbb{P}_{k}\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k})\,d\nu_{t,x}^{k}(\xi)=\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k})\,d\nu_{t,x}^{k}(\xi),\quad\text{a.e. }\,t\in[0,T],\,x\in D^{k}.

Therefore,

0TDkUk(ξ1ξk):kφt(t,x)dνt,xk(ξ)dxdt\displaystyle\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\frac{\partial\mathbb{P}_{k}\varphi}{\partial t}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt
=\displaystyle= 0TDkk(Uk(ξ1ξk)𝑑νt,xk(ξ)):φt(t,x)dxdt\displaystyle\int_{0}^{T}\!\!\int_{D^{k}}\!\!\mathbb{P}_{k}\left(\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k})\,d\nu_{t,x}^{k}(\xi)\right):\frac{\partial\varphi}{\partial t}(t,x)\,dx\,dt
=\displaystyle= 0TDkUk(ξ1ξk)𝑑νt,xk(ξ):φt(t,x)dxdt.\displaystyle\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k})\,d\nu_{t,x}^{k}(\xi):\frac{\partial\varphi}{\partial t}(t,x)\,dx\,dt.

Similarly,

DkUk(ξ1ξk):kφ(0,x)dν0,xk(ξ)dx=DkUk(ξ1ξk):φ(0,x)dν0,xk(ξ)dx,\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\mathbb{P}_{k}\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx=\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx,

and

i=1k0TDkUk(ξ1ξk):Δxikφ(t,x)dνt,xk(ξ)dxdt\displaystyle\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\Delta_{x_{i}}\mathbb{P}_{k}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt
=\displaystyle= i=1k0TDkUk(ξ1ξk):Δxiφ(t,x)dνt,xk(ξ)dxdt,\displaystyle\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\Delta_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt,

the last one being true due to the fact that the Laplacian and the Leray projection commute on the torus. Hence, the weak formulation (3.2) can be rewritten as

(3.5) 0TDkUk(ξ1ξk):φt(t,x)dνt,xk(ξ)dxdt+DkUk(ξ1ξk):φ(0,x)dν0,xk(ξ)dx+i=1k0TDkUk(ξ1(ξiξi)ξk):xiφ(t,x)dνt,xk(ξ)dxdt=εi=1k0TDkUk(ξ1ξk):Δxiφ(t,x)dνt,xk(ξ)dxdt+i=1k0TDkUk(ξ1(ξiξi)ξk):xixiψφ,i(t,x)dνt,xk(ξ)dxdt\begin{split}&\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\frac{\partial\varphi}{\partial t}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt+\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx\\ &+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\\ =&-\varepsilon\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\Delta_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\\ &+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\nabla_{x_{i}}\psi_{\varphi,i}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt\end{split}

where φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}). The terms

(3.6) i=1k0TDkUk(ξ1(ξiξi)ξk):xixiψφ,i(t,x)dνt,xk(ξ)dxdt\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\nabla_{x_{i}}\psi_{\varphi,i}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt

correspond to the pressure in the deterministic setting.

To define statistical solutions in the sense of Foiaş and Prodi, we need to introduce some notation. We denote by Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) the space of divergence free L2(D;U)L^{2}(D;U)-vector fields and by Hdiv1(D;U)H^{1}_{\operatorname{div}}(D;U) the space of divergence free functions in H1(D;U)H^{1}(D;U) (these can be obtained as the closures of C(D;U){divu=0}C^{\infty}(D;U)\cap\{\operatorname{div}u=0\} in L2(D;U)L^{2}(D;U) and H1(D;U)H^{1}(D;U), respectively, with suitable integral conditions:

Ldiv2(D;U)={uL2(𝕋d):divu=0,𝕋du(x)𝑑x=0},Hdiv1(D;U)={uH1(𝕋d):divu=0,𝕋du(x)𝑑x=0},\begin{split}L^{2}_{\operatorname{div}}(D;U)&=\left\{u\in L^{2}(\mathbb{T}^{d})\ :\ \operatorname{div}u=0,\ \int_{\mathbb{T}^{d}}u(x)\,dx=0\right\},\\ H^{1}_{\operatorname{div}}(D;U)&=\left\{u\in H^{1}(\mathbb{T}^{d})\ :\ \operatorname{div}u=0,\ \int_{\mathbb{T}^{d}}u(x)\,dx=0\right\},\end{split}

for periodic boundary conditions. We denote the L2L^{2}-inner product by

(u,v)=Du(x)v(x)𝑑x,(u,v)=\int_{D}u(x)v(x)\,dx,

and for ε>0\varepsilon>0 the H1H^{1}-inner product by

a(u,v)=εi=1dDuxiuxi𝑑x,a(u,v)=\varepsilon\sum_{i=1}^{d}\int_{D}\frac{\partial u}{\partial x^{i}}\cdot\frac{\partial u}{\partial x^{i}}\,dx,

Define the Stokes operator AA by

Au=Δu,for all uD(A)=Hdiv1(D;U)H2(D;U),ε(Au,v)=a(u,v),for all u,vD(A1/2),\begin{split}Au&=-\mathbb{P}\Delta u,\quad\text{for all }u\in D(A)=H^{1}_{\operatorname{div}}(D;U)\cap H^{2}(D;U),\\ \varepsilon(Au,v)&=a(u,v),\quad\text{for all }u,v\in D(A^{1/2}),\end{split}

where \mathbb{P} is the Leray projector, and the skew-symmetric trilinear form bb by

(3.7) b(u,v,w)D(u)vw𝑑x=(B(u,v),w),u,v,wD(A1/2)B(u)B(u,u).\begin{split}b(u,v,w)&\coloneqq\int_{D}(u\cdot\nabla)v\cdot w\,dx=(B(u,v),w),\quad u,v,w\in D(A^{1/2})\\ B(u)&\coloneqq B(u,u).\end{split}

We can then write the Navier–Stokes equations in the functional formulation: Let T>0T>0, u0Ldiv2(D;U)u_{0}\in L^{2}_{\operatorname{div}}(D;U), find uL([0,T];Ldiv2(D;U))L2([0,T];Hdiv1(D;U))u\in L^{\infty}([0,T];L^{2}_{\operatorname{div}}(D;U))\cap L^{2}([0,T];H^{1}_{\operatorname{div}}(D;U)) with uddtuL1([0,T];D(A1/2))u^{\prime}\coloneqq\frac{d}{dt}u\in L^{1}([0,T];D(A^{-1/2})) such that

(3.8) u+εAu+B(u)=0u^{\prime}+\varepsilon Au+B(u)=0

and u(0)=u0u(0)=u_{0} in a suitable sense. This corresponds to the weak formulation

(3.9) ddt(u,v)+a(u,v)+b(u,u,v)=0for all vHdiv1(D;U).\frac{d}{dt}(u,v)+a(u,v)+b(u,u,v)=0\qquad\text{for all }v\in H^{1}_{\operatorname{div}}(D;U).

If we denote

(3.10) F(t,u)εAuB(u),F(t,u)\coloneqq-\varepsilon Au-B(u),

the functional formulation becomes

(3.11) u(t)=F(t,u(t)).u^{\prime}(t)=F(t,u(t)).

We need the following class of test functions:

Notation 3.5.

[29] Let 𝒯cyl\mathscr{T}_{\text{cyl}} denote the class of cylindrical test functions consisting of the real-valued functionals Φ=Φ(u)\Phi=\Phi(u) that depend on a finite number kk\in\mathbb{N} of components of uu, that is,

Φ(u)=φ((u,g1),,(u,gk)),\Phi(u)=\varphi\bigl{(}(u,g_{1}),\dots,(u,g_{k})\bigr{)},

where φCc1(k)\varphi\in C^{1}_{c}(\mathbb{R}^{k}) and g1,,gkH1(D;U)g_{1},\dots,g_{k}\in H^{1}(D;U). Let 𝒯cyl0\mathscr{T}_{\text{cyl}}^{0} denote the subset of such functions which satisfy g1,,gkHdiv1(D;U)g_{1},\dots,g_{k}\in H^{1}_{\operatorname{div}}(D;U). We denote by Φ\Phi^{\prime} the differential of Φ\Phi in Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U), which can be expressed as

Φ(u)=j=1kjφ((u,g1),,(u,gk))gj,\Phi^{\prime}(u)=\sum_{j=1}^{k}\partial_{j}\varphi\bigl{(}(u,g_{1}),\dots,(u,g_{k})\bigr{)}g_{j},

where jφ\partial_{j}\varphi is the derivative of φ\varphi with respect to its jjth component.

We can now define statistical solutions in the sense of Foiaş and Prodi. We will use the definition as it stated in their newer work [31, Def. 3.2]:

Definition 3.6 (Foiaş–Prodi [29, 27, 30, 31]).

A family of probability measures (μt)0tT(\mu_{t})_{0\leqslant t\leqslant T} on Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) is a Foiaş–Prodi statistical solution of the Navier–Stokes equations on Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) with initial data μ0\mu_{0} if

  1. (a)

    The function

    (3.12) tLdiv2φ(u)𝑑μt(u),t\mapsto\int_{L^{2}_{\operatorname{div}}}\varphi(u)\,d\mu_{t}(u),

    is measurable on [0,T][0,T] for every φCb(Ldiv2(D;U))\varphi\in C_{b}(L^{2}_{\operatorname{div}}(D;U));

  2. (b)

    μ\mu satisfies the weak formulation

    (3.13) Ldiv2Φ(u)𝑑μt(u)=Ldiv2Φ(u)𝑑μ0(u)+0tLdiv2(F(s,u),Φ(u))𝑑μs(u)𝑑s\int_{L^{2}_{\operatorname{div}}}\Phi(u)\,d\mu_{t}(u)=\int_{L^{2}_{\operatorname{div}}}\Phi(u)\,d\mu_{0}(u)+\int_{0}^{t}\int_{L^{2}_{\operatorname{div}}}(F(s,u),\Phi^{\prime}(u))\,d\mu_{s}(u)\,ds

    for all t[0,T]t\in[0,T] and all cylindrical test functions Φ𝒯cyl0\Phi\in\mathscr{T}_{\text{cyl}}^{0}, where FF is given in (3.10).

  3. (c)

    μ\mu satisfies the strengthened mean energy inequality: For any ψC1(,)\psi\in C^{1}(\mathbb{R},\mathbb{R}) nonnegative, nondecreasing with bounded derivative and t[0,T]t\in[0,T], the inequality

    (3.14) Ldiv2ψ(uL2(D)2)𝑑μt(u)+2ε0tLdiv2ψ(uL2(D)2)|u|H1(D)2𝑑μs(u)Ldiv2ψ(uL2(D)2)𝑑μ0(u)\int_{L^{2}_{\operatorname{div}}}\psi(\|u\|_{L^{2}(D)}^{2})\,d\mu_{t}(u)+2\varepsilon\int_{0}^{t}\int_{L^{2}_{\operatorname{div}}}\psi^{\prime}(\|u\|_{L^{2}(D)}^{2})|u|_{H^{1}(D)}^{2}\,d\mu_{s}(u)\\ \leqslant\int_{L^{2}_{\operatorname{div}}}\psi(\|u\|_{L^{2}(D)}^{2})\,d\mu_{0}(u)

    holds.

  4. (d)

    The function

    (3.15) tLdiv2ψ(uL2(D)2)𝑑μt(u)t\mapsto\int_{L^{2}_{\operatorname{div}}}\psi(\left\|u\right\|_{L^{2}(D)}^{2})\,d\mu_{t}(u)

    is continuous at t=0t=0 from the right, for any function ψC1(,)\psi\in C^{1}(\mathbb{R},\mathbb{R}) nonnegative, nondecreasing with bounded derivative.

Remark 3.7.

Note that, as a consequence of the energy inequality (3.14) for ψ(s)=s\psi(s)=s, the function

tLdiv2uL2(D)2𝑑μt(u),t\mapsto\int_{L^{2}_{\operatorname{div}}}\|u\|_{L^{2}(D)}^{2}\,d\mu_{t}(u),

belongs to L([0,T])L^{\infty}([0,T]) and the function

(3.16) tLdiv2|u|H1(D)2𝑑μt(u),t\mapsto\int_{L^{2}_{\operatorname{div}}}|u|_{H^{1}(D)}^{2}\,d\mu_{t}(u),

belongs to L1([0,T])L^{1}([0,T]). Notice also that (3.13) implies that

tLdiv2Φ(u)𝑑μt(u)t\mapsto\int_{L^{2}_{\operatorname{div}}}\Phi(u)d\mu_{t}(u)

for Φ(u)\Phi(u) a cylindrical test function, is continuous since

Ldiv2(F(s,u),Φ(u))𝑑μt(u)\int_{L^{2}_{\operatorname{div}}}(F(s,u),\Phi^{\prime}(u))\,d\mu_{t}(u)

is locally integrable. Combining this fact with condition (c), conditionn (d) follows directly.

3.3. Equivalence between the solution concepts

Next, we show that the Friedman–Keller statistical solutions in Definition 3.1 and the Foiaş–Prodi statistical solutions in Definition 3.6 are in fact the same.

Theorem 3.8 (Foiaş–Temam statistical solutions satisfy the Friedman–Keller system).

Let μ\mu be a Foiaş–Prodi statistical solution such that the initial condition μ0\mu_{0} has bounded support,

supp(μ0)BLdiv2(D;U),B{uL2(D;U):uL2(D)R}\text{supp}(\mu_{0})\subset B\subset L^{2}_{\operatorname{div}}(D;U),\qquad B\subset\big{\{}u\in L^{2}(D;U)\ :\ \|u\|_{L^{2}(D)}\leqslant R\big{\}}

for some R>0R>0. Then μ\mu corresponds (cf. Theorem 2.2) to a correlation measure 𝛎t=(ν1,ν2,)\bm{\nu}_{t}=(\nu^{1},\nu^{2},\dots) that is a statistical solution in the Friedman–Keller sense (cf. Definition 3.1).

Conversely, we have:

Theorem 3.9 (Friedman–Keller solutions are Foiaş–Prodi statistical solutions).

Let 𝛎=(𝛎t)0tT\bm{\nu}=(\bm{\nu}_{t})_{0\leqslant t\leqslant T} be a Friedman–Keller statistical solution of Navier–Stokes (cf. Definition 3.1) with bounded support, i.e.,

(3.17) DkUk|ξ1|2|ξk|2𝑑νt,xk(ξ)𝑑xRk<,\int_{D^{k}}\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{t,x}^{k}(\xi)\,dx\leqslant R^{k}<\infty,

for some 0<R<0<R<\infty, every kk\in\mathbb{N}, and almost every t[0,T]t\in[0,T]. Then 𝛎\bm{\nu} corresponds to a probability measure μ=(μt)0tT\mu=(\mu_{t})_{0\leqslant t\leqslant T} on a bounded set of Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) which is a Foiaş–Prodi statistical solution of the Navier–Stokes equations (cf. Definition 3.6).

The proofs of these two results are given in Appendix A.

Foiaş et al. have shown existence of Foiaş–Prodi statistical solutions for the (forced) Navier–Stokes equations, see e.g. [27, 28, 29]. Using these equivalence theorems, this implies existence of statistical solutions via correlation measures as in Definition 3.1.

Remark 3.10.

The equivalence theorems 3.8 and 3.9 are restricted to probability measures with bounded support. It should be possible to extend these results to probability measures having sufficiently fast decay near infinity; however, the proofs would become significantly more technical. We have therefore decided to restrict ourselves to probability measures with bounded support.

4. Vanishing viscosity limit of statistical solutions of Navier–Stokes

The goal of this section is to pass to the inviscid limit ε0\varepsilon\to 0 under the assumption of weak statistical scaling (c.f. Section 4.3, Assumption 1). We will first prove a rigorous result on the longitudinal third order structure function

(4.1) S3(τ,r)0τLdiv2𝕊2D((u(x+rn)u(x))n)3𝑑x𝑑S(n)𝑑μt(u)𝑑t,S_{\|}^{3}(\tau,r)\coloneqq\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}\big{(}(u(x+rn)-u(x))\cdot n\big{)}^{3}\,dxdS(n)\,d\mu_{t}(u)\,dt,

and then relate it to the similarly defined second order structure function using the weak scaling assumption. Together with weak statistical anisotropy, this yields diagonal continuity of the correlation measures 𝝂ε\bm{\nu}^{\varepsilon} that is needed to apply the compactness theorem 2.3 and pass to the limit. The proof of the scaling estimate for the third order structure function (4.1) in Lemma 4.2 and  4.3 largely follows the proof of a similar result for martingale solutions of stochastic Navier–Stokes equations in [6]. To simplify notation, we will omit writing the dependence of 𝝂\bm{\nu} and μ\mu on ε\varepsilon in the following sections.

4.1. Kármán–Howarth–Monin relation

The key to deriving an estimate on the behavior of the third order structure function (4.1) is the so-called Kármán–Howarth–Monin (KHM) relation [16] that describes the evolution of the second correlation marginal. Similar relations have been derived before for various settings (stochastic, forced, etc.), see [32, 50, 16, 48, 6, 21, 19]. For statistical solutions we derive:

Proposition 4.1.

Let 𝛎\bm{\nu} be a Friedman–Keller statistical solution of the Navier–Stokes equations. Then the second correlation marginal ν2\nu^{2} satisfies the Kármán–Howarth–Monin relation for correlation measures:

(4.2) ijDDU2ξ1iξ2j𝑑ντ,x,x+h2(ξ)𝑑xσij(h)𝑑hijDDU2ξ1iξ2j𝑑ν0,x,x+h2(ξ)𝑑xσij(h)𝑑h+12ijk0τDDU2(ξ2iξ1i)(ξ2jξ1j)(ξ2kξ1k)𝑑νt,x,x+h2(ξ)𝑑xhkσij(h)dhdt=εij0τDDU2(ξ1iξ2i)(ξ1jξ2j)𝑑νt,x,x+h2(ξ)𝑑xΔhσij(h)𝑑h𝑑t.\begin{split}&\quad\sum_{ij}\int_{D}\!\int_{D}\!\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\,d\nu_{\tau,x,x+h}^{2}(\xi)\,dx\,\sigma^{ij}(h)\,dh-\sum_{ij}\int_{D}\!\int_{D}\!\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\,\sigma^{ij}(h)\,dh\\ &\quad+\frac{1}{2}\sum_{ijk}\int_{0}^{\tau}\!\!\int_{D}\!\int_{D}\!\int_{U^{2}}(\xi_{2}^{i}-\xi^{i}_{1})(\xi_{2}^{j}-\xi_{1}^{j})(\xi_{2}^{k}-\xi^{k}_{1})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,\partial_{h^{k}}\sigma^{ij}(h)\,dh\,dt\\ &=-\varepsilon\sum_{ij}\int_{0}^{\tau}\!\!\int_{D}\!\int_{D}\!\int_{U^{2}}\!\!(\xi_{1}^{i}-\xi_{2}^{i})(\xi_{1}^{j}-\xi_{2}^{j})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,\Delta_{h}\sigma^{ij}(h)\,dh\,dt.\end{split}

for any τ>0\tau>0, where σ=(σij)i,j=13\sigma=(\sigma_{ij})_{i,j=1}^{3} is any smooth, compactly supported, isotropic rank 2 tensor – that is, any σCc2(d,d×d)\sigma\in C^{2}_{c}(\mathbb{R}^{d},\mathbb{R}^{d\times d}) of the form

(4.3) σ(h)=(ω1(|h|)𝐈+ω2(|h|)h^h^)where h^={h/|h|h00h=0\sigma(h)=\left(\omega_{1}(|h|)\mathbf{I}+\omega_{2}(|h|)\hat{h}\otimes\hat{h}\right)\qquad\text{where }\hat{h}=\begin{cases}h/|h|&h\neq 0\\ 0&h=0\end{cases}

for ω1,ω2Cc2()\omega_{1},\omega_{2}\in C^{2}_{c}(\mathbb{R}).

Proof.

We consider equation (3.5) for k=2k=2 with the test function φ(t,x,y)=η(t,yx)\varphi(t,x,y)=\eta(t,y-x) (for simplicity replacing x1x_{1} and x2x_{2} by xx and yy and writing in component form η=(ηij)ij\eta=(\eta^{ij})_{ij}),

(4.4) ij0TDDU2ξ1iξ2jηijt(t,yx)𝑑νt,x,y2(ξ)𝑑x𝑑y𝑑t+ijDDU2ξ1iξ2jηij(0,yx)𝑑ν0,x,y2(ξ)𝑑x𝑑y+ijk0TDDU2ξ1iξ1kξ2jxkηij(t,yx)dνt,x,y2(ξ)dxdydt+ijk0TDDU2ξ1iξ2jξ2kykηij(t,yx)dνt,x,y2(ξ)dxdydt=εij0TDDU2ξ1iξ2j(Δx+Δy)ηij(t,yx)𝑑νt,x,y2(ξ)𝑑x𝑑y𝑑t+ijk0TDDU2ξ1iξ1kξ2jxkxiψη,1j(t,yx)dνt,x,y2(ξ)dxdydt+ijk0TDDU2ξ1iξ2jξ2kykyjψη,2i(t,yx)dνt,x,y2(ξ)dxdydt\begin{split}&\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\frac{\partial\eta^{ij}}{\partial t}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\\ &\quad+\sum_{ij}\int_{D}\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\eta^{ij}(0,y-x)\,d\nu_{0,x,y}^{2}(\xi)\,dx\,dy\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{x^{k}}\eta^{ij}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{y^{k}}\eta^{ij}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\\ &=-\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}(\Delta_{x}+\Delta_{y})\eta^{ij}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{x^{k}}\partial_{x^{i}}\psi^{j}_{\eta,1}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{y^{k}}\partial_{y^{j}}\psi^{i}_{\eta,2}(t,y-x)\,d\nu_{t,x,y}^{2}(\xi)\,dx\,dy\,dt\end{split}

Since ψη,1\psi_{\eta,1} solves Δxψη,1=divxη\Delta_{x}\psi_{\eta,1}=\operatorname{div}_{x}\eta and ψη,2\psi_{\eta,2} solves Δyψη,2=divyη=divxη\Delta_{y}\psi_{\eta,2}=\operatorname{div}_{y}\eta=-\operatorname{div}_{x}\eta, we have ψη,1=(Δx)1divxη\psi_{\eta,1}=(\Delta_{x})^{-1}\operatorname{div}_{x}\eta and ψη,2=(Δy)1divxη=(Δx)1divxη\psi_{\eta,2}=-(\Delta_{y})^{-1}\operatorname{div}_{x}\eta=-(\Delta_{x})^{-1}\operatorname{div}_{x}\eta and so ψηψη,1=ψη,2\psi_{\eta}\coloneqq\psi_{\eta,1}=-\psi_{\eta,2} (up to additive constants). Using this and changing the integration variables to xx and hyxh\coloneqq y-x, we obtain

(4.5) ij0TDDU2ξ1iξ2jηijt(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t+ijDDU2ξ1iξ2jηij(0,h)𝑑ν0,x,x+h2(ξ)𝑑x𝑑hijk0TDDU2ξ1iξ1kξ2jhkηij(t,h)dνt,x,x+h2(ξ)dxdhdt+ijk0TDDU2ξ1iξ2jξ2khkηij(t,h)dνt,x,x+h2(ξ)dxdhdt=2εij0TDDU2ξ1iξ2jΔhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t+ijk0TDDU2ξ1iξ1kξ2jhkhiψηj(t,h)dνt,x,x+h2(ξ)dxdhdtijk0TDDU2ξ1iξ2jξ2khkhjψηi(t,h)dνt,x,x+h2(ξ)dxdhdt\begin{split}&\sum_{ij}\int_{0}^{T}\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\frac{\partial\eta^{ij}}{\partial t}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ij}\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\eta^{ij}(0,h)\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\,dh\\ &\quad-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &=-2\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{h^{k}}\partial_{h^{i}}\psi^{j}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{h^{k}}\partial_{h^{j}}\psi^{i}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\end{split}

where ψη=ψη,1\psi_{\eta}=\psi_{\eta,1}. The cubic terms can be rewritten using the following simple fact (which can also be found in Frisch [32, equation (6.13)], and in a similar, weak form in [6]):

(4.6) 2ijk0TDDU2ξ1iξ2j(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt=ijk0TDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt.\begin{split}&\quad-2\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &=\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt.\end{split}

The proof of this is postponed to the end of this proof. Using this, we can rewrite (4.5) as

(4.7) ij0TDDU2ξ1iξ2jηijt(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t+ijDDU2ξ1iξ2jηij(0,h)𝑑ν0,x,x+h2(ξ)𝑑x𝑑h12ijk0TDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt=2εij0TDDU2ξ1iξ2jΔhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t+ijk0TDDU2ξ1iξ1kξ2jhkhiψηj(t,h)dνt,x,x+h2(ξ)dxdhdtijk0TDDU2ξ1iξ2jξ2khkhjψηi(t,h)dνt,x,x+h2(ξ)dxdhdt.\begin{split}&\quad\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\frac{\partial\eta^{ij}}{\partial t}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ij}\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\eta^{ij}(0,h)\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\,dh\\ &\quad-\frac{1}{2}\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &=-2\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{h^{k}}\partial_{h^{i}}\psi^{j}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{h^{k}}\partial_{h^{j}}\psi^{i}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt.\end{split}

Since

ij0TDDU2ξ1iξ1jΔhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t\displaystyle\quad\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{1}^{j}\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
=ij0TDUξ1iξ1j𝑑νt,x1(ξ)𝑑xDΔhηij(t,h)𝑑h𝑑t=0,\displaystyle=\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\int_{U}\!\!\xi_{1}^{i}\xi_{1}^{j}\,d\nu_{t,x}^{1}(\xi)\,dx\int_{D}\Delta_{h}\eta^{ij}(t,h)\,dh\,dt=0,

we can rewrite

2εij0TDDU2ξ1iξ2jΔhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t\displaystyle\quad-2\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
=εij0TDDU2(ξ1iξ2i)(ξ1jξ2j)Δhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t.\displaystyle=\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!(\xi_{1}^{i}-\xi_{2}^{i})(\xi_{1}^{j}-\xi_{2}^{j})\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt.

Moreover, we have for symmetric, smooth and compactly supported rank 2 tensors η\eta of the form

(4.8) η(t,h)=ω1(t,|h|)𝐈+ω2(t,|h|)h^h^,\eta(t,h)=\omega_{1}(t,|h|)\mathbf{I}+\omega_{2}(t,|h|)\hat{h}\otimes\hat{h},

with ψηi=(Δh)1divhηi,\psi^{i}_{\eta}=-(\Delta_{h})^{-1}\operatorname{div}_{h}\eta^{i,\cdot},

(4.9) ijk0TDDU2ξ1iξ1kξ2jhkhiψηj(t,h)dνt,x,x+h2(ξ)dxdhdt=0,ijk0TDDU2ξ1iξ2jξ2khkhjψηi(t,h)dνt,x,x+h2(ξ)dxdhdt=0,\begin{split}&\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{1}^{k}\xi_{2}^{j}\partial_{h^{k}}\partial_{h^{i}}\psi^{j}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt=0,\\ &\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{h^{k}}\partial_{h^{j}}\psi^{i}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt=0,\end{split}

whose proof is postponed to the end of this proof. Using this, (4.7) becomes

(4.10) ij0TDDU2ξ1iξ2jηijt(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t+ijDDU2ξ1iξ2jηij(0,h)𝑑ν0,x,x+h2(ξ)𝑑x𝑑h12ijk0TDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt=εij0TDDU2(ξ1iξ2i)(ξ1jξ2j)Δhηij(t,h)𝑑νt,x,x+h2(ξ)𝑑x𝑑h𝑑t.\begin{split}&\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\frac{\partial\eta^{ij}}{\partial t}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &\quad+\sum_{ij}\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\eta^{ij}(0,h)\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\,dh\\ &\quad-\frac{1}{2}\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ &=\varepsilon\sum_{ij}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!(\xi_{1}^{i}-\xi_{2}^{i})(\xi_{1}^{j}-\xi_{2}^{j})\Delta_{h}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt.\end{split}

Let θδ\theta_{\delta} be a sequence of smooth, uniformly bounded functions with the property that θδ𝟏(0,τ](t)\theta_{\delta}\to\mathbf{1}_{(0,\tau]}(t) for every tt as δ0\delta\to 0. If we now use a test function

(4.11) η(t,h)=σ(h)θδ(t),\eta(t,h)=\sigma(h)\theta_{\delta}(t),

where σ\sigma is of the form (4.3), then we can use the weak continuity in time of the moments Ukξ1ξk𝑑νt,xk(ξ)\int_{U^{k}}\xi_{1}\otimes\dots\otimes\xi_{k}\,d\nu^{k}_{t,x}(\xi) to obtain for any τ>0\tau>0, as δ0\delta\to 0,

(4.12) ijDDU2ξ1iξ2j𝑑ντ,x,x+h2(ξ)𝑑xσij(h)𝑑h+ijDDU2ξ1iξ2j𝑑ν0,x,x+h2(ξ)𝑑xσij(h)𝑑h12ijk0τDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)𝑑νt,x,x+h2(ξ)𝑑xhkσij(h)dhdt=εij0τDDU2(ξ1iξ2i)(ξ1jξ2j)𝑑νt,x,x+h2(ξ)𝑑xΔhσij(h)𝑑h𝑑t.\begin{split}&\quad-\sum_{ij}\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\,d\nu_{\tau,x,x+h}^{2}(\xi)\,dx\sigma^{ij}(h)\,dh+\sum_{ij}\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!\xi_{1}^{i}\xi_{2}^{j}\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\sigma^{ij}(h)\,dh\\ &\quad-\frac{1}{2}\sum_{ijk}\int_{0}^{\tau}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\partial_{h^{k}}\sigma^{ij}(h)\,dh\,dt\\ &=\varepsilon\sum_{ij}\int_{0}^{\tau}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\!\!(\xi_{1}^{i}-\xi_{2}^{i})(\xi_{1}^{j}-\xi_{2}^{j})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\Delta_{h}\sigma^{ij}(h)\,dh\,dt.\qed\end{split}
Proof of (4.6).

We expand the right hand side:

ijk0TDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt\displaystyle\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
=ijk0TDDU2(ξ1iξ1jξ1kξ2iξ2jξ2k)𝑑νt,x,x+h2(ξ)𝑑xhkηij(t,h)dhdt\displaystyle=\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}\xi_{1}^{j}\xi_{1}^{k}-\xi_{2}^{i}\xi_{2}^{j}\xi_{2}^{k})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,\partial_{h^{k}}\eta^{ij}(t,h)\,dh\,dt
+ijk0TDDU2(ξ2iξ2jξ1kξ1iξ1jξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt\displaystyle\quad+\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{2}^{i}\xi_{2}^{j}\xi_{1}^{k}-\xi_{1}^{i}\xi_{1}^{j}\xi_{2}^{k})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
ijk0TDDU2(ξ1iξ2jξ1kξ2iξ1jξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt\displaystyle\quad-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}\xi_{2}^{j}\xi_{1}^{k}-\xi_{2}^{i}\xi_{1}^{j}\xi_{2}^{k})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
ijk0TDDU2(ξ2iξ1jξ1kξ1iξ2jξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt.\displaystyle\quad-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{2}^{i}\xi_{1}^{j}\xi_{1}^{k}-\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt.

The first term on the right hand side is zero since η\eta is compactly supported (after changing the integration variable from xx to xhx-h in one of the terms). The second term on the right hand side vanishes using the divergence constraint (3.3). Using that η\eta and ν\nu are symmetric, the last two terms are identical and so

ijk0TDDU2(ξ1iξ2i)(ξ1jξ2j)(ξ1kξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt=2ijk0TDDU2(ξ2iξ1jξ1kξ1iξ2jξ2k)hkηij(t,h)dνt,x,x+h2(ξ)dxdhdt,\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{1}^{i}-\xi^{i}_{2})(\xi_{1}^{j}-\xi_{2}^{j})(\xi_{1}^{k}-\xi^{k}_{2})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt\\ =-2\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}(\xi_{2}^{i}\xi_{1}^{j}\xi_{1}^{k}-\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k})\partial_{h^{k}}\eta^{ij}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt,

which proves the claim. ∎

Proof of (4.9).

We consider the second expression, assume η\eta is of the form

(4.13) η(t,h)=ω1(t,|h|)𝐈+ω2(t,|h|)h^h^,\eta(t,h)=\omega_{1}(t,|h|)\mathbf{I}+\omega_{2}(t,|h|)\hat{h}\otimes\hat{h},

where ωi\omega_{i}, i=1,2i=1,2 are compactly supported in the torus. Then using that ψηi=(Δh)1divhηi,\psi^{i}_{\eta}=-(\Delta_{h})^{-1}\operatorname{div}_{h}\eta^{i,\cdot} (the first term is treated in a similar way)

E\displaystyle E ijk0TDDU2ξ1iξ2jξ2khkhjψηi(t,h)dνt,x,x+h2(ξ)dxdhdt\displaystyle\coloneqq-\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\partial_{h^{k}}\partial_{h^{j}}\psi^{i}_{\eta}(t,h)\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,dh\,dt
=ijk0TDDU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑xhkhjΔ1(hηi(t,h))dhdt\displaystyle=\sum_{ijk\ell}\int_{0}^{T}\!\!\int_{D}\!\!\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\,\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\partial_{h^{\ell}}\eta^{i\ell}(t,h)\right)\,dh\,dt
=ijk0TDhkhjΔ1(DU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑x)hηi(t,h)dhdt.\displaystyle=\sum_{ijk\ell}\int_{0}^{T}\!\!\int_{D}\!\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\right)\partial_{h^{\ell}}\eta^{i\ell}(t,h)dh\,dt.

We note that since ωi\omega_{i} have compact support, we can write in polar coordinates

hηi(t,h)=(ω1(t,|h|)+ω2(t,|h|)+2ω2(t,|h|)|h|G(t,|h|))h^i\sum_{\ell}\partial_{h^{\ell}}\eta^{i\ell}(t,h)=\biggl{(}\underbrace{\omega^{\prime}_{1}(t,|h|)+\omega_{2}^{\prime}(t,|h|)+2\frac{\omega_{2}(t,|h|)}{|h|}}_{\eqqcolon\,G(t,|h|)}\biggr{)}\hat{h}^{i}

and so

E=ijk0TDhkhjΔ1(DU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)hηi(t,h)dx)dhdt=ijk0TDhkhjΔ1(DU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑x)G(t,|h|)h^idhdt=jk0T0|h|=rhkhjΔ1(DU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑x)h^idS(h)G(t,r)drdt=ijk0T0|h|rhihkhjΔ1(DU2ξ1iξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑x)dhG(t,r)drdt=jk0T0|h|rhkhjΔ1divh(DU2ξ1ξ2jξ2k𝑑νt,x,x+h2(ξ)𝑑x)dhG(t,r)drdt=0,\begin{split}E&=\sum_{ijk\ell}\int_{0}^{T}\!\!\int_{D}\!\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\partial_{h^{\ell}}\eta^{i\ell}(t,h)\,dx\right)dh\,dt\\ &=\sum_{ijk}\int_{0}^{T}\!\!\int_{D}\!\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\int_{D}\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\right)G(t,|h|)\hat{h}^{i}\,dh\,dt\\ &=\sum_{jk}\int_{0}^{T}\!\!\int_{0}^{\infty}\int_{|h|=r}\!\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\int_{D}\!\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\right)\hat{h}^{i}dS(h)\,G(t,r)\,dr\,dt\\ &=\sum_{ijk}\int_{0}^{T}\!\!\int_{0}^{\infty}\int_{|h|\leqslant r}\!\partial_{h^{i}}\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\left(\int_{D}\!\int_{U^{2}}\xi_{1}^{i}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\right)dh\,G(t,r)\,dr\,dt\\ &=\sum_{jk}\int_{0}^{T}\!\!\int_{0}^{\infty}\int_{|h|\leqslant r}\!\partial_{h^{k}}\partial_{h^{j}}\Delta^{-1}\operatorname{div}_{h}\left(\int_{D}\!\int_{U^{2}}\xi_{1}\xi_{2}^{j}\xi_{2}^{k}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\right)dh\,G(t,r)\,dr\,dt=0,\end{split}

where we used the divergence theorem in the second to last identity and the divergence constraint (3.3) for the last identity. ∎

4.2. Scaling of third order structure functions

Next, we use the KHM-relation (4.2) to derive a scaling relation for the averaged third order structure function in terms of the measure μt\mu_{t},

(4.14) S03(τ,r)=0τLdiv2𝕊2D|u(x)u(x+rn)|2(u(x+rn)u(x))n𝑑x𝑑S(n)𝑑μt(u)𝑑t,S^{3}_{0}(\tau,r)=\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\fint_{\mathbb{S}^{2}}\int_{D}\big{|}u(x)-u(x+rn)\big{|}^{2}\big{(}u(x+rn)-u(x)\big{)}\cdot n\,dx\,dS(n)\,d\mu_{t}(u)\,dt,

which will be more convenient to work with for this purpose. We have:

Lemma 4.2.

Let μt\mu_{t} be a Foiaş–Prodi statistical solution of the Navier–Stokes equations (cf. Definition 3.6). Then

(4.15) |S03(τ,r)r|2E0,\left|\frac{S^{3}_{0}(\tau,r)}{r}\right|\leqslant 2E_{0},

where E0E_{0} is the initial energy,

(4.16) E0Ldiv2(D;U)u(x)L2(D)2𝑑μ0(u).E_{0}\coloneqq\int_{L^{2}_{\operatorname{div}}(D;U)}\left\|u(x)\right\|^{2}_{L^{2}(D)}\,d\mu_{0}(u).
Proof.

We take a test function of the form σ(h)=ω(|h|)𝐈\sigma(h)=\omega(|h|)\mathbf{I} in the KHM-relation (4.2) with ω\omega having compact support in [0,0.5)[0,0.5). A little bit of algebra yields (denoting h^k=hk/|h|\hat{h}_{k}=h^{k}/|h|)

hkω(|h|)=ω(|h|)h^k,\partial_{h^{k}}\omega(|h|)=\omega^{\prime}(|h|)\hat{h}^{k},

and so (4.2) for this particular test function reads

(4.17) DDU2ξ1ξ2𝑑ντ,x,x+h2(ξ)𝑑xω(|h|)𝑑hDDU2ξ1ξ2𝑑ν0,x,x+h2(ξ)𝑑xω(|h|)𝑑h+120τDDU2|ξ1ξ2|2(ξ2ξ1)h^𝑑νt,x,x+h2(ξ)𝑑xω(|h|)𝑑h𝑑t=ε0τDDU2|ξ1ξ2|2𝑑νt,x,x+h2(ξ)𝑑xΔhω(|h|)𝑑h𝑑t.\begin{split}&\int_{D}\int_{D}\!\int_{U^{2}}\!\!\xi_{1}\cdot\xi_{2}\,d\nu_{\tau,x,x+h}^{2}(\xi)\,dx\omega(|h|)\,dh-\int_{D}\int_{D}\!\int_{U^{2}}\!\!\xi_{1}\cdot\xi_{2}\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\omega(|h|)\,dh\\ &+\frac{1}{2}\int_{0}^{\tau}\!\!\int_{D}\int_{D}\!\int_{U^{2}}|\xi_{1}-\xi_{2}|^{2}(\xi_{2}-\xi_{1})\cdot\hat{h}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\omega^{\prime}(|h|)\,dh\,dt\\ =&-\varepsilon\int_{0}^{\tau}\!\!\int_{D}\int_{D}\!\int_{U^{2}}\!\!|\xi_{1}-\xi_{2}|^{2}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\Delta_{h}\omega(|h|)\,dh\,dt.\end{split}

In terms of the statistical solution (μt)t>0(\mu_{t})_{t>0} this is

(4.18) Ldiv2DDu(x)u(x+h)𝑑xω(|h|)𝑑h𝑑μτ(u)Ldiv2DDu(x)u(x+h)𝑑xω(|h|)𝑑h𝑑μ0(u)+120τLdiv2DD|u(x)u(x+h)|2(u(x+h)u(x))h^𝑑xω(|h|)𝑑h𝑑μt(u)𝑑t=ε0τLdiv2DD|u(x)u(x+h)|2𝑑xΔhω(|h|)𝑑h𝑑μt(u)𝑑t.\begin{split}&\quad\int_{L^{2}_{\operatorname{div}}}\!\int_{D}\int_{D}\!\!u(x)\cdot u(x+h)\,dx\omega(|h|)\,dh\,d\mu_{\tau}(u)-\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!u(x)u(x+h)\,dx\,\omega(|h|)\,dh\,d\mu_{0}(u)\\ &\quad+\frac{1}{2}\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!|u(x)-u(x+h)|^{2}(u(x+h)-u(x))\cdot\hat{h}\,\,dx\omega^{\prime}(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=-\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!|u(x)-u(x+h)|^{2}\,dx\Delta_{h}\omega(|h|)\,dh\,d\mu_{t}(u)\,dt.\end{split}

The last term can also be written as

(4.19) ε0τLdiv2DD|u(x)u(x+h)|2𝑑xΔhω(|h|)𝑑h𝑑μt(u)𝑑t=2ε0τLdiv2DDhu(x+h)(u(x)u(x+h))𝑑xhω(|h|)𝑑h𝑑μt(u)𝑑t=2ε0τLdiv2DDxu(x+h)(u(x)u(x+h))𝑑xhω(|h|)𝑑h𝑑μt(u)𝑑t=2ε0τLdiv2DDxu(x)(u(xh)u(x))𝑑xhω(|h|)𝑑h𝑑μt(u)𝑑t=2ε0τLdiv2DDxu(x):hu(xh)dxω(|h|)dhdμt(u)dt=2ε0τLdiv2DDxu(x+h):xu(x)dxω(|h|)dhdμt(u)dt\begin{split}&\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!|u(x)-u(x+h)|^{2}\,dx\Delta_{h}\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=2\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!\nabla_{h}u(x+h)\cdot(u(x)-u(x+h))\,dx\nabla_{h}\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=2\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!\nabla_{x}u(x+h)\cdot(u(x)-u(x+h))\,dx\nabla_{h}\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=2\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!\nabla_{x}u(x)\cdot(u(x-h)-u(x))\,dx\nabla_{h}\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=-2\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!\nabla_{x}u(x):\nabla_{h}u(x-h)\,dx\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\\ &=2\varepsilon\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!\nabla_{x}u(x+h):\nabla_{x}u(x)\,dx\omega(|h|)\,dh\,d\mu_{t}(u)\,dt\end{split}

Changing to spherical coordinates and using the definition of S03S^{3}_{0}, (4.14), we obtain

120S03(τ,r)r2ω(r)𝑑r\displaystyle\frac{1}{2}\int_{0}^{\infty}S_{0}^{3}(\tau,r)r^{2}\omega^{\prime}(r)dr
=2ε00τLdiv2𝕊2Dxu(x):xu(x+rn)dxdS(n)dμt(u)dtr2ω(r)dr\displaystyle=-2\varepsilon\int_{0}^{\infty}\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!\nabla_{x}u(x):\nabla_{x}u(x+rn)\,dxdS(n)\,d\mu_{t}(u)\,dtr^{2}\omega(r)dr
0Ldiv2𝕊2Du(x)u(x+rn)𝑑x𝑑S(n)𝑑μτ(u)r2ω(r)𝑑r\displaystyle\quad-\int_{0}^{\infty}\int_{L^{2}_{\operatorname{div}}}\!\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot u(x+rn)\,dxdS(n)\,d\mu_{\tau}(u)r^{2}\omega(r)dr
+0Ldiv2𝕊2Du(x)u(x+rn)𝑑x𝑑S(n)𝑑μ0(u)r2ω(r)𝑑r.\displaystyle\quad+\int_{0}^{\infty}\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot u(x+rn)\,dxdS(n)\,d\mu_{0}(u)r^{2}\omega(r)dr.

We denote

m2(τ,r)\displaystyle m_{2}(\tau,r) Ldiv2𝕊2Du(x)u(x+rn)𝑑x𝑑S(n)𝑑μτ(u)\displaystyle\coloneqq\int_{L^{2}_{\operatorname{div}}}\!\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot u(x+rn)\,dxdS(n)\,d\mu_{\tau}(u)
v(τ,r)\displaystyle v(\tau,r) 0τLdiv2𝕊2Dxu(x):xu(x+rn)dxdS(n)dμt(u)dt.\displaystyle\coloneqq\int_{0}^{\tau}\!\!\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!\nabla_{x}u(x):\nabla_{x}u(x+rn)\,dxdS(n)\,d\mu_{t}(u)\,dt.

Since (μt)t>0(\mu_{t})_{t>0} is supported on functions in L([0,);Ldiv2(D;U))L2([0,);Hdiv1(D;U))L^{\infty}([0,\infty);L^{2}_{\operatorname{div}}(D;U))\cap L^{2}([0,\infty);H^{1}_{\operatorname{div}}(D;U)), S03S_{0}^{3} is a continuous function. Moreover, notice that due to the a priori bounds following from the energy inequality (3.14), both m2m_{2} and vv are uniformly bounded and continuous in rr and τ\tau (for the continuity in τ\tau of the first quantity, one needs weak time continuity of the moments which follows the fact that they satisfy the equations (4.2) where all the terms are integrable). We obtain

120S03(τ,r)r2ω(r)𝑑r\displaystyle\frac{1}{2}\int_{0}^{\infty}S_{0}^{3}(\tau,r)r^{2}\omega^{\prime}(r)dr =2ε0v(τ,r)r2ω(r)𝑑r\displaystyle=-2\varepsilon\int_{0}^{\infty}v(\tau,r)r^{2}\omega(r)dr
0m2(τ,r)r2ω(r)𝑑r+0m2(0,r)r2ω(r)𝑑r,\displaystyle\quad-\int_{0}^{\infty}m_{2}(\tau,r)r^{2}\omega(r)dr+\int_{0}^{\infty}m_{2}(0,r)r^{2}\omega(r)dr,

which is an ODE in the sense of distributions for S03(τ,)S_{0}^{3}(\tau,\cdot), and because the right hand side is uniformly bounded and continuous, we can consider it in the strong sense (note that boundary terms when integrating the S03S_{0}^{3} term by parts vanish):

1r2r(r2S03(r))=4εv(τ,r)+2m2(τ,r)2m2(0,r),\frac{1}{r^{2}}\partial_{r}(r^{2}S_{0}^{3}(r))=4\varepsilon v(\tau,r)+2m_{2}(\tau,r)-2m_{2}(0,r),

or

S03(r)r=2r30rs2(2εv(τ,s)+m2(τ,s)m2(0,s))𝑑s.\frac{S_{0}^{3}(r)}{r}=\frac{2}{r^{3}}\int_{0}^{r}s^{2}\big{(}2\varepsilon v(\tau,s)+m_{2}(\tau,s)-m_{2}(0,s)\big{)}\,ds.

The energy inequality (3.14) and the Cauchy–Schwarz inequality imply that 2εv(τ,s)2\varepsilon v(\tau,s) and m2(τ,s)m_{2}(\tau,s) are both bounded by E0E_{0} (defined in (4.16)), uniformly in τ,s\tau,s. Hence,

(4.20) |S03(r)r|2r0r(sr)2(2ε|v(τ,s)|+|m2(τ,s)|+|m2(0,s)|)𝑑s2E0.\left|\frac{S_{0}^{3}(r)}{r}\right|\leqslant\frac{2}{r}\int_{0}^{r}\left(\frac{s}{r}\right)^{2}\big{(}2\varepsilon|v(\tau,s)|+|m_{2}(\tau,s)|+|m_{2}(0,s)|\big{)}\,ds\leqslant 2E_{0}.

(See also [6, Proposition 1.9] for a related result.) ∎

Using this lemma, we can derive a scaling relation for the averaged longitudinal structure function S3S^{3}_{\|}, where

(4.21) Sp(τ,r)=0τLdiv2𝕊2D((u(x+rn)u(x))n)p𝑑x𝑑S(n)𝑑μt(u)𝑑t.S_{\|}^{p}(\tau,r)=\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}\big{(}(u(x+rn)-u(x))\cdot n\big{)}^{p}\,dxdS(n)\,d\mu_{t}(u)\,dt.
Lemma 4.3.

Let μt\mu_{t} be a Foiaş–Prodi statistical solution of the Navier–Stokes equations (cf. Definition 3.6). Then

(4.22) |S3(τ,r)r|2E0,\left|\frac{S^{3}_{\|}(\tau,r)}{r}\right|\leqslant 2E_{0},

where C>0C>0 is some constant independent of ε\varepsilon and E0E_{0} is the initial energy,

(4.23) E0Ldiv2(D;U)u(x)L2(D)2𝑑μ0(u)E_{0}\coloneqq\int_{L^{2}_{\operatorname{div}}(D;U)}\left\|u(x)\right\|^{2}_{L^{2}(D)}\,d\mu_{0}(u)
Proof.

Again, we start with the KHM relation (4.2). This time we use the test function σ(h)=ω(|h|)h^h^\sigma(h)=\omega(|h|)\hat{h}\otimes\hat{h} where ωCc()\omega\in C_{c}^{\infty}(\mathbb{R}) is an even function. We have

hk(ω(|h|)h^ih^j)=ω(|h|)h^ih^jh^k+1|h|(δikh^j+δjkh^i2h^ih^jh^k)ω(|h|)=(ω(|h|)2ω(|h|)|h|)h^ih^jh^k+1|h|(δikh^j+δjkh^i)ω(|h|)\begin{split}\partial_{h^{k}}(\omega(|h|)\hat{h}^{i}\hat{h}^{j})&=\omega^{\prime}(|h|)\hat{h}^{i}\hat{h}^{j}\hat{h}^{k}+\frac{1}{|h|}\left(\delta_{ik}\hat{h}^{j}+\delta_{jk}\hat{h}^{i}-2\hat{h}^{i}\hat{h}^{j}\hat{h}^{k}\right)\omega(|h|)\\ &=\left(\omega^{\prime}(|h|)-2\frac{\omega(|h|)}{|h|}\right)\hat{h}^{i}\hat{h}^{j}\hat{h}^{k}+\frac{1}{|h|}\left(\delta_{ik}\hat{h}^{j}+\delta_{jk}\hat{h}^{i}\right)\omega(|h|)\end{split}

Therefore, (4.2) becomes

(4.24) DDU2ξ1h^ξ2h^𝑑ντ,x,x+h2(ξ)𝑑xω(|h|)𝑑h+DDU2ξ1h^ξ2h^𝑑ν0,x,x+h2(ξ)𝑑xω(|h|)𝑑h120τDDU2((ξ2ξ1)h^)3𝑑νt,x,x+h2(ξ)𝑑x(ω(|h|)2|h|1ω(|h|))𝑑h𝑑t+0τDDU2|ξ1ξ2|2(ξ1ξ2)h^𝑑νt,x,x+h2(ξ)𝑑x|h|1ω(|h|)𝑑h𝑑t=εij0τDDU2(ξ1iξ2i)(ξ1jξ2j)𝑑νt,x,x+h2(ξ)𝑑xΔhσij(h)𝑑h𝑑t.\begin{split}&-\int_{D}\int_{D}\!\int_{U^{2}}\!\!\xi_{1}\cdot\hat{h}\,\xi_{2}\cdot\hat{h}\,d\nu_{\tau,x,x+h}^{2}(\xi)\,dx\omega(|h|)\,dh\\ &+\int_{D}\int_{D}\!\int_{U^{2}}\!\!\xi_{1}\cdot\hat{h}\,\xi_{2}\cdot\hat{h}\,\,d\nu_{0,x,x+h}^{2}(\xi)\,dx\omega(|h|)\,dh\\ &-\frac{1}{2}\int_{0}^{\tau}\!\!\int_{D}\int_{D}\!\int_{U^{2}}\big{(}(\xi_{2}-\xi_{1})\cdot\hat{h}\big{)}^{3}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\left(\omega^{\prime}(|h|)-2|h|^{-1}\omega(|h|)\right)\,dh\,dt\\ &+\int_{0}^{\tau}\!\!\int_{D}\int_{D}\!\int_{U^{2}}|\xi_{1}-\xi_{2}|^{2}(\xi_{1}-\xi_{2})\cdot\hat{h}\,d\nu_{t,x,x+h}^{2}(\xi)\,dx|h|^{-1}\omega(|h|)\,dh\,dt\\ =&\ \varepsilon\sum_{ij}\int_{0}^{\tau}\!\!\int_{D}\int_{D}\!\int_{U^{2}}\!\!(\xi_{1}^{i}-\xi_{2}^{i})(\xi_{1}^{j}-\xi_{2}^{j})\,d\nu_{t,x,x+h}^{2}(\xi)\,dx\Delta_{h}\sigma^{ij}(h)\,dh\,dt.\end{split}

Again, in terms of (μt)t>0(\mu_{t})_{t>0}, this means

(4.25) Ldiv2DDu(x)h^u(x+h)h^𝑑xω(|h|)𝑑h𝑑μτ(u)+Ldiv2DDu(x)h^u(x+h)h^𝑑xω(|h|)𝑑h𝑑μ0(u)120τDLdiv2D((u(x+h)u(x))h^)3𝑑x𝑑μt(u)(ω(|h|)2|h|1ω(|h|))𝑑h𝑑t+0τDLdiv2D|u(x)u(x+h)|2(u(x)u(x+h))h^𝑑x𝑑μt(u)|h|1ω(|h|)𝑑h𝑑t=εij0τDLdiv2D(ui(x)ui(x+h))(uj(x)uj(x+h))𝑑x𝑑μt(u)Δhσij(h)𝑑h𝑑t.\begin{split}&\quad-\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!u(x)\cdot\hat{h}\,u(x+h)\cdot\hat{h}\,\,dx\omega(|h|)\,dh\,d\mu_{\tau}(u)\\ &\quad+\int_{L^{2}_{\operatorname{div}}}\int_{D}\int_{D}\!u(x)\cdot\hat{h}\,u(x+h)\cdot\hat{h}\,dx\omega(|h|)\,dh\,d\mu_{0}(u)\\ &\quad-\frac{1}{2}\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}\big{(}(u(x+h)-u(x))\cdot\hat{h}\big{)}^{3}\,dx\,d\mu_{t}(u)\left(\omega^{\prime}(|h|)-2|h|^{-1}\omega(|h|)\right)\,dh\,dt\\ &\quad+\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}|u(x)-u(x+h)|^{2}\big{(}u(x)-u(x+h)\big{)}\cdot\hat{h}\,dx\,d\mu_{t}(u)|h|^{-1}\omega(|h|)\,dh\,dt\\ &=\varepsilon\sum_{ij}\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}\!\big{(}u^{i}(x)-u^{i}(x+h)\big{)}\big{(}u^{j}(x)-u^{j}(x+h)\big{)}\,dx\,d\mu_{t}(u)\Delta_{h}\sigma^{ij}(h)\,dh\,dt.\end{split}

Similar to the computation in (4.19), we have

εij0τDLdiv2D(ui(x)ui(x+h))(uj(x)uj(x+h))𝑑x𝑑μt(u)Δhσij(h)𝑑h𝑑t\displaystyle\quad\varepsilon\sum_{ij}\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}\!\big{(}u^{i}(x)-u^{i}(x+h)\big{)}\big{(}u^{j}(x)-u^{j}(x+h)\big{)}\,dx\,d\mu_{t}(u)\Delta_{h}\sigma^{ij}(h)\,dh\,dt
=2εij0τDLdiv2Dxui(x+h)xuj(x)𝑑x𝑑μt(u)σij(h)𝑑h𝑑t\displaystyle=2\varepsilon\sum_{ij}\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}\!\nabla_{x}u^{i}(x+h)\nabla_{x}u^{j}(x)\,dx\,d\mu_{t}(u)\sigma^{ij}(h)\,dh\,dt
=2ε0τDLdiv2D(xu(x+h)h^)(xu(x)h^)𝑑x𝑑μt(u)ω(|h|)𝑑h𝑑t,\displaystyle=2\varepsilon\int_{0}^{\tau}\int_{D}\int_{L^{2}_{\operatorname{div}}}\!\!\int_{D}\!\big{(}\nabla_{x}u(x+h)\cdot\hat{h}\big{)}\cdot\big{(}\nabla_{x}u(x)\cdot\hat{h}\big{)}\,dx\,d\mu_{t}(u)\omega(|h|)\,dh\,dt,

so equation (4.25) becomes (after switching to polar coordinates)

(4.26) 0Ldiv2𝕊2Du(x)nu(x+rn)n𝑑x𝑑S(n)𝑑μτ(u)r2ω(r)𝑑r+0Ldiv2𝕊2Du(x)nu(x+rn)n𝑑x𝑑S(n)𝑑μ0(u)r2ω(r)𝑑r1200τLdiv2𝕊2D((u(x+rn)u(x))n)3𝑑x𝑑S(n)𝑑μt(u)𝑑t(r2ω(r)2rω(r))𝑑r+00τLdiv2𝕊2D|u(x)u(x+rn)|2(u(x)u(x+rn))n𝑑x𝑑S(n)𝑑μt(u)𝑑trω(r)𝑑r=2ε00τLdiv2𝕊2D(xu(x+rn)n)(xu(x)n)𝑑x𝑑S(n)𝑑μt(u)𝑑tr2ω(r)𝑑r\begin{split}&\quad-\int_{0}^{\infty}\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot n\,u(x+rn)\cdot n\,\,dx\,dS(n)\,d\mu_{\tau}(u)r^{2}\omega(r)\,dr\\ &\quad+\int_{0}^{\infty}\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot n\,u(x+rn)\cdot n\,dx\,dS(n)\,d\mu_{0}(u)r^{2}\omega(r)\,dr\\ &\quad-\frac{1}{2}\int_{0}^{\infty}\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}\big{(}\big{(}u(x+rn)-u(x)\big{)}\cdot n\big{)}^{3}\,dx\,dS(n)\,d\mu_{t}(u)\,dt\left(r^{2}\omega^{\prime}(r)-2r\omega(r)\right)\,dr\\ &\quad+\int_{0}^{\infty}\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}|u(x)-u(x+rn)|^{2}\big{(}u(x)-u(x+rn)\big{)}\cdot n\,dx\,dS(n)\,d\mu_{t}(u)\,dtr\omega(r)\,dr\\ &=2\varepsilon\int_{0}^{\infty}\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}\!\big{(}\nabla_{x}u(x+rn)\cdot n\big{)}\cdot\big{(}\nabla_{x}u(x)\cdot n\big{)}\,dx\,dS(n)\,d\mu_{t}(u)\,dtr^{2}\omega(r)\,dr\end{split}

Denote

m~2(τ,r)\displaystyle\widetilde{m}_{2}(\tau,r) Ldiv2𝕊2Du(x)nu(x+rn)n𝑑x𝑑S(n)𝑑μτ(u)\displaystyle\coloneqq\int_{L^{2}_{\operatorname{div}}}\fint_{\mathbb{S}^{2}}\int_{D}\!u(x)\cdot n\,u(x+rn)\cdot n\,\,dx\,dS(n)\,d\mu_{\tau}(u)
v~(τ,r)\displaystyle\widetilde{v}(\tau,r) 0τLdiv2𝕊2D(xu(x+rn)n)(xu(x)n)𝑑x𝑑S(n)𝑑μt(u)𝑑t.\displaystyle\coloneqq\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}\!\big{(}\nabla_{x}u(x+rn)\cdot n\big{)}\cdot\big{(}\nabla_{x}u(x)\cdot n\big{)}\,dx\,dS(n)\,d\mu_{t}(u)\,dt.

Writing ω(r)2r1ω(r)=r2ddr(ω(r)r2)\omega^{\prime}(r)-2r^{-1}\omega(r)=r^{2}\frac{d}{dr}\big{(}\omega(r)r^{-2}\big{)}, (4.26) becomes

120r4S3(τ,r)r(r2ω(r))dr+0S03(τ,r)rω(r)𝑑r\displaystyle\quad\frac{1}{2}\int_{0}^{\infty}r^{4}S_{\|}^{3}(\tau,r)\partial_{r}\big{(}r^{-2}\omega(r)\big{)}\,dr+\int_{0}^{\infty}S_{0}^{3}(\tau,r)r\omega(r)\,dr
=0m~2(τ,r)r2ω(r)𝑑r+0m~2(0,r)r2ω(r)𝑑r2ε0v~(τ,r)r2ω(r)𝑑r.\displaystyle=-\int_{0}^{\infty}\widetilde{m}_{2}(\tau,r)r^{2}\omega(r)\,dr+\int_{0}^{\infty}\widetilde{m}_{2}(0,r)r^{2}\omega(r)\,dr-2\varepsilon\int_{0}^{\infty}\widetilde{v}(\tau,r)r^{2}\omega(r)\,dr.

Again, we note that due to the estimates from the energy inequality (3.14), S3S^{3}_{\|}, m~\widetilde{m} and v~\widetilde{v} are continuous and bounded quantities in τ\tau and rr. And so we can consider this ODE in the sense of distributions as an ODE in the strong sense,

r(r4S3(τ,r))=2r4(S03(τ,r)r+m~2(τ,r)m~2(0,r)+2εv~(τ,r)),\partial_{r}\big{(}r^{4}S_{\|}^{3}(\tau,r)\big{)}=2r^{4}\left(\frac{S_{0}^{3}(\tau,r)}{r}+\widetilde{m}_{2}(\tau,r)-\widetilde{m}_{2}(0,r)+2\varepsilon\widetilde{v}(\tau,r)\right),

or

(4.27) S3(τ,r)=20rs4r4(S03(τ,s)s+m~2(τ,s)m~2(0,s)+2εv~(τ,s))𝑑s.S_{\|}^{3}(\tau,r)=2\int_{0}^{r}\frac{s^{4}}{r^{4}}\left(\frac{S_{0}^{3}(\tau,s)}{s}+\widetilde{m}_{2}(\tau,s)-\widetilde{m}_{2}(0,s)+2\varepsilon\widetilde{v}(\tau,s)\right)\,ds.

By the energy bound and Cauchy–Schwarz inequality, m~2\widetilde{m}_{2} and εv~\varepsilon\widetilde{v} are uniformly bounded in ε\varepsilon for all s,τ>0s,\tau>0. Moreover from Lemma 4.2, we have that s1S03(s,τ)s^{-1}S_{0}^{3}(s,\tau) is uniformly bounded in ε\varepsilon by 2E02E_{0}. Hence,

(4.28) |S3(τ,r)r|2E0\left|\frac{S_{\|}^{3}(\tau,r)}{r}\right|\leqslant 2E_{0}

for some C>0C>0 independent of ε\varepsilon. ∎

Remark 4.4.

Combining (4.20) and (4.28), we also obtain a uniform bound on r1S3(τ,r)r^{-1}S^{3}_{\perp}(\tau,r), where S3S^{3}_{\perp} is the transversal structure function

(4.29) S3(τ,r)0τLdiv2𝕊2D|δrnu(x)|2δrnu(x)n𝑑x𝑑S(n)𝑑μt(u)𝑑t=S03(τ,r)S3(τ,r),S^{3}_{\perp}(\tau,r)\coloneqq\int_{0}^{\tau}\int_{L^{2}_{\operatorname{div}}}\!\!\fint_{\mathbb{S}^{2}}\int_{D}|\delta^{\perp}_{rn}u(x)|^{2}\delta_{rn}u(x)\cdot n\,dx\,dS(n)\,d\mu_{t}(u)\,dt=S^{3}_{0}(\tau,r)-S^{3}_{\|}(\tau,r),

where

(4.30) δhu(x)=u(x+h)u(x),δhu=(𝐈h^h^)δhu,h^h|h|.\delta_{h}u(x)=u(x+h)-u(x),\quad\delta^{\perp}_{h}u=(\mathbf{I}-\hat{h}\otimes\hat{h})\delta_{h}u,\quad\hat{h}\coloneqq\frac{h}{|h|}.

None of these quantities has a sign and therefore the previously derived bounds do not imply compactness without further assumptions.

4.3. Scaling assumption

In order to pass to the limit ε0\varepsilon\to 0, we need to an additional assumption about the behavior of structure functions. Specifically, we need

Assumption 1 (Weak statistical scaling).

For any ε>0\varepsilon>0, let με\mu^{\varepsilon} be a Foiaş–Prodi statistical solution of the incompressible Navier–Stokes equations. We assume that for r1r\ll 1, the second and third order longitudinal structure functions (4.21) are related by

(4.31) |S2(τ,r)|C|S3(τ,r)|α,\left|S^{2}_{\|}(\tau,r)\right|\leqslant C\left|S^{3}_{\|}(\tau,r)\right|^{\alpha},

where CC is a constant independent of ε\varepsilon and α>0\alpha>0.

Remark 4.5 (Weak statistical scaling).

Assumption 1 is inspired by the following stronger scaling assumption often encountered in turbulence theory: For any p,qp,q with qp1q\geqslant p\geqslant 1, the pp-th and qq-th order longitudinal structure functions (4.21) are related by

(4.32) |Sp(τ,r)|C|Sq(τ,r)|λ(p)λ(q),\left|S^{p}_{\|}(\tau,r)\right|\leqslant C\left|S^{q}_{\|}(\tau,r)\right|^{\frac{\lambda(p)}{\lambda(q)}},

where CC is a constant independent of ε\varepsilon and λ(p)>0\lambda(p)>0 for pp0p\leqslant p_{0} where 3p0{}3\leqslant p_{0}\in\mathbb{R}\cup\{\infty\}. In Kolmogorov’s 1941 (“K41”) theory [42, 41, 43], λ(p)=p\lambda(p)=p. However, this cannot be confirmed with physical experiments [4, 54]. Various physicists therefore suggested intermittency corrections to account for the deviation from Kolmogorov’s original theory, among others, Kolmogorov himself in 1962 [44] in his refined theory of turbulence, Frisch et al. the β\beta-model [33], as well as Novikov and Stewart [51]. Assumption (4.32) can also accommodate the frequently used model by She and Leveque [55] who suggested

(4.33) λ(p)=p9+2(1(23)p/3).\lambda(p)=\frac{p}{9}+2\left(1-\left(\frac{2}{3}\right)^{p/3}\right).
Remark 4.6.

Combining the bound on the third order structure function in Lemma 4.3 with Assumption 1, we obtain

|S2(τ,r)|Crα.\left|S^{2}_{\|}(\tau,r)\right|\leqslant Cr^{\alpha}.

We will combine Assumption 1 with the following lemma, which is Lemma 1 by Drivas [18], translated to the setting of statistical solutions. The proof is given in Appendix C:

Lemma 4.7 (Weak anisotropy).

Let μt\mu_{t} be a statistical solution of the Navier–Stokes equation. Then μ\mu satisfies

(4.34) 30TDLdiv2(D;U)Br(0)(δrnun)2𝑑S(n)𝑑x𝑑μt(u)𝑑t=0TDLdiv2(D;U)Br(0)|δu(x)|2𝑑𝑑x𝑑μt(u)𝑑t.\begin{split}&3\int_{0}^{T}\int_{D}\int_{L^{2}_{\operatorname{div}}(D;U)}\fint_{\partial B_{r}(0)}(\delta_{rn}u\cdot n)^{2}dS(n)\,dx\,d\mu_{t}(u)\,dt\\ &=\int_{0}^{T}\int_{D}\int_{L^{2}_{\operatorname{div}}(D;U)}\fint_{B_{r}(0)}|\delta_{\ell}u(x)|^{2}d\ell\,dx\,d\mu_{t}(u)\,dt.\end{split}

Under Assumption 1, we obtain

(4.35) 0TDLdiv2(D;U)Br(0)|δu(x)|2𝑑𝑑x𝑑μt(u)𝑑tCrα.\int_{0}^{T}\int_{D}\int_{L^{2}_{\operatorname{div}}(D;U)}\fint_{B_{r}(0)}|\delta_{\ell}u(x)|^{2}d\ell\,dx\,d\mu_{t}(u)\,dt\leqslant Cr^{\alpha}.

Using the equivalence theorem 2.2, we can write this as

(4.36) 0TDBr(0)U2|ξ1ξ2|2𝑑νx,x+y2(ξ)𝑑y𝑑x𝑑tCrα.\int_{0}^{T}\int_{D}\fint_{B_{r}(0)}\int_{U^{2}}|\xi_{1}-\xi_{2}|^{2}\,d\nu_{x,x+y}^{2}(\xi)\,dy\,dx\,dt\leqslant Cr^{\alpha}.

and since this is uniform with respect to the viscosity coefficient ε\varepsilon (by the weak scaling assumption), it implies uniform diagonal continuity of the sequence {𝝂}ε>0\{\bm{\nu}\}_{\varepsilon>0}.

4.4. Passage to the limit ε0\varepsilon\to 0

Now we are in a position to prove our main result. We will keep track of the superscript ε\varepsilon again in order to distinguish between the approximating sequence {𝝂ε}ε>0\{\bm{\nu}^{\varepsilon}\}_{\varepsilon>0} and the limiting measure 𝝂\bm{\nu} for ε=0\varepsilon=0.

Theorem 4.8.

Let {𝛎ε}ε>0\{\bm{\nu}^{\varepsilon}\}_{\varepsilon>0} be a sequence of (either Foiaş–Temam or Friedman–Keller) statistical solutions to the Navier–Stokes equations with initial data μ0\mu_{0} with bounded support (cf. (3.1)). Assume that 𝛎ε\bm{\nu}^{\varepsilon} all satisfy Assumption 1. Then, as ε0\varepsilon\to 0, 𝛎ε\bm{\nu}^{\varepsilon} converges (along a subsequence) to a correlation measure 𝛎\bm{\nu} on L2L^{2} with bounded support (cf. (3.17))

(4.37) DkUk|ξ1|2|ξk|2𝑑νt,xk(ξ)𝑑xRk<,\int_{D^{k}}\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{t,x}^{k}(\xi)\,dx\leqslant R^{k}<\infty,

for some 0<R<0<R<\infty, any kk\in\mathbb{N} and that satisfies the “inviscid Friedman–Keller system”:

(4.38) 0TDkUk(ξ1ξk):φt(t,x)dνt,xk(ξ)dxdt+DkUk(ξ1ξk):φ(0,x)dν0,xk(ξ)dx+i=1k0TDkUk(ξ1(ξiξi)ξk):xiφ(t,x)dνt,xk(ξ)dxdt=0\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\frac{\partial\varphi}{\partial t}(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt+\int_{D^{k}}\!\!\int_{U^{k}}\!\!(\xi_{1}\otimes\cdots\otimes\xi_{k}):\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx\\ +\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\!\!\int_{U^{k}}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt=0

for all kk\in\mathbb{N}, for all φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}) with divxiφ(x)=0\operatorname{div}_{x_{i}}\varphi(x)=0, a.e. xDkx\in D^{k} for all i=1,,ki=1,\dots,k and (corresponding to the divergence constraint)

(4.39) DkUkξ1ξα+1(ξ+1)αk(ξk)𝑑νt,xk(ξ)x1,,xψ(x)𝑑x=0,\int_{D^{k}}\int_{U^{k}}\xi_{1}\otimes\dots\otimes\xi_{\ell}\otimes\alpha_{\ell+1}(\xi_{\ell+1})\otimes\dots\otimes\alpha_{k}(\xi_{k})\,d\nu_{t,x}^{k}(\xi)\cdot\nabla_{x_{1},\dots,x_{\ell}}\psi(x)\,dx=0,

where x1,,x=(x1,,x)\nabla_{x_{1},\dots,x_{\ell}}=(\nabla_{x_{1}},\dots,\nabla_{x_{\ell}})^{\top}, 1k1\leqslant\ell\leqslant k\in\mathbb{N}, for all ψH1(Dk;Uk)\psi\in H^{1}(D^{k};U^{k-\ell}), αjC(U;U)\alpha_{j}\in C(U;U), αj(v)C(1+|v|2)\alpha_{j}(v)\leqslant C(1+|v|^{2}) and j=1,,kj=1,\dots,k.

Proof.

From the condition on the initially bounded support (3.1) and the energy inequality (3.4), we obtain that the sequence 𝝂ε\bm{\nu}^{\varepsilon} satisfies (2.9) for p=2p=2 uniformly in ε>0\varepsilon>0. The reasoning of Subsection 4.14.2 and 4.3 resulting in (4.36) imply that 𝝂ε\bm{\nu}^{\varepsilon} is uniformly diagonal continuous as in (2.10). Hence, using Theorem 2.3, we obtain, up to subsequence, the existence of a limiting correlation measure 𝝂2([0,T),D;U)\bm{\nu}\in\mathcal{L}^{2}([0,T),D;U). So it remains to check whether 𝝂\bm{\nu} satisfies the equations (4.38) and (4.39). We note that the functions

(4.40) g1(t,x,ξ)(ξ1ξk):φt(t,x),g2(t,x,ξ)i=1k(ξ1ξk):Δxiφ(t,x),g3(t,x,ξ)i=1k(ξ1(ξiξi)ξk):xiφ(t,x)\begin{split}g_{1}(t,x,\xi)&\coloneqq(\xi_{1}\otimes\cdots\otimes\xi_{k}):\frac{\partial\varphi}{\partial t}(t,x),\\ g_{2}(t,x,\xi)&\coloneqq\sum_{i=1}^{k}(\xi_{1}\otimes\cdots\otimes\xi_{k}):\Delta_{x_{i}}\varphi(t,x),\\ g_{3}(t,x,\xi)&\coloneqq\sum_{i=1}^{k}(\xi_{1}\otimes\cdots\otimes(\xi_{i}\otimes\xi_{i})\otimes\cdots\xi_{k}):\nabla_{x_{i}}\varphi(t,x)\end{split}

for φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}), 1k1\leqslant\ell\leqslant k\in\mathbb{N}, are all functions in 1k,p([0,T],D;U)\mathcal{H}^{k,p}_{1}([0,T],D;U). Hence, we can pass to the limit in all the terms in the Friedman–Keller system (3.2). The term that is multiplied by ε\varepsilon vanishes because it is a uniformly bounded in ε>0\varepsilon>0 quantity that is multiplied by ε\varepsilon. For the divergence constraint (4.39), we note that the function

g4(t,x,ξ)θ(t)ξ1ξα+1(ξ+1)αk(ξk)x1,,xψ(x),g_{4}(t,x,\xi)\coloneqq\theta(t)\xi_{1}\otimes\dots\otimes\xi_{\ell}\otimes\alpha_{\ell+1}(\xi_{\ell+1})\otimes\dots\otimes\alpha_{k}(\xi_{k})\cdot\nabla_{x_{1},\dots,x_{\ell}}\psi(x),

lies in 1k,p([0,T],D;U)\mathcal{H}^{k,p}_{1}([0,T],D;U) for any θCc((0,T))\theta\in C^{\infty}_{c}((0,T)), and ψH1(Dk;Uk)\psi\in H^{1}(D^{k};U^{k-\ell}), and αjC(U;U)\alpha_{j}\in C(U;U) with |αj(v)|C(1+|v|2)|\alpha_{j}(v)|\leqslant C(1+|v|^{2}). Passing ε0\varepsilon\to 0 in νε,k,g4\langle\nu^{\varepsilon,k},\,g_{4}\rangle and using that νε,k\nu^{\varepsilon,k} satisfy the divergence constraint (3.3), we can conclude, by the arbitrariness of θ\theta, that (4.39) holds for a.e. t[0,T]t\in[0,T]. ∎

Remark 4.9.

By the equivalence theorem 2.2, we know that the limiting correlation measure 𝝂\bm{\nu} corresponds to a parametrized measure μ=(μt)t>0:[0,T)𝒫(Ldiv2(D;U))\mu=(\mu_{t})_{t>0}:[0,T)\to\mathscr{P}(L^{2}_{\operatorname{div}}(D;U)) that satisfies

(4.41) L2(D)Φ(u)𝑑μt(u)=L2(D)Φ(u)𝑑μ0(u)+0tL2(D)D(u(x)u(x)):xΦ(u)(x)dxdμs(u)ds\int_{L^{2}(D)}\Phi(u)\,d\mu_{t}(u)=\int_{L^{2}(D)}\Phi(u)\,d\mu_{0}(u)+\int_{0}^{t}\int_{L^{2}(D)}\int_{D}(u(x)\otimes u(x)):\nabla_{x}\Phi^{\prime}(u)(x)\,dx\,d\mu_{s}(u)\,ds

for all cylindrical test functions Φ𝒯cyl0\Phi\in\mathscr{T}_{\text{cyl}}^{0} that satisfy gjC2(D)g_{j}\in C^{2}(D), and the energy inequality

(4.42) L2(D)uL2(D)2𝑑μt(u)L2(D)uL2(D)2𝑑μ0(u),for all t[0,T].\int_{L^{2}(D)}\|u\|_{L^{2}(D)}^{2}\,d\mu_{t}(u)\leqslant\int_{L^{2}(D)}\|u\|_{L^{2}(D)}^{2}\,d\mu_{0}(u),\quad\text{for all }t\in[0,T].

The proof of this fact follows along the lines of the proof of Theorem 3.9 while ignoring the terms involving ε\varepsilon and not attempting to recover B(u)B(u) as it may be unbounded.

5. Discussion

It is well-known that many incompressible fluid flows of interest are characterized by very-high Reynolds number. Hence, a precise characterization of the vanishing viscosity (ε0\varepsilon\to 0) limit of the Navier-Stokes equations (1.1) is of great interest. Formally, one would expect that the vanishing viscosity limit of Navier-Stokes equations is related to the incompressible Euler equations. However as mentioned in the introduction, rigorous results in this direction are only available in two space dimensions, even in the case of periodic boundary conditions. The key aim of this article was to investigate the vanishing viscosity limit of the Navier-Stokes equations, including in three space dimensions.

It is well known that fluid flows at high Reynolds numbers are characterized by turbulence, loosely speaking, marked by the presence of energy containing eddies at smaller and smaller scales. This phenomenon is clearly linked to the lack of compactness in the Leray-Hopf Navier-Stokes solutions as well as their possible instabilities/non-uniqueness.

Hence, one needs to make further assumptions on the Leray-Hopf solutions that can yield additional information and facilitate passage to the limit. One avenue for making such assumptions, which are realistic and possibly observed in experiments, comes from physical theories of turbulence. In particular, Kolmogorov’s well-known K41 theory is based on several verifiable assumptions on the underlying fluid flow and results in a precise characterization of quantities such as structure functions and energy spectra.

In [12], Chen and Glimm relate the K41 energy spectra to compactness results on the Leray-Hopf solutions, in appropriate Sobolev and Hölder spaces. Consequently, under the assumption of the K41 energy spectrum, the authors prove that the underlying Leray-Hopf solutions converge to weak solutions of the incompressible Euler equations as ε0\varepsilon\to 0. However, Kolmogorov’s derivation of the decay of energy spectra is based on a probabilistic charectization of the underlying fluid flow. In particular, assumptions such as (statistical) homogeneity, isotropy and scaling, which form the foundation of Kolmogorov’s theory, are too stringent if imposed at the deterministic level, as done in [12]. Moreover, it is now well-established that the strong scaling assumptions of Kolmogorov might not hold in real fluid flows and intermittent corrections are necessary. Hence, the applicability of the assumptions and results of [12] can be questioned from this perspective.

Nevertheless, the connection with Kolmogorov’s theories of turbulence and their variants forms the basis of our work. We start with the realization that a probabilistic description of the solutions of Navier-Stokes equations is necessary to relate physical theories of turbulence to rigorous mathematical statements. To this end, we focus on statistical solutions of Navier-Stokes equations. Two possible frameworks of such statistical solutions are available, namely the Foiaş-Prodi statistical solutions (see Definition 3.6) and the Friedman-Keller statistical solutions (see Definition 3.1), which is based on the concept of correlation measures of [24]. We prove that both these solution concepts are equivalent as long as a statistical version of the energy inequality holds. This also allows us to prove the existence of Friedmann-Keller statistical solutions of the incompressible Navier-Stokes equations.

Then, we investigated the vanishing viscosity limit of the statistical solutions of the incompressible Navier-Stokes equations. To this end, we derived a suitable statistical version of the well-known Kármán-Howarth-Monin relation and used it to prove precise rates for the asymptotic decay of a averaged third-order structure function in Lemma 4.2. However, these estimates do not suffice to pass to the ε0\varepsilon\to 0 limit. To this end, we assumed a weak statistical scaling of the Navier-Stokes statistical solutions (see Assumption 1). This assumption is a weaker version of Kolmogorov’s scaling assumptions in his K41 theory. Moreover, it is consistent with and incorporates different variants of scaling that are proposed in the physics literature to explain intermittent corrections to Kolmogorov’s theory. Under this assumption, we proved a weak anisotropy result and invoked compactness results of [25] to rigorously prove that the statistical solutions of the Navier-Stokes equations converge, in a suitable sense, to a statistical solution of the incompressible Euler equations. Thus, we were able to characterize the vanishing viscosity limit of the Navier-Stokes equations in a relevant regime.

At this juncture, it is essential to point that that no assumption, other than weak statistical scaling, is made in our results and all other estimates are derived rigorously. This should be contrasted with the results of [12] where the authors directly assume a decay of the energy spectrum for the weak solutions of the Navier-Stokes equations. It is currently unclear if one can relax the weak statistical scaling assumption or even if it holds for all incompressible fluid flows. Experimental evidence strongly supports that this assumption is verified in practice, see e.g., [4, 55, 54].

To the best of our knowledge, the only rigorous study of the vanishing viscosity limit of the (Foiaş-Prodi) statistical solutions was carried out by Chae in [11] where he proved that these statistical solutions converge to a measure-valued solution of the incompressible Euler equations. In contrast, we prove convergence to statistical solutions of the incompressible Euler equations and recall that statistical solutions are much more informative than measure-valued solutions as they also incorporate knowledge of all multi-point correlations.

Finally, our characterization of the vanishing viscosity limit can be viewed in connection to recent results in [46] where the authors proved convergence of numerical spectral viscosity approximations to the statistical solutions of the Euler equations under very similar weak scaling assumptions.

Appendix A Equivalence of different definitions of statistical solutions for the incompressible Navier–Stokes equations

This appendix is devoted to the proof of Theorems 3.8 and 3.9. For convenience, we restate the result:

Theorem A.1 (Foiaş–Prodi statistical solutions are Friedman–Keller solutions).

Let μ\mu be a Foiaş–Prodi statistical solution such that the initial condition μ0\mu_{0} has bounded support, supp(μ0)BLdiv2(D;U)\text{supp}(\mu_{0})\subset B\subset L^{2}_{\operatorname{div}}(D;U), B={uL2(D;U):uL2(D)R}B=\{u\in L^{2}(D;U)\,:\,\|u\|_{L^{2}(D)}\leqslant R\} for some 0<R0<R\in\mathbb{R} large enough. Then μ\mu is a Friedmann–Keller statistical solution (cf. Definition 3.1).

Proof.

It is shown in [27, Theorem 2, Section 3] that Foiaş–Prodi statistical solutions with initial measure μ0\mu_{0} having bounded support in BLdiv2(D;U)(R){uLdiv2(D;U):uL2(D)R}B_{L^{2}_{\operatorname{div}}(D;U)}(R)\coloneqq\{u\in L^{2}_{\operatorname{div}}(D;U):\|u\|_{L^{2}(D)}\leqslant R\} have bounded support for all times, i.e., supp(μt)BLdiv2(D;U)(R)\text{supp}(\mu_{t})\subset B_{L^{2}_{\operatorname{div}}(D;U)}(R). Therefore, we can assume that μt\mu_{t} has uniformly bounded support. This implies in particular that μt\mu_{t} have bounded moments:

(A.1) Ldiv2(D;U)uL2(D)2k𝑑μt(u)R2kfor a.e. t[0,T].\int_{L^{2}_{\operatorname{div}}(D;U)}\|u\|_{L^{2}(D)}^{2k}\,d\mu_{t}(u)\leqslant R^{2k}\quad\text{for a.e. }t\in[0,T].

Moreover, by [27, Lemma 5, Section 3], we have that statistical solutions of Navier–Stokes satisfy

(A.2) 0TLdiv2(D;U)[tΦ(t,u)+a(u,uΦ(t,u))+b(u,u,uΦ(t,u))]𝑑μt(u)𝑑t=Ldiv2(D;U)Φ(0,u)𝑑μ0(u)\int_{0}^{T}\int_{L^{2}_{\operatorname{div}}(D;U)}\left[-\partial_{t}\Phi(t,u)+a(u,\partial_{u}\Phi(t,u))+b(u,u,\partial_{u}\Phi(t,u))\right]\,d\mu_{t}(u)\,dt\\ =\int_{L^{2}_{\operatorname{div}}(D;U)}\Phi(0,u)\,d\mu_{0}(u)

for any test function Φ(t,u)\Phi(t,u) that is Fréchet differentiable on [0,T]×Ldiv2(D;U)[0,T]\times L^{2}_{\operatorname{div}}(D;U) with Φ(t,)=0\Phi(t,\cdot)=0 near t=Tt=T and |uΦ(t,u)|C|\partial_{u}\Phi(t,u)|\leqslant C and |tΦ(t,u)|C1+C2uL2(D)|\partial_{t}\Phi(t,u)|\leqslant C_{1}+C_{2}\|u\|_{L^{2}(D)} for all uu and tt and some constants C,C1,C2C,C_{1},C_{2} (equation (3.13I3.13_{\mathrm{I}}) and condition (3.8) in [27]). Therefore, we can choose test functions

Φ(u)=q((u,φ1),,(u,φk))θ(t)\Phi(u)=q((u,\varphi_{1}),\dots,(u,\varphi_{k}))\theta(t)

for qq a polynomial on k\mathbb{R}^{k} and φjHdiv1(D;U)Cc2(D,U)\varphi_{j}\in H^{1}_{\operatorname{div}}(D;U)\cap C^{2}_{c}(D,U) and θCc1([0,T))\theta\in C^{1}_{c}([0,T)), j=1,,kj=1,\dots,k in (A.2), so that we get

0TLdiv2(D;U)θ(t)q((u,φ1),,(u,φk))dμt(u)dt+0TLdiv2(D;U)[θ(t)i=1kiq((u,φ1),,(u,φk))[a(u,φi)+b(u,u,φi)]]𝑑μt(u)𝑑t=Ldiv2(D;U)θ(0)q((u,φ1),,(u,φk))𝑑μ0(u),\int_{0}^{T}\!\!\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!-\theta^{\prime}(t)q((u,\varphi_{1}),\dots,(u,\varphi_{k}))d\mu_{t}(u)dt\\ +\int_{0}^{T}\!\!\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!\Big{[}\theta(t)\sum_{i=1}^{k}\partial_{i}q((u,\varphi_{1}),\dots,(u,\varphi_{k}))\left[a(u,\varphi_{i})+b(u,u,\varphi_{i})\right]\Big{]}\,d\mu_{t}(u)\,dt\\ =\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!\!\!\!\theta(0)q((u,\varphi_{1}),\dots,(u,\varphi_{k}))\,d\mu_{0}(u),

Note that we can integrate by parts in the terms involving a(u,φi)a(u,\varphi_{i}) and b(u,u,φi)b(u,u,\varphi_{i}), i=1,,ki=1,\dots,k, so that all the derivatives are on the test functions φi\varphi_{i}, i=1,,ki=1,\dots,k and θ\theta:

0=\displaystyle 0= 0TLdiv2(D;U)θ(t)q((u,φ1),,(u,φk))𝑑μt(u)𝑑t+Ldiv2(D;U)θ(0)q((u,φ1),,(u,φk))𝑑μ0(u)\displaystyle\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!\theta^{\prime}(t)q((u,\varphi_{1}),\dots,(u,\varphi_{k}))\,d\mu_{t}(u)\,dt+\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!\theta(0)q((u,\varphi_{1}),\dots,(u,\varphi_{k}))\,d\mu_{0}(u)
+i=1k0TLdiv2(D;U)θ(t)iq((u,φ1),,(u,φk))\displaystyle\quad+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!\theta(t)\partial_{i}q((u,\varphi_{1}),\dots,(u,\varphi_{k}))
×D[εu(xi)Δxiφi(xi)+(u(xi)u(xi)):xiφi(xi)]dxidμt(u)dt\displaystyle\hphantom{\quad+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\!\!}\times\int_{D}\left[\varepsilon u(x_{i})\Delta_{x_{i}}\varphi_{i}(x_{i})+(u(x_{i})\otimes u(x_{i})):\nabla_{x_{i}}\varphi_{i}(x_{i})\right]\,dx_{i}\,d\mu_{t}(u)\,dt

Now take q(s1,,sk)=s1skq(s_{1},\dots,s_{k})=s_{1}\cdots s_{k}, so that the last identity becomes (denote dx=dx1dxk\,dx=\,dx_{1}\dots\,dx_{k})

0=0TLdiv2(D;U)Dkθ(t)u(x1)φ1(x1)u(xk)φk(xk)𝑑x𝑑μt(u)𝑑t+i=1k0TLdiv2(D;U)Dkθ(t)(u(x1)φ1(x1)εu(xi)Δxiφi(xi)u(xk)φk(xk))𝑑x𝑑μt(u)𝑑t+i=1k0TLdiv2(D;U)Dkθ(t)(u(x1)φ1(x1)(u(xi)u(xi)):xiφi(xi)u(xk)φk(xk))dxdμt(u)dt+Ldiv2(D;U)Dkθ(0)u(x1)φ1(x1)u(xk)φk(xk)𝑑x𝑑μ0(u),0=\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}\theta^{\prime}(t)u(x_{1})\cdot\varphi_{1}(x_{1})\cdots u(x_{k})\cdot\varphi_{k}(x_{k})\,dx\,d\mu_{t}(u)\,dt\\ +\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}\theta(t)\big{(}u(x_{1})\cdot\varphi_{1}(x_{1})\cdots\varepsilon u(x_{i})\cdot\Delta_{x_{i}}\varphi_{i}(x_{i})\cdots u(x_{k})\cdot\varphi_{k}(x_{k})\big{)}\,dx\,d\mu_{t}(u)\,dt\\ +\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}\theta(t)\big{(}u(x_{1})\cdot\varphi_{1}(x_{1})\cdots(u(x_{i})\otimes u(x_{i})):\nabla_{x_{i}}\varphi_{i}(x_{i})\cdots u(x_{k})\cdot\varphi_{k}(x_{k})\big{)}\,dx\,d\mu_{t}(u)\,dt\\ +\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}\theta(0)u(x_{1})\cdot\varphi_{1}(x_{1})\cdots u(x_{k})\cdot\varphi_{k}(x_{k})\,dx\,d\mu_{0}(u),

which is, denoting x=(x1,,xk)x=(x_{1},\dots,x_{k}) and φ(t,x)θ(t)φ1(x1)φk(xk)\varphi(t,x)\coloneqq\theta(t)\varphi_{1}(x_{1})\otimes\dots\otimes\varphi_{k}(x_{k}), equivalent to

0=\displaystyle 0= 0TLdiv2(D;U)Dk(u(x1)u(xk)):tφ(t,x)dxdμt(u)dt\displaystyle\ \int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}(u(x_{1})\otimes\dots\otimes u(x_{k})):\partial_{t}\varphi(t,x)\,dx\,d\mu_{t}(u)\,dt
+εi=1k0TLdiv2(D;U)Dk(u(x1)u(xk)):Δxiφ(t,x)dxdμt(u)dt\displaystyle+\varepsilon\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}(u(x_{1})\otimes\dots\otimes u(x_{k})):\Delta_{x_{i}}\varphi(t,x)\,dx\,d\mu_{t}(u)\,dt
+i=1k0TLdiv2(D;U)Dk(u(x1)(u(xi)u(xi))u(xk)):xiφ(t,x)dxdμt(u)dt\displaystyle+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}(u(x_{1})\otimes\dots\otimes(u(x_{i})\otimes u(x_{i}))\otimes\dots\otimes u(x_{k})):\nabla_{x_{i}}\varphi(t,x)\,dx\,d\mu_{t}(u)\,dt
+Ldiv2(D;U)Dk(u(x1)u(xk)):φ(0,x)dxdμ0(u),\displaystyle+\int_{L^{2}_{\operatorname{div}}(D;U)}\int_{D^{k}}(u(x_{1})\otimes\dots\otimes u(x_{k})):\varphi(0,x)\,dx\,d\mu_{0}(u),

Since we assume that the support of μ0\mu_{0} is bounded (and therefore also the support of μt\mu_{t} for almost all t[0,T]t\in[0,T]), we can use a density argument to conclude that the above identity holds for all φL2([0,T];(Hdiv1(D;U))k)Cc2([0,T]×Dk;Uk)\varphi\in L^{2}([0,T];(H^{1}_{\operatorname{div}}(D;U))^{k})\cap C^{2}_{c}([0,T]\times D^{k};U^{k}), that is, all φCc2([0,T]×Dk;Uk)\varphi\in C^{2}_{c}([0,T]\times D^{k};U^{k}) with divxiφ=0\operatorname{div}_{x_{i}}\varphi=0 for all i=1,,ki=1,\dots,k. Observe next that all the terms in the above identity have the form required to apply Theorem 2.2 (recall that μt\mu_{t} has bounded support). Therefore, by the same theorem, there exists a unique correlation measure 𝝂=(𝝂t)0tT=((νt1,νt2,))0tT\bm{\nu}=(\bm{\nu}_{t})_{0\leqslant t\leqslant T}=\big{(}(\nu^{1}_{t},\nu^{2}_{t},\dots)\big{)}_{0\leqslant t\leqslant T} corresponding to {μt}\{\mu_{t}\} satisfying

0=\displaystyle 0= 0TDkUk(ξ1ξk):tφ(t,x)dνt,xk(ξ)dxdt\displaystyle\ \int_{0}^{T}\!\!\int_{D^{k}}\int_{U^{k}}(\xi_{1}\otimes\dots\otimes\xi_{k}):\partial_{t}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt
+εi=1k0TDkUk(ξ1ξk):Δxiφ(t,x)dνt,xk(ξ)dxdt\displaystyle+\varepsilon\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\int_{U^{k}}(\xi_{1}\otimes\dots\otimes\xi_{k}):\Delta_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt
+i=1k0TDkUk(ξ1(ξiξi)ξk):xiφ(t,x)dνt,xk(ξ)dxdt\displaystyle+\sum_{i=1}^{k}\int_{0}^{T}\!\!\int_{D^{k}}\int_{U^{k}}(\xi_{1}\otimes\dots\otimes(\xi_{i}\otimes\xi_{i})\otimes\dots\otimes\xi_{k}):\nabla_{x_{i}}\varphi(t,x)\,d\nu_{t,x}^{k}(\xi)\,dx\,dt
+DkUk(ξ1ξk):φ(0,x)dν0,xk(ξ)dx,\displaystyle+\int_{D^{k}}\int_{U^{k}}(\xi_{1}\otimes\dots\otimes\xi_{k}):\varphi(0,x)\,d\nu_{0,x}^{k}(\xi)\,dx,

which is (3.2). By previous arguments, each measure μt\mu_{t} has bounded support (and in particular bounded moments), so we can take any nonnegative, nondecreasing polynomial ψ(s)=pK(s)\psi(s)=p_{K}(s), of degree KK, for KK\in\mathbb{N} in the strengthened energy inequality (3.14), so that we get

(A.3) L2(D)pK(uL2(D)2)𝑑μt(u)+2ε0tL2(D)pK(uL2(D)2)|u|H1(D)2𝑑μs(u)𝑑sL2(D)pK(uL2(D)2)𝑑μ0(u)for a.e. t[0,T].\int_{L^{2}(D)}p_{K}(\|u\|_{L^{2}(D)}^{2})\,d\mu_{t}(u)+2\varepsilon\int_{0}^{t}\int_{L^{2}(D)}p_{K}^{\prime}(\|u\|_{L^{2}(D)}^{2})|u|_{H^{1}(D)}^{2}\,d\mu_{s}(u)\,ds\\ \leqslant\int_{L^{2}(D)}p_{K}(\|u\|_{L^{2}(D)}^{2})\,d\mu_{0}(u)\qquad\text{for a.e. }t\in[0,T].

Thanks to the bounded support of μt\mu_{t}, all the moment terms have the form required to apply Theorem 2.2, moreover, with Lemma B.3, we can rewrite the gradient term such that we obtain for a.e. t[0,T]t\in[0,T], the energy inequality (3.4). Indeed, we have with Theorem 2.2

L2(D)pK(uL2(D)2)𝑑μt(u)\displaystyle\int_{L^{2}(D)}p_{K}(\|u\|_{L^{2}(D)}^{2})\,d\mu_{t}(u) =k=0KakL2(D)uL2(D)2k𝑑μt(u)\displaystyle=\sum_{k=0}^{K}a_{k}\int_{L^{2}(D)}\|u\|_{L^{2}(D)}^{2k}\,d\mu_{t}(u)
=k=0KakL2(D)Dk|u(x1)|2|u(xk)|2𝑑x𝑑μt(u)\displaystyle=\sum_{k=0}^{K}a_{k}\int_{L^{2}(D)}\int_{D^{k}}|u(x_{1})|^{2}\dots|u(x_{k})|^{2}dx\,d\mu_{t}(u)
=k=0KakDkUk|ξ1|2|ξk|2𝑑νt,xk(ξ)𝑑x.\displaystyle=\sum_{k=0}^{K}a_{k}\int_{D^{k}}\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}d\nu_{t,x}^{k}(\xi)dx.

Similarly, with Lemma B.3, (via Lemma B.1)

0tL2(D)pK(uL2(D)2)|u|H1(D)2𝑑μs(u)𝑑s\displaystyle\int_{0}^{t}\int_{L^{2}(D)}p_{K}^{\prime}\big{(}\|u\|_{L^{2}(D)}^{2}\big{)}|u|_{H^{1}(D)}^{2}\,d\mu_{s}(u)\,ds
=k=0Kakk0tL2(D)uL2(D)2(k1)|u|H1(D)2𝑑μs(u)𝑑s\displaystyle\qquad=\sum_{k=0}^{K}a_{k}k\int_{0}^{t}\int_{L^{2}(D)}\|u\|_{L^{2}(D)}^{2(k-1)}|u|_{H^{1}(D)}^{2}\,d\mu_{s}(u)\,ds
=k=0Kakk0tL2(D)Dk|u(x1)|2|u(xk1)|2|u(xk)|2𝑑x𝑑μs(u)𝑑s\displaystyle\qquad=\sum_{k=0}^{K}a_{k}k\int_{0}^{t}\int_{L^{2}(D)}\int_{D^{k}}|u(x_{1})|^{2}\dots|u(x_{k-1})|^{2}|\nabla u(x_{k})|^{2}dx\,d\mu_{s}(u)\,ds
=k=0Kaki=1k0tL2(D)Dk|u(x1)|2|u(xi)|2|u(xk)|2𝑑x𝑑μs(u)𝑑s\displaystyle\qquad=\sum_{k=0}^{K}a_{k}\sum_{i=1}^{k}\int_{0}^{t}\int_{L^{2}(D)}\int_{D^{k}}|u(x_{1})|^{2}\dots|\nabla u(x_{i})|^{2}\dots|u(x_{k})|^{2}dx\,d\mu_{s}(u)\,ds
=k=0Kaki=1kj=1dlimh01h20tDkUk+1|ξ1|2|ξiξk+1|2|ξk|2𝑑νt,(x,xi+h𝐞j)k+1(ξ,ξk+1)𝑑x𝑑s\displaystyle\qquad=\sum_{k=0}^{K}a_{k}\sum_{i=1}^{k}\sum_{j=1}^{d}\lim_{h\to 0}\frac{1}{h^{2}}\int_{0}^{t}\int_{D^{k}}\int_{U^{k+1}}|\xi_{1}|^{2}\dots|\xi_{i}-\xi_{k+1}|^{2}\dots|\xi^{k}|^{2}d\nu^{k+1}_{t,(x,x_{i}+h\mathbf{e}_{j})}(\xi,\xi_{k+1})dx\,ds

and a similar computation for the term involving μ0\mu_{0} yields (3.4). The bounded support of μ0\mu_{0} implies, using the equivalence theorem [24, Theorem 2.7],

DkUk|ξ1|2|ξk|2𝑑ν0,xk(ξ)𝑑x<\int_{D^{k}}\!\!\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{0,x}^{k}(\xi)\,dx<\infty

for any kk\in\mathbb{N} and thus the boundedness of all terms in (3.4). It remains to show that each νk\nu^{k} satisfies (3.3). We let φCc1(Dk;Uk)\varphi\in C^{1}_{c}(D^{k};U^{k}), 1k1\leqslant\ell\leqslant k, αjC(U;U)\alpha_{j}\in C(U;U), with αj(v)C(1+|v|2)\alpha_{j}(v)\leqslant C(1+|v|^{2}), j=,,kj=\ell,\dots,k and compute using the equivalence theorem 2.2,

DkUkξ1ξα+1(ξ+1)αk(ξk)𝑑νt,xk(ξ)x1,,xφ(x)𝑑x=L2(D)Dku(x1)u(x)α+1(u(x+1))αk(u(xk))x1,,xφ(x)dxdμt(u)=0,\begin{split}&\int_{D^{k}}\int_{U^{k}}\xi_{1}\otimes\dots\otimes\xi_{\ell}\otimes\alpha_{\ell+1}(\xi_{\ell+1})\otimes\dots\otimes\alpha_{k}(\xi_{k})\,d\nu_{t,x}^{k}(\xi)\cdot\nabla_{x_{1},\dots,x_{\ell}}\varphi(x)\,dx\\ &\qquad=\int_{L^{2}(D)}\int_{D^{k}}\!\!\!\begin{aligned} &u(x_{1})\otimes\dots\otimes u(x_{\ell})\otimes\alpha_{\ell+1}(u(x_{\ell+1}))\otimes\dots\otimes\alpha_{k}(u(x_{k}))\\ &\cdot\nabla_{x_{1},\dots,x_{\ell}}\varphi(x)\,dx\,d\mu_{t}(u)\end{aligned}\\ &\qquad=0,\end{split}

since μt\mu_{t} is supported on divergence free functions. This concludes the proof. ∎

Let us show the reverse direction now, that is, that any correlation measure that solves the Friedman–Keller system and satisfies in addition an energy inequality, is a statistical solution of the Navier–Stokes equations in the sense of Foiaş–Prodi.

Theorem A.2 (Friedman–Keller solutions are Foiaş–Prodi solutions).

Let 𝛎=(𝛎t)0tT\bm{\nu}=(\bm{\nu}_{t})_{0\leqslant t\leqslant T} be a Friedman–Keller statistical solution of Navier–Stokes (cf. Definition 3.1) with bounded support, i.e.,

(A.4) DkUk|ξ1|2|ξk|2𝑑νt,xk(ξ)𝑑xRk<,\int_{D^{k}}\int_{U^{k}}|\xi_{1}|^{2}\dots|\xi_{k}|^{2}\,d\nu_{t,x}^{k}(\xi)\,dx\leqslant R^{k}<\infty,

for some 0<R<0<R<\infty, every kk\in\mathbb{N}, and almost every t[0,T]t\in[0,T]. Then 𝛎\bm{\nu} corresponds to a probability measure μ=(μt)0tT\mu=(\mu_{t})_{0\leqslant t\leqslant T} on a bounded set of Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) which is a Foiaş–Prodi statistical solution of the Navier–Stokes equations (cf. Definition 3.6).

Proof.

From the equivalence theorem 2.2, we obtain that (𝝂t)0tT(\bm{\nu}_{t})_{0\leqslant t\leq T} corresponds to a family of measures (μt)0tT𝒫(L2(D;U))(\mu_{t})_{0\leqslant t\leqslant T}\subset\mathscr{P}(L^{2}(D;U)) with bounded support, that is supp(μt){uL2(D;U):uL2(D;U)R}\text{supp}(\mu_{t})\subset\{u\in L^{2}(D;U)\,:\,\|u\|_{L^{2}(D;U)}\leqslant R\} for almost every t[0,T]t\in[0,T]. Property (a) of Definition 3.6 – the fact that (3.12) is measurable for all φCb(L2)\varphi\in C_{b}(L^{2}) – follows from a monotone class argument, which we include here. By property (i) of Theorem 2.2, the function

tL2(D;U)Lf(u)𝑑μt(u)=L2(D;U)Df(x,u(x))𝑑x𝑑μt(u)t\mapsto\int_{L^{2}(D;U)}L_{f}(u)\,d\mu_{t}(u)=\int_{L^{2}(D;U)}\int_{D}f(x,u(x))\,dx\,d\mu_{t}(u)

is measurable for every fL1(Dk,C0(Uk))f\in L^{1}(D^{k},C_{0}(U^{k})). Let 𝐌\mathbf{M} be the collection of sets

𝐌={E(L2(D;U)) such that tL2(D;U)𝟙E(u)𝑑μt(u) is measurable}.\mathbf{M}=\left\{E\in\mathscr{B}(L^{2}(D;U))\text{ such that }t\mapsto\int_{L^{2}(D;U)}\mathbbm{1}_{E}(u)\,d\mu_{t}(u)\text{ is measurable}\right\}.

By the monotone convergence theorem, 𝐌\mathbf{M} is a monotone class, that is, it is closed under (countable) unions of increasing sequences of sets and intersections of decreasing sequences of sets. By the same argument as in the proof of [24, Proposition 2.12], 𝐌\mathbf{M} contains the collection of cylinder sets Cyl(L2)={all cylinder sets on L2(D;U)}\mathrm{Cyl}(L^{2})=\{\text{all cylinder sets on }L^{2}(D;U)\}, that is, all sets of the form E={uL2:(φ1,u,,φn,u)F}E=\big{\{}u\in L^{2}\ :\ (\langle\varphi_{1},\,u\rangle,\dots,\langle\varphi_{n},\,u\rangle)\in F\big{\}} for some nn\in\mathbb{N}, a Borel set FnF\subset\mathbb{R}^{n} and φ1,,φnL2(D;U)\varphi_{1},\dots,\varphi_{n}\in L^{2}(D;U). Since Cyl(L2)\mathrm{Cyl}(L^{2}) is an algebra which generates (L2)\mathscr{B}(L^{2}) (cf. e.g. [24, Appendix]), it follows from the monotone class lemma that 𝐌=σ(Cyl(L2))=(L2)\mathbf{M}=\sigma(\mathrm{Cyl}(L^{2}))=\mathscr{B}(L^{2}). Approximating an arbitrary φCb(L2(D;U))\varphi\in C_{b}(L^{2}(D;U)) by simple functions now gives the desired conclusion.

We claim that μt\mu_{t} is supported on Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U). This follows from Lemma B.5: Since μ\mu has bounded support on L2(D)L^{2}(D) we may take φ(ξ)=|ξ|2\varphi(\xi)=|\xi|^{2} in that lemma, which is continuous and bounded on any compact subset of \mathbb{R} and satisfies φ(0)=0\varphi(0)=0. Then by the lemma, for any gH1(D)g\in H^{1}(D),

L2(D)|Dugdx|2𝑑μt(u)=0.\int_{L^{2}(D)}\left|\int_{D}u\cdot\nabla g\,dx\right|^{2}\,d\mu_{t}(u)=0.

Hence, by Chebychev’s inequality,

μt({uL2(D):Dugdx0})=0.\mu_{t}\left(\left\{u\in L^{2}(D):\,\int_{D}u\cdot\nabla g\,dx\neq 0\right\}\right)=0.

Since gH1(D;U)g\in H^{1}(D;U) was arbitrary and H1(D;U)H^{1}(D;U) is separable, this implies that μt\mu_{t} is supported on L2L^{2}-functions that are weakly divergence free, which is exactly the space Ldiv2(D;U)L^{2}_{\operatorname{div}}(D;U) (see e.g. [57, Section 1, Chapter 1]).

Next we claim that μ\mu satisfies condition (c) in Definition 3.6. As 𝝂\bm{\nu} is assumed to satisfy (3.4), we can apply [24, Theorem 2.7] combined with Lemma B.4 to each of the terms and obtain that μt\mu_{t} satisfies (A.3) for any nonnegative and nondecreasing polynomial pK(s)=k=0Kakskp_{K}(s)=\sum_{k=0}^{K}a_{k}s^{k} on [0,R][0,R]. This implies in particular that μt\mu_{t} is supported on H1(D;U)H^{1}(D;U), for a.e. t[0,T)t\in[0,T), and (3.16). Now any differentiable nondecreasing function ψ\psi on a bounded interval can be approximated by nondecreasing polynomials [56], from which (3.14) follows after passing to the limit in a suitable polynomial approximation. Specifically, for a given ψ\psi, let {pKn}n\{p_{K_{n}}\}_{n\in\mathbb{N}} be a sequence of nonnegative, nondecreasing polynomials (KnK_{n}\to\infty) satisfying

pKnψC1([0,R])1n.\left\|p_{K_{n}}-\psi\right\|_{C^{1}([0,R])}\leqslant\frac{1}{n}.

Then, for t0t\geqslant 0, by the compact support property of μ\mu on BB,

|L2(D)ψ(uL2(D)2)𝑑μt(u)L2(D)pKn(uL2(D)2)𝑑μt(u)|pKnψC1([0,R])1n,\left|\int_{L^{2}(D)}\psi(\left\|u\right\|_{L^{2}(D)}^{2})d\mu_{t}(u)-\int_{L^{2}(D)}p_{K_{n}}(\left\|u\right\|^{2}_{L^{2}(D)})d\mu_{t}(u)\right|\leqslant\left\|p_{K_{n}}-\psi\right\|_{C^{1}([0,R])}\leqslant\frac{1}{n},

and

|0tL2(D)ψ(uL2(D)2)|u|H1(D)2𝑑μt(u)𝑑s0tL2(D)pKn(uL2(D)2)|u|H1(D)2𝑑μt(u)𝑑s|\displaystyle\left|\int_{0}^{t}\int_{L^{2}(D)}\psi^{\prime}(\left\|u\right\|_{L^{2}(D)}^{2})|u|^{2}_{H^{1}(D)}d\mu_{t}(u)ds-\int_{0}^{t}\int_{L^{2}(D)}p_{K_{n}}^{\prime}(\left\|u\right\|^{2}_{L^{2}(D)})|u|^{2}_{H^{1}(D)}d\mu_{t}(u)ds\right|
0tL2(D)|ψ(uL2(D)2)pKn(uL2(D)2)||u|H1(D)2𝑑μt(u)𝑑s\displaystyle\qquad\leqslant\int_{0}^{t}\int_{L^{2}(D)}\left|\psi^{\prime}(\left\|u\right\|_{L^{2}(D)}^{2})-p_{K_{n}}^{\prime}(\left\|u\right\|^{2}_{L^{2}(D)})\right||u|^{2}_{H^{1}(D)}d\mu_{t}(u)ds
pKnψC1([0,R])0tL2(D)|u|H1(D)2𝑑μt(u)𝑑s\displaystyle\qquad\leqslant\left\|p_{K_{n}}-\psi\right\|_{C^{1}([0,R])}\int_{0}^{t}\int_{L^{2}(D)}|u|^{2}_{H^{1}(D)}d\mu_{t}(u)ds
Cn,\displaystyle\qquad\leqslant\frac{C}{n},

where the last inequality follows from (3.16).

It remains to show that the correlation measures satisfy the evolution equation (3.13) for all Φ𝒯cyl0\Phi\in\mathscr{T}_{\text{cyl}}^{0}. From this also part (d) of the definition will follow (see Remark 3.7). Let Φ(u)=φ((u,g1),,(u,gk))\Phi(u)=\varphi((u,g_{1}),\dots,(u,g_{k})) be an arbitrary function in 𝒯cyl0\mathscr{T}_{\text{cyl}}^{0} with gjH1(D;U)g_{j}\in H^{1}(D;U) and let R~=max1jk(gjL2)\widetilde{R}=\max_{1\leqslant j\leqslant k}(\left\|g_{j}\right\|_{L^{2}}). Since φ\varphi is continuously differentiable and bounded on B[RR~1,RR~+1]kB\coloneqq[-R\widetilde{R}-1,R\widetilde{R}+1]^{k}, we can approximate it arbitrarily well by polynomials thanks to the Weierstrass approximation theorem. Let

pn(ζ)|j¯|=0Nnβj¯ζ1j1ζkjk,ji0,j¯=(j1,,jk),|j¯|j1++jk,p_{n}(\zeta)\coloneqq\sum_{|\underline{j}|=0}^{N_{n}}\beta_{\underline{j}}\zeta_{1}^{j_{1}}\dots\zeta_{k}^{j_{k}},\quad j_{i}\geqslant 0,\quad\underline{j}=(j_{1},\dots,j_{k}),\quad|\underline{j}|\coloneqq j_{1}+\dots+j_{k},

for some NnN_{n}\in\mathbb{N} large enough, be approximations of φ\varphi that satisfy

φpnC1(B)1n,n.\left\|\varphi-p_{n}\right\|_{C^{1}(B)}\leqslant\frac{1}{n},\quad n\in\mathbb{N}.

Let θCc1((0,T))\theta\in C_{c}^{1}((0,T)) be an arbitrary compactly supported test function. Since by the equivalence theorem, for any j¯=(j1,,jk)\underline{j}=(j_{1},\dots,j_{k}), ji0j_{i}\geqslant 0, |j¯|=j1++jk|\underline{j}|=j_{1}+\dots+j_{k},

0Tθ(t)D|j¯|(U|j¯|ξ1ξ|j¯|𝑑νt,x|j¯|(ξ))\displaystyle\int_{0}^{T}\theta^{\prime}(t)\int_{D^{|\underline{j}|}}\left(\int_{U^{|\underline{j}|}}\xi_{1}\otimes\cdots\otimes\xi_{|\underline{j}|}\,d\nu_{t,x}^{|\underline{j}|}(\xi)\right)
:(g1(x1)g1(xj1)gk(x|j¯|jk+1)gk(x|j¯|))dxdt\displaystyle:\Big{(}g_{1}(x_{1})\otimes\cdots\otimes g_{1}(x_{j_{1}})\otimes\cdots\otimes g_{k}(x_{|\underline{j}|-j_{k}+1})\otimes\cdots\otimes g_{k}(x_{|\underline{j}|})\Big{)}\,dx\,dt
=0Tθ(t)L2(D)(u,g1)j1(u,gk)jk𝑑μt(u)𝑑t,\displaystyle=\int_{0}^{T}\theta^{\prime}(t)\int_{L^{2}(D)}(u,g_{1})^{j_{1}}\dots(u,g_{k})^{j_{k}}\,d\mu_{t}(u)\,dt,

and if j~=j1++ji1\widetilde{j}=j_{1}+\dots+j_{i-1}, then

=j~+1j~+ji0Tθ(t)D|j¯|(U|j¯|ξ1ξ|j¯|𝑑νt,x|j¯|(ξ))\displaystyle\sum_{\ell=\widetilde{j}+1}^{\widetilde{j}+j_{i}}\int_{0}^{T}\!\!\!\theta(t)\!\int_{D^{|\underline{j}|}}\left(\int_{U^{|\underline{j}|}}\xi_{1}\otimes\cdots\otimes\xi_{|\underline{j}|}\,d\nu_{t,x}^{|\underline{j}|}(\xi)\right)
:(g1(x1)gi(xj~+1)Δxgi(x)gi(xj~+ji)gk(x|j¯|))dxdt\displaystyle:\Big{(}g_{1}(x_{1})\otimes\cdots\otimes g_{i}(x_{\widetilde{j}+1})\otimes\dots\otimes\Delta_{x_{\ell}}g_{i}(x_{\ell})\otimes\dots\otimes g_{i}(x_{\widetilde{j}+j_{i}})\otimes\dots\otimes g_{k}(x_{|\underline{j}|})\Big{)}\,dx\,dt
=\displaystyle= ji0Tθ(t)L2(D;U)(u,g1)j1(u,gi)ji1(u,Δgi)(u,gk)jk𝑑μt(u)𝑑t\displaystyle\ j_{i}\int_{0}^{T}\!\!\theta(t)\int_{L^{2}(D;U)}(u,g_{1})^{j_{1}}\cdots(u,g_{i})^{j_{i}-1}(u,\Delta g_{i})\cdots(u,g_{k})^{j_{k}}\,d\mu_{t}(u)\,dt
=\displaystyle= ji0Tθ(t)L2(D;U)(u,g1)j1(u,gi)ji1(Au,gi)(u,gk)jk𝑑μt(u)𝑑t,\displaystyle\ j_{i}\int_{0}^{T}\!\!\theta(t)\int_{L^{2}(D;U)}(u,g_{1})^{j_{1}}\cdots(u,g_{i})^{j_{i}-1}(Au,g_{i})\cdots(u,g_{k})^{j_{k}}\,d\mu_{t}(u)\,dt,

where the last equality follows because gig_{i} is divergence free because μt\mu_{t} is supported on H1(D;U)H^{1}(D;U)-functions by the energy inequality (3.14). Furthermore, we have

=j~+1j~+ji0Tθ(t)D|j¯|(U|j¯|ξ1ξj~+1(ξξ)ξj~+jiξ|j¯|𝑑νt,x|j¯|(ξ))\displaystyle\sum_{\ell=\widetilde{j}+1}^{\widetilde{j}+j_{i}}\int_{0}^{T}\!\!\!\theta(t)\!\int_{D^{|\underline{j}|}}\left(\int_{U^{|\underline{j}|}}\xi_{1}\otimes\cdots\otimes\xi_{\widetilde{j}+1}\otimes\cdots\otimes(\xi_{\ell}\otimes\xi_{\ell})\otimes\cdots\otimes\xi_{\widetilde{j}+j_{i}}\otimes\cdots\otimes\xi_{|\underline{j}|}\,d\nu_{t,x}^{|\underline{j}|}(\xi)\right)
:(g1(x1)xgi(x)gk(x|j¯|))dxdt\displaystyle:\Big{(}g_{1}(x_{1})\otimes\cdots\otimes\nabla_{x_{\ell}}g_{i}(x_{\ell})\otimes\cdots\otimes g_{k}(x_{|\underline{j}|})\Big{)}\,dx\,dt
=\displaystyle= ji0Tθ(t)L2(D)(u,g1)j1(u,gi)ji1(uu,gi)(u,gk)jk𝑑μt(u)𝑑t\displaystyle\ j_{i}\int_{0}^{T}\!\!\theta(t)\int_{L^{2}(D)}(u,g_{1})^{j_{1}}\cdots(u,g_{i})^{j_{i}-1}(u\otimes u,\nabla g_{i})\cdots(u,g_{k})^{j_{k}}\,d\mu_{t}(u)\,dt
=\displaystyle= ji0Tθ(t)L2(D)(u,g1)j1(u,gi)ji1(B(u),gi)(u,gk)jk𝑑μt(u)𝑑t,\displaystyle\ -j_{i}\int_{0}^{T}\!\!\theta(t)\int_{L^{2}(D)}(u,g_{1})^{j_{1}}\cdots(u,g_{i})^{j_{i}-1}(B(u),g_{i})\cdots(u,g_{k})^{j_{k}}\,d\mu_{t}(u)\,dt,

again, the last equality following because gig_{i} is divergence free and μt\mu_{t} is supported on H1(D;U)H^{1}(D;U)-functions for almost every t[0,T]t\in[0,T]. We obtain, by combining the terms to the Friedman–Keller system, that if

Φn(u)pn((u,g1),,(u,gk))\Phi_{n}(u)\coloneqq p_{n}\big{(}(u,g_{1}),\dots,(u,g_{k})\big{)}

then μ\mu satisfies the equation

(A.5) 0TL2(D)θ(s)Φn(u)𝑑μs(u)𝑑s+0Tθ(s)L2(D)(F(s,u),uΦn(u))𝑑μs(u)𝑑s=0\int_{0}^{T}\int_{L^{2}(D)}\theta^{\prime}(s)\Phi_{n}(u)\,d\mu_{s}(u)\,ds+\int_{0}^{T}\theta(s)\int_{L^{2}(D)}(F(s,u),\partial_{u}\Phi_{n}(u))\,d\mu_{s}(u)\,ds=0

for every test function θCc1((0,T))\theta\in C_{c}^{1}((0,T)). Next, note that

(A.6) |L2(D)Φn(u)𝑑μt(u)L2(D)Φ(u)𝑑μt(u)|=|L2(D)(Φn(u)Φ(u))𝑑μt(u)|L2(D)|pn((u,g1),,(u,gk))φ((u,g1),,(u,gk))|𝑑μt(u)L2(D)supζB|pn(ζ)φ(ζ)|dμt(u)=φpnC0(B)1n,\begin{split}&\left|\int_{L^{2}(D)}\Phi_{n}(u)\,d\mu_{t}(u)-\int_{L^{2}(D)}\Phi(u)\,d\mu_{t}(u)\right|=\left|\int_{L^{2}(D)}\left(\Phi_{n}(u)-\Phi(u)\right)\,d\mu_{t}(u)\right|\\ &\quad\leqslant\int_{L^{2}(D)}\big{|}p_{n}\big{(}(u,g_{1}),\dots,(u,g_{k})\big{)}-\varphi\big{(}(u,g_{1}),\dots,(u,g_{k})\big{)}\big{|}\,d\mu_{t}(u)\\ &\quad\leqslant\int_{L^{2}(D)}\sup_{\zeta\in B}\left|p_{n}(\zeta)-\varphi(\zeta)\right|\,d\mu_{t}(u)\\ &\quad=\left\|\varphi-p_{n}\right\|_{C^{0}(B)}\leqslant\frac{1}{n},\end{split}

because μt\mu_{t} has support on {uL2(D):uL2(D)R}\{u\in L^{2}(D)\,:\,\|u\|_{L^{2}(D)}\leqslant R\}. Moreover,

(A.7) |0Tθ(s)L2(D)(uΦn(u),F(s,u))𝑑μs(u)𝑑s0Tθ(s)L2(D)(uΦ(u),F(s,u))𝑑μs(u)𝑑s|θL0TL2(D)|(u(Φ(u)Φn(u)),F(s,u))|𝑑μs(u)𝑑sθLj=1k0TL2(D)|jφ((u,g1),,(u,gk))jpn((u,g1),,(u,gk))||(gj,F(s,u))|𝑑μs(u)𝑑sθLj=1ksupB|j(φpn)|0TL2(D)|(gj,F(s,u))|𝑑μs(u)𝑑sθLφpnC1(B)j=1k0TL2(D)|(gj,F(s,u))|𝑑μs(u)𝑑s12nθLj=1k(gjH1(D)2+0TL2(D)uL2(D)2𝑑μs(u)𝑑s)\begin{split}&\left|\int_{0}^{T}\theta(s)\int_{L^{2}(D)}\big{(}\partial_{u}\Phi_{n}(u),F(s,u)\big{)}\,d\mu_{s}(u)\,ds-\int_{0}^{T}\theta(s)\int_{L^{2}(D)}\big{(}\partial_{u}\Phi(u),F(s,u)\big{)}\,d\mu_{s}(u)\,ds\right|\\ &\quad\leqslant\|\theta\|_{L^{\infty}}\int_{0}^{T}\int_{L^{2}(D)}\big{|}\big{(}\partial_{u}(\Phi(u)-\Phi_{n}(u)),F(s,u)\big{)}\big{|}\,d\mu_{s}(u)\,ds\\ &\quad\leqslant\|\theta\|_{L^{\infty}}\sum_{j=1}^{k}\int_{0}^{T}\int_{L^{2}(D)}\big{|}\partial_{j}\varphi\big{(}(u,g_{1}),\dots,(u,g_{k})\big{)}-\partial_{j}p_{n}\big{(}(u,g_{1}),\dots,(u,g_{k})\big{)}\big{|}\big{|}(g_{j},F(s,u))\big{|}\,d\mu_{s}(u)\,ds\\ &\quad\leqslant\|\theta\|_{L^{\infty}}\sum_{j=1}^{k}\sup_{B}\left|\partial_{j}(\varphi-p_{n})\right|\int_{0}^{T}\int_{L^{2}(D)}\big{|}(g_{j},F(s,u))\big{|}\,d\mu_{s}(u)\,ds\\ &\quad\leqslant\|\theta\|_{L^{\infty}}\left\|\varphi-p_{n}\right\|_{C^{1}(B)}\sum_{j=1}^{k}\int_{0}^{T}\int_{L^{2}(D)}\big{|}(g_{j},F(s,u))\big{|}\,d\mu_{s}(u)\,ds\\ &\quad\leqslant\frac{1}{2n}\|\theta\|_{L^{\infty}}\sum_{j=1}^{k}\left(\left\|g_{j}\right\|_{H^{1}(D)}^{2}+\int_{0}^{T}\int_{L^{2}(D)}\left\|\nabla u\right\|_{L^{2}(D)}^{2}\,d\mu_{s}(u)\,ds\right)\end{split}

where we used Young’s inequality for the last inequality. Thanks to (A.6) and (A.7), we can pass to the limit nn\rightarrow\infty in (A.5) and obtain

(A.8) 0Tθ(s)L2(D)Φ(u)𝑑μs(u)𝑑s+0Tθ(s)L2(D)(F(s,u),uΦ(u))𝑑μs(u)𝑑s=0\int_{0}^{T}\theta^{\prime}(s)\int_{L^{2}(D)}\Phi(u)\,d\mu_{s}(u)\,ds+\int_{0}^{T}\theta(s)\int_{L^{2}(D)}(F(s,u),\partial_{u}\Phi(u))\,d\mu_{s}(u)\,ds=0

for every θCc1((0,T))\theta\in C_{c}^{1}((0,T)). It follows that the distributional derivative of φ(s)L2(D)Φ(u)𝑑μs(u)\varphi(s)\coloneqq\int_{L^{2}(D)}\Phi(u)\,d\mu_{s}(u) is

ddsφ(s)=L2(D)(F(s,u),uΦ(s,u))𝑑μs(u),\frac{d}{ds}\varphi(s)=\int_{L^{2}(D)}(F(s,u),\partial_{u}\Phi(s,u))\,d\mu_{s}(u),

and since the right-hand side lies in L1((0,T))L^{1}((0,T)) we see that φ\varphi is absolutely continuous. Consequently, we can (through a standard approximation procedure) insert θ=𝟙(0,t)\theta=\mathbbm{1}_{(0,t)} in (A.8) and conclude that (3.13) is true.

Appendix B Auxiliary lemmata

Lemma B.1 (Representation of the gradient).

Let (μt)0tT(\mu_{t})_{0\leqslant t\leqslant T} be a measure on L2(D)L^{2}(D) satisfying

(B.1) 0TL2(D)uL2(D)2𝑑μt(u)𝑑tC<.\int_{0}^{T}\int_{L^{2}(D)}\left\|\nabla u\right\|_{L^{2}(D)}^{2}\,d\mu_{t}(u)\,dt\leqslant C<\infty.

Define for hh\in\mathbb{R} the finite difference gradient h\nabla_{h} by

hf(x)=(D1hf(x),,Ddhf(x)),Djhf(x)=f(x+h𝐞j)f(x)h,\nabla_{h}f(x)=\left(D^{h}_{1}f(x),\dots,D^{h}_{d}f(x)\right),\quad D^{h}_{j}f(x)=\frac{f(x+h\mathbf{e}_{j})-f(x)}{h},

where 𝐞jd\mathbf{e}_{j}\in\mathbb{R}^{d} is the jjth unit vector. Then we have for any VDV\subset\subset D (or in the case that DD is the torus also for V=DV=D)

(B.2) 0TL2(