Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-01-09T16:34:38.594Z Has data issue: false hasContentIssue false

On the reconstruction of unknown driving forces from low-mode observations in the 2D Navier–Stokes equations

Published online by Cambridge University Press:  01 April 2024

Vincent R. Martinez*
Affiliation:
Department of Mathematics & Statistics, CUNY Hunter College, New York, NY, USA Department of Mathematics, New York, NY, CUNY Graduate Center, USA ([email protected])
Rights & Permissions [Opens in a new window]

Abstract

This article is concerned with the problem of determining an unknown source of non-potential, external time-dependent perturbations of an incompressible fluid from large-scale observations on the flow field. A relaxation-based approach is proposed for accomplishing this, which makes use of a nonlinear property of the equations of motions to asymptotically enslave small scales to large scales. In particular, an algorithm is introduced that systematically produces approximations of the flow field on the unobserved scales in order to generate an approximation to the unknown force; the process is then repeated to generate an improved approximation of the unobserved scales, and so on. A mathematical proof of convergence of this algorithm is established in the context of the two-dimensional Navier–Stokes equations with periodic boundary conditions under the assumption that the force belongs to the observational subspace of phase space; at each stage in the algorithm, it is shown that the model error, represented as the difference between the approximating and true force, asymptotically decreases to zero in a geometric fashion provided that sufficiently many scales are observed and certain parameters of the algorithm are appropriately tuned.

Type
Research Article
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh

1. Introduction

In the derivation of any model, parameters arise that capture intrinsic properties of the phenomenon of interest. In the case of modelling a turbulent, incompressible fluid flow via the Navier–Stokes equations (NSE), assuming a constant density, the two relevant parameters are essentially the kinematic viscosity of the fluid and the external force. For instance, in eddy diffusivity models, turbulent viscosity coefficients must be specified [Reference Berselli, Iliescu and Layton6]. In practice, these parameters must be empirically determined. On the contrary, in proposing any turbulent closure, one inevitably commits model errors. These errors may themselves then be modelled as a body force, whose exact form in terms of the mean field is unknown.

In the current study, we propose an algorithm for determining all large-scale features of an external driving force in the two-dimensional incompressible NSE (2D NSE) down to the observation scale. In practice, such an algorithm can be used to filter model errors that are represented as external driving forces. In our idealized set-up, the fluid occupies a periodic domain, $\Omega =[0,2\pi ]^2$, the density of the fluid is normalized to unity, the kinematic viscosity of the fluid is known perfectly, but the external driving force is not. We point out that the problem of determining the viscosity based on direct observation of the velocity field has recently been studied in [Reference Biswas and Hudson9, Reference Carlson, Hudson and Larios15, Reference Cialenco and Glatt-Holtz21, Reference Clark Di Leoni, Mazzino and Biferale22, Reference Martinez61], where estimators for the viscosity were proposed, their consistency and asymptotic normality were established [Reference Cialenco and Glatt-Holtz21], convergence analyses for viscosity-recovery algorithms were carried out [Reference Martinez61], and numerical tests were performed [Reference Carlson, Hudson and Larios15, Reference Clark Di Leoni, Mazzino and Biferale22]. The problem of multi-parameter recovery in chaotic dynamical systems was studied in [Reference Carlson, Hudson, Larios, Martinez, Ng and Whitehead16, Reference Pachev, Whitehead and McQuarrie63], while a recent study [Reference Carlson and Larios17] analysed the sensitivity of the 2D NSE to changes in the viscosity, as well as its implications for a certain downscaling algorithm for data assimilation.

Generally speaking, one of the objectives of this article is to study the extent to which external forces can be determined based on error-free, but partial observations of flow field through a practically implementable algorithm. In our ideal set-up, we will assume that we have access to a time series of the velocity field, but only through sufficiently small, but nevertheless finitely many, length scales. Ultimately, we introduce an algorithm for reconstructing large-scale features of the unknown force and establish its convergence under the assumption that the force acts only on length scales that are directly observed, that is to say, that the force belongs to the span of the observational field.

To be more precise, we recall that the 2D NSE on $\Omega =[0,2\pi ]^2$ is given by

(1.1)\begin{equation} \partial_tu+(u\cdotp\nabla) u=\nu\Delta u-\nabla p+f,\quad \nabla\cdotp u=0, \end{equation}

where the kinematic viscosity, $\nu >0$, is given and fixed, and the time series, $\{P_Nu(t)\}_{t\in [0,T]}$, is known, up to some time $T>0$, where $P_N$ is the $L^2$-orthogonal projection onto Fourier wavenumbers corresponding to $|k|\leq N$, and $u\cdotp \nabla =u^j\partial _j$, where repeated indices indicate summation over those indices. However, the external force, $f$, is time-independent, but unknown. The scalar pressure field is denoted by $p$; upon taking the divergence of (1.1), one sees that $p$ is determined entirely by $u, f$ via the Poisson equation $-\Delta p=\partial _i\partial _j(u^iu^j)+\partial _jf^j$. We will assume that $u$, $p$, and $f$ are both mean-free and periodic over $\Omega$, and that $f$ is divergence-free. The main result is that for forces satisfying $P_Nf=f$, where $N$ is sufficiently large, there exists a sequence of times $\{t_n\}_n$ and sequence of approximating forces, $\{f_n\}_n$, depending only on the observations $\{P_Nu(t)\}_{t\geq 0}$ such that $f_n$ converges to $f$. We will in fact address the more general case of time-dependent forces.

The motivating idea for the algorithm we propose is based on the notion of ‘asymptotic completeness’ for nonlinear dissipative systems. For systems that possess this property, it asserts that having direct (observational) access to a sufficiently rich, but nevertheless finite-dimensional, set of scales, is enough to asymptotically determine the unobserved scales. This property was rigorously shown to hold for (1.1) by Foiaş and Prodi [Reference Foiaş and Prodi44] in the case when access to sufficiently many Fourier modes of the velocity field is available; for the case of three dimensions, the reader is referred to [Reference Biswas and Price11, Reference Cheskidov and Dai20]. Specifically, given two solutions $u_1, u_2$ of (1.1) corresponding to external forces $f_1, f_2$, it was shown that there exists $N\geq 1$, depending on $\nu$ and $f_1,f_2$, but only through their size, such that $P_N(u_1(t)-u_2(t))\rightarrow 0$ and $f_1(t)-f_2(t)\rightarrow 0$ as $t\rightarrow \infty$ together imply $u_1(t)-u_2(t)\rightarrow 0$ as $t\rightarrow \infty$. In this case, it is then said that the Fourier modes corresponding to wavenumbers $|k|\leq N$ are determining for system (1.1); the smallest such number, $N$, is referred to as the number of determining modes. In light of this result, one sees that the problem of inferring an unknown force may be possible to solve provided that sufficiently many modes are observed and that one identifies an algorithm that asymptotically reconstructs higher modes that can subsequently be made use of to approximate the force.

To see the main difficulty that must be overcome in doing so, suppose that one is given access to $P_Nu(t)$, for $t\geq t_0$. Then, a naïve, but reasonable first approximation to the low modes of the force may be given by simply evaluating the nonlinear differential operator determined by (1.1) along $P_Nu$. This yields

(1.2)\begin{equation} \partial_tP_Nu-\nu\Delta P_Nu+P_N((P_Nu\cdotp\nabla)P_Nu)+\nabla p_N=:f_0, \end{equation}

where $p_N$ is determined by enforcing $f_0$ to be divergence-free. On the contrary, by applying the low-pass filter $P_N$ to (1.1), one also obtains:

(1.3)\begin{equation} \partial_tP_Nu-\nu\Delta P_Nu+P_N((P_Nu\cdotp\nabla)P_Nu)+\nabla p_N=P_Nf+\mathcal{R}_N, \end{equation}

where $\mathcal {R}_N$ denotes the Reynolds stress, defined as:

(1.4)\begin{equation} \mathcal{R}_N:=P_N((P_Nu\cdotp\nabla)P_Nu)-P_N((u\cdotp\nabla)u)+\nabla p_N-\nabla P_Np. \end{equation}

One therefore has the following relation:

(1.5)\begin{equation} \mathcal{R}_N=f_0-P_Nf. \end{equation}

In particular, determination of $P_Nf$ is equivalent to determination of $\mathcal {R}_N$. The fundamental difficulty of the problem arises from the elementary fact that $\mathcal {R}_N$ depends on both $P_Nu$ and $(I-P_N)u$, which is tantamount to the closure problem of turbulence. In comparison, if (1.1) were replaced by the linear heat equation, then observation of $P_Nu$ along a trajectory of $u$ provides an exact reconstruction for $P_Nf$ along that same trajectory. Thus, any resolution of this difficulty must address a way to reconstruct high-mode information from low-mode data. In § 3, we introduce an algorithm that, under suitable conditions, accomplishes this and systematically decreases the ‘Reynolds stress’ at each stage in a geometric fashion, yielding a convergent scheme for approximating the force on the low modes. From this point of view, the proposed scheme may be viewed as a ‘nonlinear filtering’ of large-scale error.

For systems that possess an inertial manifold, reconstruction of small scales from large scales would be possible. Indeed, the existence of an inertial manifold implies a strong form of enslavement of scales as it asserts the existence of a map $\Phi$ such that $P_Nu|_{t=t_0}\mapsto (I-P_N)u|_{t\geq t_0}$, that is, $u(t)=P_Nu(t)+\Phi (P_Nu(t_0))(t)$, for all $t\geq t_0$. In other words, knowledge of low modes at a single time is sufficient to determine the high-mode behaviour at all future times. The inertial manifold would then be determined by the graph of $\Phi$ (see [Reference Foias, Sell and Temam45, Reference Foias and Temam47]). However, the existence of such a map for (1.1) remains an outstanding open problem. One of the main points of this article is that one needs to only rely on the weaker property of asymptotic completeness to reconstruct sufficiently high-mode components of the state variables in order to eventually recover the low modes of the forcing.

There are at least two ways available in the literature for doing, one due to Foias et al. in [Reference Foias, Jolly, Kravchenko and Titi39] and another by Azouani et al. in [Reference Azouani, Olson and Titi4]. The first construction [Reference Foias, Jolly, Kravchenko and Titi39] allows one to encode projections of solutions on the global attractor of (1.1) as travelling wave solutions to some infinite-dimensional ordinary differential equation. However, it effectively requires one to solve the evolution equation of (1.1) corresponding to the high modes [Reference Foias, Jolly, Lithio and Titi40]. The second method introduces a downscaling algorithm for data assimilation [Reference Azouani, Olson and Titi4], in which large-scale observations are exogenously inserted into (1.1) as a feedback-control term that serves to drive the corresponding solution of this modified system towards the reference solution, but only on large scales. By virtue of the Foias–Prodi property of determining modes, it was then shown that by tuning the strength of the feedback-control system appropriately, the generated approximating signal asymptotically synchronizes to the reference signal to which the large-scale observations correspond. We will make use of the latter approach to systematically reconstruct high-mode information on the reference field.

The remainder of the paper is organized as follows. In § 2, we establish the notation and functional setting in which the result will be proved, as well as classical well-posedness results that we will make use of. A derivation of the algorithm that reconstructs the external force is presented in § 3. We provide formal statements of the convergence results in § 4, then outline their proof in § 5. The main step is to reduce the analysis of the convergence to establish suitable ‘sensitivity estimates’. These estimates are proved in § 6. We finally supply rigorous proofs of the main results in § 7. Technical auxiliary results are relegated to Appendix A.

2. Notation and functional setting

Let $\Omega =[0,2\pi ]^2$ and $L^p_\sigma =L^p_\sigma (\Omega )$, for $p\in [1,\infty ]$, denote the space of $p$-integrable (in the sense of Lebesgue), mean-free, solenoidal vector fields over $\Omega$, which are periodic in each direction; its norm is given by

(2.1)\begin{equation} \lVert u\rVert_{L^p}^p:=\int_{\Omega} |u(x)|^p\,{\rm d}x,\quad \text{for}\ p\in[1,\infty),\quad \lVert u\rVert_{L^\infty}:=\operatorname{ess\ sup}_{x\in\Omega}|u(x)|, \end{equation}

where $\operatorname {ess\ sup}$ denotes the essential supremum. We let $H^k_\sigma =H^k_\sigma (\Omega )$ denote the space of periodic, mean-free, solenoidal vector fields over $\Omega$ whose weak derivatives (in the sense of Sobolev) up to order $k$ belong to $L^2_\sigma$; its norm is given by

(2.2)\begin{equation} \lVert u\rVert_{H^k}^2:=\sum_{|\alpha|\leq k}\int_{\Omega}|\partial^\alpha u(x)|^2\,{\rm d}x, \end{equation}

where $\alpha \in (\mathbb {N}\cup \{0\})^2$ is a multi-index. We will abuse notation and use $L^2_\sigma, H^k_\sigma$ to denote the corresponding spaces for scalar functions as well; in this case, $|u|$ is interpreted as absolute value, rather than Euclidean norm.

We will make use of the functional form of the NSE, which is given by

(2.3)\begin{equation} \frac{{\rm d}}{{\rm d}t}u+\nu Au+B(u,u)=Pf, \end{equation}

where

\[ Au={-}P\Delta u\quad\text{and}\quad B(u,v)=P(u\cdotp\nabla) v, \]

and $P$ denotes the Leray–Helmholtz projection, that is, the orthogonal projection onto divergence-free vector fields; we refer to $A$ as the Stokes operator. Note that by orthogonality, $\lVert P \rVert _{L^2\rightarrow L^2}\leq 1$. Moreover,

(2.4)\begin{equation} \widehat{Pu}(n):=\hat{u}(n)-\frac{n\cdotp\hat{u}(n)}{|n|^2}n,\quad n\in\mathbb{Z}^2\setminus\{(0,0)\}, \end{equation}

where $\hat {u}(n)$ denotes the Fourier coefficient of $u$ corresponding to wavenumber $n\in \mathbb {Z}^2$. Also, powers of $A$ can be defined spectrally via:

(2.5)\begin{equation} \widehat{A^{k/2}u}(n):=|n|^k\hat{u}(n). \end{equation}

Hence, $P$ commutes with $\Delta$ in the setting of periodic boundary conditions. Recall that for $u\in H^1_\sigma$ the Poincaré inequality states:

(2.6)\begin{equation} \lVert u\rVert_{L^2}\leq \lVert A^{1/2} u\rVert_{L^2}=\lVert A^{1/2}u\rVert_{L^2}. \end{equation}

It follows that for each integer $k\geq 1$, there exists a constant $c_k>0$ such that

(2.7)\begin{equation} c_k^{{-}2}\lVert u\rVert_{H^k}^2\leq \sum_{|\alpha|=k}\int_\Omega|\partial^\alpha u(x)|^2\,{\rm d}x\leq c_k^2\lVert u\rVert_{H^k}^2, \end{equation}

whenever $u\in H^k_\sigma$. In particular, by the Parseval identity, it follows that there also exist constants $c_k>0$ such that

(2.8)\begin{equation} c_k^{{-}1}\lVert u\rVert_{H^k}\leq \lVert A^{k/2}u\rVert_{L^2}\leq c_k\lVert u\rVert_{H^k}, \end{equation}

for all $u\in H^k_\sigma$.

Given $t_0\geq 0$ and $f\in L^\infty (t_0,\infty ;L^2_\sigma )$, we define the Grashof-type number by

(2.9)\begin{equation} \tilde{G}:=\frac{\sup_{t\geq t_0}\lVert f(t)\rVert_{L^2}}{\kappa_0^2\nu^2}, \end{equation}

where $\kappa _0=(2\pi )/L$, where $L$ is the linear length of the domain; since $\Omega =[0,2\pi ]^2$, we see that $\kappa _0=1$. We note that the Grashof number is traditionally denoted by $G$, that is, undecorated, when the force is independent of time. Since we allow for time-dependence in the force, we will distinguish between the two notations by making use of tilde. See, for instance [Reference Balci, Foias and Jolly5], where this distinction is also maintained.

Observe that $Pf=f$, for any $f\in L^2_\sigma$. One has the following classical results regarding the existence theory for (2.3) [Reference Constantin and Foias26, Reference Foias, Manley, Rosa and Temam41, Reference Temam65, Reference Temam66].

Theorem 2.1 Let $t_0\geq 0$ and $f\in L^\infty (t_0,\infty ;L^2_\sigma )$. For all $u_0\in H^1_\sigma$, there exists a unique $u\in C([t_0,T];H_\sigma ^1)\cap L^2(t_0,T;H^{2})$ satisfying (2.3), for all $T>0$, such that $({{\rm d}}/{{\rm d}t})u\in L^2(t_0,T;L^2_\sigma )$ and

(2.10)\begin{equation} \lVert A^{1/2}u(t)\rVert_{L^2}^2\leq \lVert A^{1/2} u(t_0)\rVert_{L^2}^2\,{\rm e}^{-\nu (t-t_0)}+\nu^2\tilde{G}^2(1-{\rm e}^{-\nu(t-t_0)}), \end{equation}

for all $t\geq t_0\geq 0$.

From (2.10), we see that for $t_0=0$ and all $t>0$ sufficiently large, depending on $\lVert A ^{1/2}u_0\rVert _{L^2}$, one has:

(2.11)\begin{equation} \lVert A^{1/2}u(t)\rVert_{L^2}\leq \sqrt{2}\nu \tilde{G}=:{\tilde{c}}_1{\tilde{R}}_1,\quad {\tilde{c}}_1=\sqrt{2}. \end{equation}

Let ${\tilde {B}}_1$ denote the ball of radius $\tilde {R}_1$, centred at the origin in $L^2_\sigma$. Observe that $A^{1/2}u_0\in \tilde {B}_1$ implies via (2.10) that $u(t;u_0,f)\in \tilde {B}_1$, for all $t\geq 0$, where $u(t;u_0,f)$ denotes the unique solution of (2.3) with initial data $u_0$ and external forcing $f$. Moreover, there exists a constant ${\tilde {c}}_2>0$ such that if $f\in L^\infty (0,\infty ;H^1_\sigma )$ and $Au_0\in \alpha {\tilde {B}}_2$, where $\alpha >0$ is arbitrary, ${\tilde {B}}_2$ denotes the ball of radius $\tilde {R}_2={\tilde {c}}_2\nu (\tilde {G}+\tilde {\sigma }_1)\tilde {G}$, centred at the origin in $L^2_\sigma$, and $\alpha {\tilde {B}}_2$ the same ball of radius $\alpha {\tilde {R}}_2$, then

(2.12)\begin{equation} \lVert Au(t)\rVert_{L^2}^2\leq (1+\alpha^2)\tilde{c}_2^2\nu^2(\tilde{\sigma}_1+\tilde{G})^2\tilde{G}^2, \end{equation}

for all $t\geq 0$, where $\tilde {\sigma }_1$ denotes a ‘shape factor’ defined by

(2.13)\begin{equation} \tilde{\sigma}_1:=\frac{\sup_{t\geq t_0}\lVert A^{1/2} f(t)\rVert_{L^2}}{\sup_{t\geq t_0}\lVert f(t)\rVert_{L^2}}. \end{equation}

In other words, $Au_0\in \alpha {\tilde {B}}_2$ implies $Au(t)\in (1+\alpha ^2)^{1/2}{\tilde {B}}_2$, for all $t\geq 0$. Observe that $\tilde {\sigma }_1\geq 1$ by Poincaré's inequality. In particular, if $Au_0\in {\tilde {B}}_2$, then

(2.14)\begin{equation} \lVert Au(t)\rVert_{L^2}\leq \sqrt{2}{\tilde{c}}_2\nu(\tilde{G}+\tilde{\sigma}_1)\tilde{G}=\sqrt{2}{\tilde{R}}_2. \end{equation}

Bounds sharper than (2.14) were established in [Reference Dascaliuc, Foias and Jolly27] in the setting where $f$ was time-independent. In this particular case, one has that:

(2.15)\begin{equation} \lVert Au(t)\rVert_{L^2}\leq c_2\nu(\sigma_{1}^{1/2}+G)G=:R_2, \end{equation}

holds for all $t\geq 0$, for some $c_2>0$, provided that $u_0\in B_2$, the ball of radius $R_2$ in $H^2_\sigma$. Here, $G$ denotes the Grashof number, which is simply given by (2.9) when $f$ is time-independent. Similarly, $\sigma _1$ is given by (2.13) when $f$ is time-independent. For the sake of completeness, we supply a short proof of (2.15) in Appendix A.

3. Description of the algorithm

We consider the following feedback control system:

(3.1)\begin{equation} \frac{{\rm d}}{{\rm d}t}v+\nu Av+B(v,v)=h-\mu P_N(v-u), \end{equation}

where $h$, possibly time-dependent, is given. The well-posedness theory and synchronization properties of this model were originally developed in [Reference Azouani, Olson and Titi4] for a more general class of observables, which includes projection onto finitely many Fourier modes as a special case, in [Reference Bessaih, Olson and Titi7, Reference Blömker, Law, Stuart and Zygalakis12] in the setting of noisy observations, while the issue of synchronization in higher-order topologies was studied in [Reference Biswas, Brown and Martinez8, Reference Biswas and Martinez10]. In the idealized setting considered in this article, the existence, uniqueness results in [Reference Azouani, Olson and Titi4] will suffice for our purposes. This is stated in the following theorem.

Theorem 3.1 Let $t_0\geq 0$ and $h\in L^\infty (t_0,\infty ;L^2_\sigma )$. Let $u$ denote the unique solution of (2.3) corresponding to initial data $u_0\in H^1_\sigma$ guaranteed by theorem 2.1. There exists a constant ${\tilde {c}}>0$ such if $\mu,N$ satisfy

(3.2)\begin{equation} \mu \leq {\tilde{c}}\nu N^2 \end{equation}

then given $v_0\in H^1_\sigma$, there exists a unique solution, $v$, to the initial value problem corresponding to (3.1) such that

(3.3)\begin{equation} v\in C([t_0,T];H^1_\sigma)\cap L^2(t_0,T; H^2_\sigma)\quad\text{and}\quad \frac{{\rm d}}{{\rm d}t}v\in L^2(t_0,T;L^2_\sigma), \end{equation}

for any $T>0$.

The feedback control system (3.1) was originally conceived as a way to assimilate observations on the flow field into the equations of motion in order to reconstruct the unobserved scales of motion. There exists a considerable body of research analysing the extent to which this is possible in various situations in hydrodynamics such as Rayleigh–Bénard convection [Reference Altaf, Titi, Gebrael, Knio, Zhao and McCabe3, Reference Cao, Jolly, Titi and Whitehead14, Reference Farhat, Glatt-Holtz, Martinez, McQuarrie and Whitehead32Reference Farhat, Lunasin and Titi35, Reference Farhat, Lunasin and Titi37], turbulence [Reference Albanez, Nussenzveig Lopes and Titi2, Reference Chen, Li and Lunasin19, Reference Clark Di Leoni, Mazzino and Biferale23, Reference Farhat, Lunasin and Titi38, Reference Larios and Pei58, Reference Yu, Giorgini, Jolly and Pakzad67, Reference Zauner, Mons, Marquet and Leclaire68], geophysical fluids [Reference Albanez and Benvenutti1, Reference Desamsetti, Dasari, Langodan, Titi, Knio and Hoteit30, Reference Farhat, Lunasin and Titi36, Reference Jolly, Martinez, Olson and Titi52, Reference Jolly, Martinez and Titi53, Reference Pei64], dispersive equations [Reference Jolly, Sadigov and Titi54, Reference Jolly, Sadigov and Titi55], as well as various numerical analytical and computational studies [Reference Blocher, Martinez and Olson13, Reference Celik, Olson and Titi18, Reference Diegel and Rebholz31, Reference Foias, Mondaini and Titi42, Reference García-Archilla and Novo48Reference Ibdah, Mondaini and Titi51, Reference Larios, Rebholz and Zerfas59].

Given $h$, we may thus obtain from (3.1) an approximate reconstruction of the high modes of $u$ via $Q_Nv=(I-P_N)v$. We therefore propose the following algorithm: let $h=f_0$ denote the initial guess for the forcing field, defined for all $t\geq t_0$, for some fixed initial time $t_0\geq 0$, and let $v^0$ to be an arbitrary initial state; $f_0$ is considered to be the approximation at stage $n=0$ that is prescribed by the user and may, in fact, be chosen arbitrarily. Suppose that $P_Nf_0=f_0$. Then, at stage $n=1$, we consider:

(3.4)\begin{align} \frac{{\rm d}}{{\rm d}t}v_1+\nu Av_1+B(v_1,v_1)=f_0-\mu P_N(v_1-u),\quad v_1(t_0)=v_1^0,\quad t\in I_0:=[t_0,\infty). \end{align}

We define the first approximation to the flow field by

(3.5)\begin{equation} u_1=P_Nu+Q_Nv_1, \quad \text{for}\ t\in I_0. \end{equation}

By design, $u_1$ will relax towards $u$ after a transient period, $\rho _1:=t_1-t_0$, that is proportional to the relaxation timescale, $\mu ^{-1}$, where $t_1\gg t_0$, but only up to an error of size $O(g_0)$, where $g_0:=f_0-f$, which represents the ‘model error’. Only after this period has transpired we will extract the first approximation to $f$ via the formula:

(3.6)\begin{align} f_1(t)& :=\frac{{\rm d}}{{\rm d}t}P_Nu_1+\nu AP_Nu_1+P_NB(u_1,u_1),\nonumber\\ & \quad \text{for all}\ t\in I_1:=[t_0+\rho_1,\infty),\ \text{for some}\ \rho_1\gg 0. \end{align}

Observe that $P_Nf_1=f_1$. To obtain new approximations in subsequent stages, we proceed recursively: suppose that at stage $n-1$, a force, $f_{n-1}=P_Nf_{n-1}$ over $t\in I_{n-1}:=[t_{n-2}+\rho _{n-1},\infty )$, has been produced, where the relaxation period satisfies $\rho _{n-1}\gg 0$, and an arbitrary initial state, $v_n^0$, has been given. Then, consider:

(3.7)\begin{align} \frac{{\rm d}}{{\rm d}t}v_n+\nu Av_n+B(v_n,v_n)=f_{n-1}-\mu P_N(v_n-u),\quad v_n(t_{n-1})=v_n^0,\quad t\in I_{n-1}. \end{align}

Let $u_n=P_Nu+Q_Nv_n$, for $t\in I_{n-1}$, and define the approximation to the force at stage $n$ by

(3.8)\begin{align} f_n(t)& :=\frac{{\rm d}}{{\rm d}t}P_Nu_n+\nu AP_Nu_n+P_NB(u_n,u_n),\nonumber\\ & \quad \text{for all}\ t\in I_n:=[t_{n-1}+\rho_n,\infty),\ \text{for some}\ \rho_n\gg 0. \end{align}

This procedure produces a sequence of forces $f_1|_{I_1}, f_2|_{I_2}, f_3|_{I_3},\dots$ that approximates $f$ on time intervals $I_n$ whose left-hand endpoints are increasing with $n$. In particular, the sequence $\{f_n|_{I_n}\}_n$ asymptotically approximates $f$.

The key step to ensuring convergence of the generated sequence $\{f_n\}_{n\geq 1}$ to the true forcing, $f$, is to control model errors, $g_n:=f_n-f$, at each stage, in terms of the synchronization errors, $w_n$, which, in turn, are controlled by the model error from the previous stage; this will be guaranteed to be the case after transient periods of length $\rho _n:=t_n-t_{n-1}$, which allows relaxation in (3.4) to take place. However, it will be shown that this convergence can only be guaranteed to occur on the ‘observational subspace’ $P_NL^2_\sigma$, for $N$ sufficiently large. Indeed, a crucial observation at this point is that if $f=P_Nf$, then $f_n=P_Nf_n$, for all $N\geq 1$. We refer the reader to remarks 4.6 and 5.2 for additional remarks on the basic expectations for recovering force from low-mode data and the underlying limitations of this algorithm. We refer the reader to remark 7.1 for a more detailed discussion on the size of the transient periods, $\rho _n$.

Remark 3.2 We remark that the first guess, $f_0$, need not be arbitrary and can be chosen to be:

(3.9)\begin{equation} f_0:=\frac{{\rm d}}{{\rm d}t}P_Nu+\nu AP_Nu+P_NB (P_Nu,P_Nu), \end{equation}

as suggested in (1.2), where $t_{0}>0$ is chosen sufficiently large so that $u(t_0)$ is contained in an absorbing ball for (2.3). Similarly, the initial states, $v_n^0$, at each stage need not be arbitrary. Since $\{P_Nu(t)\}_{t\geq t_0}$ is assumed to be given, one can initialize the system governing $v_n$ at time $t=t_{n-1}$ with $v_n^0=P_Nu(t_{n-1})$ at each stage. These natural choices would presumably aid in the implementation of the proposed algorithm; we refer the reader to remark 7.1 for additional remarks related to this point.

Before outlining the proofs, we rigorously state the main results of the article.

4. Statements of main results

Let $A^{1/2}u_0\in \tilde {B}_1$ and $Au_0\in \tilde {B}_2$ and $f\in C([t_0,\infty );L^2_\sigma )\cap L^\infty (t_0,\infty ;L^2_\sigma )$, for some $t_0\geq 0$, where $\tilde {B}_1 ,\tilde {B}_2$ were defined as in § 2. Let $u$ denote the unique strong solution of (2.3) corresponding to forcing $f$ and initial data $u_0$ guaranteed by theorem 2.1 corresponding to initial velocity $u_0$ and external forcing $f$. Given $t_0\geq 0$, let $\gamma _0$ denote the initial relative error defined by

(4.1)\begin{equation} \gamma_0:=\frac{\sup_{t\geq t_0}\lVert f_0(t)-f(t)\rVert_{L^2}}{\sup_{t\geq t_0}\lVert f(t)\rVert_{L^2}}, \end{equation}

where $f_0$ is a user-prescribed initial guess for the force which satisfies $f_0\in C([t_0,\infty );L^2_\sigma )\cap L^\infty (t_0,\infty ;L^2_\sigma )$. Recall that the assumptions on the initial data imply [via (2.10), (2.15)] that $A^{1/2}u(t)\in \tilde {B}_1$ and $Au(t)\in 2\tilde {B}_2$, for all $t\geq 0$.

Theorem 4.1 Suppose that $f=P_Nf$, for some $N\geq 1$, and that $\{P_Nu(t)\}_{t\geq t_0}$ is given. There exists a positive constant $c$, depending on $\gamma _0$, such that for any $\beta \in (0,1)$, if $N\geq 1$ satisfies:

(4.2)\begin{equation} \beta^{2}N^2>c(\tilde{\sigma}_1+\tilde{G})\tilde{G}^2, \end{equation}

then there exists a choice for the tuning parameter $\mu$ and increasing sequence of times $t_{n-1}\leq t_n$, for $n\geq 2$, such that

(4.3)\begin{equation} \sup_{t\geq t_n}\lVert f_n(t)-f(t)\rVert_{L^2}\leq \beta\left(\sup_{t\geq t_{n-1}}\lVert f_{n-1}(t)-f(t)\rVert_{L^2}\right), \end{equation}

for all $n\geq 1$, where $f_n=P_Nf_n$, $f_n\in C([t_0,\infty );L^2_\sigma )\cap L^\infty (t_0,\infty ;L^2_\sigma )$, and each $f_n$ is given by (3.8).

Remark 4.2 It is worth pointing out that (4.2) depends on the unknown forcing $f$. However, it must be emphasized that this condition only depends on $f$ through its size. From a practical perspective, one must always approach the problem of parameter estimation with prior knowledge in hand. In this light, what condition (4.2) indicates is that if one has access to the Grashof number of the flow, for instance through measurement of the Reynolds number (see remark 4.7), then the only prior knowledge on the force that is needed to achieve exact recovery is knowledge of its shape factor ${\tilde {\sigma }}_1$. It moreover indicates that having such knowledge in one's possession can quantitatively inform what balance is needed between the number of observations and the algorithmic parameter, $\mu$, to ensure a full reconstruction of the unknown force.

Remark 4.3 Regarding condition (4.2), we remind the reader that $\tilde {G}$ also depends on viscosity. Thus, the intuition that the number of observations needed should increase as the viscosity decreases or that fewer observations are needed when viscosity is large is reflected in the statement.

Note that when $f$ is time-dependent, theorem 4.1 only asserts recovery of the external force asymptotically in time. However, when the force is time-periodic or time-independent, then theorem 4.1 immediately implies that the external force is eventually recovered; we provide a statement of the time-independent case in the following corollary since the corresponding approximating forces are obtained by evaluating the sequence of approximating forces asserted in theorem 4.1 at certain times.

Corollary 4.4 Suppose that $f=P_Nf$, for some $N\geq 1$, is time-independent, and that $\{P_Nu(t)\}_{t\geq t_0}$ is given. There exists a positive constant $c_0$, depending on $\gamma _0$, such that for any $\beta \in (0,1)$, if $N\geq 1$ satisfies:

(4.4)\begin{equation} \beta^2N^2>c_0(\tilde{\sigma}_1+G)G^2, \end{equation}

then there exists a choice for the tuning parameter $\mu$ and an increasing sequence of times $t_{n-1}\leq t_n$ such that

(4.5)\begin{equation} \lVert f_n-f\rVert_{L^2}\leq \beta\lVert f_{n-1}-f\rVert_{L^2}, \end{equation}

for all $n\geq 1$, where $f_n=P_Nf_n$ and each $f_n$ is given by (3.8) evaluated at $t=t_n$.

In the setting of time-independent forcing, one can in fact ‘recycle’ the data provided that a sufficiently long time series is available. By ‘recycle’ we mean that once an approximation to the force is proposed by the algorithm at a given stage, we may use this proposed forcing to re-run the algorithm over the same time window to produce the subsequent approximation of the force in the following stage, and so on. In this way, the same data set is used over and over in order to generate new approximations to the force.

Theorem 4.5 Let $T>0$. Suppose that $f=P_Nf$, for some $N\geq 1$, is time-independent, and that $\{P_Nu(t)\}_{t\in [0,T)}$ is given, for some $t_0\geq 0$. There exists a positive constant $c_0$, depending on $\gamma _0$, such that for any $\beta \in (0,1)$, if $N\geq 1$ satisfies:

(4.6)\begin{equation} \beta^2N^2>c_0(\tilde{\sigma}_1+G)G^2, \end{equation}

then for each $n\geq 1$, there exists a choice for the tuning parameter $\mu$ and a sequence of times, $t_n>0$, such that if $T>t_0+t_n$, for all $n\geq 1$, then

(4.7)\begin{equation} \lVert f_n-f\rVert_{L^2}\leq \beta\lVert f_{n-1}-f\rVert_{L^2}, \end{equation}

for all $n\geq 1$, where $f_n=P_Nf_n$, $f_n\in L^2_\sigma$, and each is determined by a procedure similar to that in § 3, except that $v_n, u_n, f_n$ are always derived on the interval $[0,T)$.

The proof of theorem 4.5 follows along the same lines as that of theorem 4.1, except that one requires a few technical modifications of the set-up described in § 3; we supply the relevant details of these modifications in § 7.

Remark 4.6 A well-known example given by Marchioro [Reference Marchioro60] exhibits a scenario where observation of low modes below the spectrum of the forcing are sufficient to determine the forcing. In particular, the study [Reference Marchioro60] identifies a class of low-mode body forces for which the asymptotic behaviour of solutions to (2.3) are characterized by a one-point global attractor for (2.3) whose unique stationary point is supported on a frequency shell strictly smaller than that of the force. Of course, this phenomenon is made possible due to the presence of nonlinearity in (2.3).

The extent to which it is possible to determine features of the forcing beyond the scales which are observed is, in general, not known and in particular, not addressed by theorem 4.5. Indeed, condition (4.6) identifies an upper bound on the number of modes that one should observe in order to uniquely determine the forcing provided that scales beyond those that are directly observed are not forced to begin with. In particular, it would be interesting to (1) study the sharpness of condition (4.6), and (2) whether one may identify classes of forces, beyond the Marchioro class, which inject energy into scales larger than those which are directly observed, preferably much larger, but can nevertheless be reconstructed by these observations. These issues are left to be investigated in a future study.

Remark 4.7 There are at least three natural directions that warrant further investigation. First and foremost, a systematic computational study that probes the efficacy and limitations of this method for recovering the force should be carried out, especially in the context of noisy observations. For instance, each of the above theorems suggests that a large number of modes are required to achieve convergence. In the context of a turbulent flow, $G\sim \text {Re}^2$, where $\text {Re}$ denotes the Reynolds number of the flow [Reference Dascaliuc, Foias and Jolly28, Reference Dascaliuc, Foias and Jolly29], which may be intractably large in practice. On the contrary, the analysis performed here inherently takes into account ‘worst-case’ scenarios that may saturate various inequalities that were used, but which may occur rarely in reality. This direction will be explored in future research.

Secondly, the choice of Fourier modes as the form of observations is chosen due to its conceptual and analytical convenience. In principle, other observations on the velocity, such as nodal values or local spatial averages can also be used. However, the choice of Fourier modes allows one to commute the ‘observation operator’, $P_N$ with derivatives, which greatly facilitates the analysis. The failure of this commutation introduces further analytical difficulties. Indeed, the original feedback control system was introduced with a general interpolant observable operator, $I_h$, replacing $P_N$ in (3.1). In connection to this, the reader is referred to the classical studies [Reference Cockburn, Jones and Titi24, Reference Cockburn, Jones and Titi25, Reference Foias and Temam46, Reference Jones and Titi56, Reference Jones and Titi57], where the notion of ‘finite determining parameters’, properly generalizing ‘determining modes’, was developed.

Thirdly, although the existence of an inertial manifold for the 2D NSE is an open problem, there are several systems, which do possesses inertial manifolds [Reference Foias, Sell and Temam45, Reference Temam66], such as the Kuramoto–Sivashinsky equation [Reference Foias, Nicolaenko, Sell and Temam43]; it would be interesting to explore what can be gained in this particular context. In a similar vein, it would also be interesting to explore the usage of approximate inertial manifolds for the 2D NSE, as it is used, for instance, in post-processing Galerkin methods. Indeed, these ideas have been successfully used in the context of downscaling data assimilation algorithms in [Reference Mondaini and Titi62].

5. Outline of the convergence argument

To prove theorem 4.1, the object of interest will be the error in the forcing, which we also refer to as ‘model error’. This is denoted by

(5.1)\begin{equation} g_n:=f_n-f, \end{equation}

where $f_n$ is generated from the scheme described in § 3. We claim the following: there exists a sequence of times $t_n\geq t_{n-1}$, for all $n\geq 1$, such that

(5.2)\begin{equation} \sup_{t\geq t_n}\lVert A^{1/2} g_n(t)\rVert_{L^2}\leq \beta\sup_{t\geq t_{n-1}}\lVert A^{1/2} g_{n-1}(t)\rVert_{L^2}, \end{equation}

for some $\beta \in (0,1)$. The lengths of the relaxation periods, $\rho _n:=t_n-t_{n-1}$, between the moments, $t_{n-1}, t_n$, at which we choose to reconstitute new forces, $f_{n-1}, f_n$, respectively, are prescribed to be sufficiently large in order to allow time for system (3.7) to relax. As we will see, the length of these relaxation periods, $\rho _n$, will essentially be determined by the relaxation parameter $\mu$; the reader is referred to remark 7.1 for a precise relation.

Let us denote the synchronization error by

\[ w_n:=v_n-u. \]

Observe that the evolution of $w_n$ over the time interval $[t_{n-1},\infty )$ is governed by

(5.3)\begin{align} \frac{{\rm d}}{{\rm d}t}w_n+\nu Aw_n+B(w_n,w_n)+DB(u)w_n& =g_{n-1}-\mu P_Nw_n, \nonumber\\ w_n(t_{n-1})& =v_n^0-u(t_{n-1}), \end{align}

where $DB(u)v:=B(u,v)+B(v,u)$. Let $B_N=P_NB$ and $DB_N=P_NDB$. Then, using (2.3), the facts that $P_Nf=f$ and $u_n=P_Nu+Q_Nv_N$, and (3.8), we obtain

(5.4)\begin{align} g_n& =\left(\frac{{\rm d}}{{\rm d}t}P_Nu_n+\nu AP_Nu_n+B_N(u_n,u_n)\right)-\left(\frac{{\rm d}}{{\rm d}t}P_Nu+\nu AP_Nu+B_N(u,u)\right)\nonumber\\ & =B_N(P_Nu+Q_Nv_n,P_Nu+Q_Nv_n)-B_N(P_Nu+Q_Nu,P_Nu+Q_Nu)\nonumber\\ & =B_N(Q_Nw_n,Q_Nw_N)+DB_N(u)Q_Nw_n, \end{align}

which holds over $[t_{n-1},\infty )$. In analogy to (1.5), we define the ‘Reynolds stress’ at stage $n$ by

(5.5)\begin{equation} \mathcal{R}_{N}^{(n)}:=B_N(Q_Nw_n,Q_Nw_n)+DB_N(u)Q_Nw_n=g_n. \end{equation}

Ultimately, (5.5) enables us to envision a recursion in the model error at each stage through the dependence of $\mathcal {R}_N^{(n)}$ on $\mathcal {R}_N^{(n-1)}$ via the synchronization error $w_n$. Hence, in order to prove that $\mathcal {R}_{N}^{(n)}$ vanishes in the limit as $n\rightarrow \infty$, we require sensitivity-type estimates, that is, estimates on $w_n$. The estimates will take on the following form:

(5.6)\begin{equation} \lVert A^{1/2} w_n(t)\rVert_{L^2}\leq \nu\left(\frac{\nu}{\mu}\right)^{1/2}O\left(\frac{\sup_{t\geq t_n}\lVert g_{n-1}(t)\rVert_{L^2}}{\nu^2}\right), \end{equation}

for all $t\geq t_n$, for some sufficiently large $t_n\geq t_{n-1}$. Before we go on to develop estimates of form (5.6) in § 6, let us first determine the precise manner in which their application will arise.

Remark 5.1 In what follows and for the remainder of the paper, we make use of the convention that $C$ denotes a generic dimensionless constant, which may change line-to-line and possibly be large, but will always be independent of $N,\nu, \gamma _0$.

To estimate (5.5), we will invoke the following inequalities for $B(u,v)$, which follows from a direct application of Hölder's inequality:

(5.7)\begin{equation} \lVert B(u,v)\rVert_{L^2}\leq \min\{\lVert u\rVert_{L^\infty}\lVert A^{1/2}v\rVert_{L^2},\lVert u\rVert_{L^4}\lVert A^{1/2}v\rVert_{L^4}\}. \end{equation}

We will also make use of the fact that $P_N, Q_N$ commute with $A^{m/2}$, for all integers $m$, and that the following inequalities hold for all $N>0$, $\ell >0$, and $m\in \mathbb {Z}$:

(5.8)\begin{equation} \lVert A^{m/2}P_Nv\rVert_{L^2}\leq N^m\lVert P_Nv\rVert_{L^2},\quad \lVert A^{m/2}Q_Nv\rVert_{L^2}\leq N^{-\ell}\lVert A^{(m+\ell)/2}Q_Nv\rVert_{L^2}. \end{equation}

Lastly, we make the following elementary, but important observation for treating the first term appearing in (5.5): for vector fields $u=(u_1,u_2)$ and $v=(v_1,v_2)$, we have:

\[ B(u,v)_i=\sum_{j=1}^2P(u\cdotp\nabla v_i)=P\nabla\cdotp(uv_i),\quad i=1,2, \]

where $B(u,v)_i$ denotes the $i$-th component of the vector field $B(u,v)$. Thus, upon applying (5.8) and Hölder's inequality, we obtain:

(5.9)\begin{equation} \lVert B_N(u,v)\rVert_{L^2}^2\leq2\sum_{i=1}^2\lVert P_NA^{1/2}(uv_i)\rVert_{L^2}^2\leq 2N^2\sum_{i=1}^2\lVert uv_i\rVert_{L^2}^2\leq 2N^2\lVert u\rVert_{L^4}^2\lVert v\rVert_{L^4}^2. \end{equation}

From (5.5), we now apply (5.9), (5.7), (5.8), and interpolation to obtain

(5.10)\begin{align} \lVert \mathcal{R}_{N}^{(n)}\rVert_{L^2}& \leq \lVert P_NA^{1/2}(Q_Nw_n\otimes Q_Nw_n)\rVert_{L^2}+\lVert DB_N(u)Q_Nw_n\rVert_{L^2}\nonumber\\ & \leq CN\lVert Q_Nw_n\rVert_{L^4}^2+\lVert u\rVert_{L^\infty}\lVert A^{1/2} Q_Nw_n\rVert_{L^2}+\lVert Q_Nw_n\rVert_{L^4}\lVert A^{1/2} u\rVert_{L^4}\nonumber\\ & \leq CN\lVert A^{1/2}Q_N w_n\rVert_{L^2}\lVert Q_Nw_n\rVert_{L^2}+C\lVert Au\rVert_{L^2}^{1/2}\lVert u\rVert_{L^2}^{1/2}\lVert A^{1/2} Q_Nw_n\rVert_{L^2}\nonumber\\ & \quad+\lVert A^{1/2} Q_Nw_n\rVert_{L^2}^{1/2}\lVert Q_Nw_n\rVert_{L^2}^{1/2}\lVert Au\rVert_{L^2}^{1/2}\lVert A^{1/2} u\rVert_{L^2}^{1/2}\nonumber\\ & \leq C\lVert A^{1/2}Q_N w_n\rVert_{L^2}^2+C\lVert Au\rVert_{L^2}^{1/2}\lVert u\rVert_{L^2}^{1/2}\lVert A^{1/2} Q_Nw_n\rVert_{L^2}\nonumber\\ & \quad+\frac{C}{N^{1/2}}\lVert A^{1/2} Q_Nw_n\rVert_{L^2}\lVert Au\rVert_{L^2}^{1/2}\lVert A^{1/2} u\rVert_{L^2}^{1/2}\nonumber\\ & \leq C_0\nu(\tilde{\sigma}_1+\tilde{G})^{1/2}\tilde{G}\left(1+\frac{\lVert A^{1/2} Q_Nw_n\rVert_{L^2}}{\nu}\right)\lVert A^{1/2} Q_Nw_n\rVert_{L^2}, \end{align}

for some universal constant $C_0>0$, independent of $n$, for all $t\geq t_{n-1}$. Note that we also invoked the assumption that $u$ belongs to the absorbing ball in $H^1_\sigma$ and $H^2_\sigma$ [see (2.11), (2.14), respectively] in obtaining the final inequality. It is at this point that one applies (5.6) in order to properly close the recursive estimate. In order to rigorously carry out this argument, let us therefore prove that (5.6) indeed holds.

Remark 5.2 In the case when the force contains modes beyond those that are observed, one can identify an obstruction that precludes a proof in the manner described above. Suppose that $f\neq P_Nf$. Then, modify the ansatz (3.8) for the force at each stage by removing the projection onto low modes. In particular, re-define $f_n$ so that:

\[ f_n=\frac{{\rm d}}{{\rm d}t}u_n+\nu Au_n+B(u_n,u_n). \]

Then, the model error becomes:

(5.11)\begin{align} g_n=\frac{{\rm d}}{{\rm d}t}Q_Nw_n+\nu AQ_Nw_n+B(Q_Nw_n,Q_Nw_n)+DB(u)Q_Nw_n. \end{align}

Upon applying the complementary projection, $Q_N$, to (5.3), and combining the result with (5.11), we see that:

\begin{align*} Q_Ng_n& ={-}B^N(P_Nw_n,P_Nw_n)-B^N(P_Nw_n,Q_Nw_n)-B^N(Q_Nw_n,P_Nw_n)\\ & \quad -DB^N(u)P_Nw_n+Q_Ng_{n-1}, \end{align*}

where $B^N=Q_NB$ and $DB^N=Q_NDB$. Due to the presence of $Q_Ng_{n-1}$ on the right-hand side, one cannot expect to obtain a convergent recursive relation of the form $\lVert Q _Ng_n\rVert _{L^2}\leq \beta \lVert Q _Ng_{n-1}\rVert _{L^2}$, for some $\beta \in (0,1)$. Although one can establish an estimate of $Q_Ng_n$ from this relation that is of the form $\lVert Q _Ng_n-Q_Ng_{n-1}\rVert _{L^2}\leq O_N(\lVert A ^{1/2}Q_Nw_n\rVert _{L^2})$, where the suppressed constant has a favourable dependence on $N$, a subsequent analysis will nevertheless be unable to establish a suitable recursion relation since we will only ever have access to an estimate of the form (5.6).

6. Sensitivity analysis

We establish a more precise form of the crucial estimates (5.6), which form the bridge to the desired recursion for the model error at each stage of the approximation. For this, we recall the notation introduced in § 5. In particular, we prove the following.

Proposition 6.1 There exist universal constants $\overline {c}_0, \underline {c}_0\geq 1$ such that if $\mu, N$ satisfy

(6.1)\begin{equation} \underline{c}_0\big(\tilde{\sigma}_1+\tilde{G}\big)\tilde{G}^2\leq \frac{\mu}{\nu}\leq \overline{c}_0 N^2, \end{equation}

then for each $n\geq 1$, there exist relaxation periods, $\rho _n=t_n-t_{n-1}$, for some $t_n>t_{n-1}$, such that

(6.2)\begin{equation} \sup_{t\in I_n}\left(\frac{\lVert A^{1/2} w_n(t)\rVert_{L^2}}{\nu}\right) \leq \left(\frac{2C_1\nu}{\mu}\right)^{1/2}\left(\frac{\sup_{t\in I_{n-1}}\lVert g_{n-1}(t)\rVert_{L^2}}{\nu^2}\right), \end{equation}

where $I_n:=[t_{n-1}+\rho _n,\infty )$, for some universal constant $C_1\geq 1$, independent of $n$. Moreover, $\rho _n$ satisfies (6.5).

Proof. Fix $n\geq 1$. The enstrophy-balance for $w_n$ is obtained by taking the $L^2$-inner product of (5.3) by $Aw_n$, which yields:

(6.3)\begin{align} \frac{1}2\frac{{\rm d}}{{\rm d}t}\lVert A^{1/2} w_n\rVert_{L^2}^2+\nu\lVert Aw_n\rVert_{L^2}^2& ={-}\langle B(u,w_n),Aw_n\rangle-\langle B(w_n,u),Aw_n\rangle\nonumber\\ & \quad +\langle g_{n-1},Aw_n\rangle-\mu \lVert A^{1/2} P_Nw_n\rVert_{L^2}^2\nonumber\\ & =I+II+III+IV. \end{align}

Observe that by interpolation, the Cauchy–Schwarz inequality, Poincaré's inequality, and (2.11), (2.14), we may estimate:

\begin{align*} |I|& \leq C\lVert u\rVert_{L^\infty}\lVert A^{1/2} w_n\rVert_{L^2}\lVert Aw_n\rVert_{L^2}\\ & \leq C\lVert Au\rVert_{L^2}^{1/2}\lVert u\rVert_{L^2}^{1/2}\lVert A^{1/2} w_n\rVert_{L^2}\lVert Aw_n\rVert_{L^2}\\ & \leq C\mu\left(\frac{\nu}{\mu}\right)\big(\tilde{\sigma}_1+\tilde{G}\big)\tilde{G}^2\lVert A^{1/2} w_n\rVert_{L^2}^2+\frac{\nu}{100}\lVert Aw_n\rVert_{L^2}^2. \end{align*}

Similarly, we obtain:

\begin{align*} |II| & \leq C\lVert A^{1/2} w_n\rVert_{L^2}^{1/2}\lVert w_n\rVert_{L^2}^{1/2}\lVert Au\rVert_{L^2}^{1/2}\lVert A^{1/2}u\rVert_{L^2}^{1/2}\lVert Aw_n\rVert_{L^2}\\ & \leq C\mu\left(\frac{\nu}{\mu}\right)\big(\tilde{\sigma}_1+\tilde{G}\big)\tilde{G}^2\lVert A^{1/2} w_n\rVert_{L^2}^2+\frac{\nu}{100}\lVert Aw_n\rVert_{L^2}^2. \end{align*}

For the third term, we apply the Cauchy–Schwarz inequality to obtain:

\[ |III|\leq C\frac{\lVert g_{n-1}\rVert_{L^2}^2}{\nu}+\frac{\nu}{100}\lVert Aw_n\rVert_{L^2}^2. \]

Lastly, we have:

\begin{align*} IV& ={-}\mu\lVert A^{1/2} w_n\rVert_{L^2}^2+\mu\lVert A^{1/2} Q_Nw_n\rVert_{L^2}^2\\ & \leq{-}\mu\lVert A^{1/2} w_n\rVert_{L^2}^2+\frac{\mu}{N^2}\lVert Aw_n\rVert_{L^2}^2. \end{align*}

Upon combining the estimates for $I$$IV$ and invoking (6.1), where $\underline {c}_0, \overline {c}_0\geq 1$ are chosen appropriately relative to the constants $C$ appearing above, we arrive at:

\[ \frac{{\rm d}}{{\rm d}t}\lVert A^{1/2} w_n\rVert_{L^2}^2+\frac{3}2\nu\lVert Aw_n\rVert_{L^2}^2+\frac{3}2\mu\lVert A^{1/2} w_n\rVert_{L^2}^2\leq C\nu^3\left(\frac{\lVert g_{n-1}\rVert_{L^2}}{\nu^2}\right)^2. \]

An application of Gronwall's inequality yields:

(6.4)\begin{align} \left(\frac{\lVert A^{1/2} w_n(t)\rVert_{L^2}}{\nu}\right)^2 & \leq {\rm e}^{-\mu(t-t_n)}\,{\rm e}^{-\mu\rho_n}\left(\frac{\lVert A^{1/2} w_n(t_{n-1})\rVert_{L^2}}{\nu}\right)^2\nonumber\\ & \quad + C_1\left(\frac{\nu}{\mu}\right)\left(\frac{\sup_{t\geq t_{n-1}}\lVert g_{n-1}(t)\rVert_{L^2}}{\nu^2}\right)^2, \end{align}

for all $t\geq t_n$, for some $C_1\geq 1$ independent of $n$. Now, recall that $w_n(t_{n-1})=v^0_n-u(t_{n-1})$. We choose $\rho _n>0$ such that:

(6.5)\begin{equation} \rho_n\geq\frac{1}{\mu}\ln\left[\left(\frac{\mu}{C_1\nu}\right)\left(\frac{\nu\lVert A^{1/2} w_n(t_{n-1})\rVert_{L^2}}{\sup_{t\in I_{n-1}}\lVert g_{n-1}(t)\rVert_{L^2}}\right)^2\right]. \end{equation}

Returning to (6.4), it follows that

\[ \sup_{t\in I_n}\left(\frac{\lVert A^{1/2} w_n(t)\rVert_{L^2}}{\nu}\right) \leq \left(\frac{2C_1\nu}{\mu}\right)^{1/2}\left(\frac{\sup_{t\in I_{n-1}}\lVert g_{n-1}(t)\rVert_{L^2}}{\nu^2}\right), \]

as desired.

7. Proofs of convergence

Let us assume that $f=P_Nf$. At the initializing stage, $n=0$, we consider any $f_0\in C([0,\infty );L^2_\sigma )\cap L^\infty ([0,\infty ;);L^2_\sigma )$. Recall that in each subsequent stage, we will produce a new approximation, $f_n$, for the force via the ansatz (3.8). We proceed by induction.

Proof Proof of theorem 4.1

Fix $\beta \in (0,1)$. We choose $c_0$ to satisfy:

\[ c_0\geq\max\left\{\underline{c}_0, 2C_1\left(\frac{C_0}{\beta}\right)^2\left[1+\left(\frac{2C_1}{\underline{c}_0(\tilde{\sigma}_1+\tilde{G})}\right)^{1/2}\gamma_0\right]^2\right\}, \]

where $C_0, C_1, \underline {c}_0$ are the universal constants appearing in (5.10), (6.1), (6.2), respectively, and $\gamma _0$ denotes the initial relative error defined by (4.1). Fix any $c_1\leq {\tilde {c}}$, where ${\tilde {c}}$ is the constant appearing in (3.2). Then, assume that $\mu, N$ satisfies (4.2) with $c=c_0c_1^{-1}$. Then, choose $\mu$ such that:

(7.1)\begin{equation} c_0(\tilde{\sigma}_1+\tilde{G})\tilde{G}^2<\beta^2\frac{\mu}{\nu}\leq c_1 \beta^2 N^2. \end{equation}

Observe that $N$ satisfies (4.2) with $c=c_0c_1^{-1}$. Since $c_0\geq \underline {c}_0$ by choice, it immediately follows that (6.1) also holds.

Let $n=1$. Denote $\mathcal {R}_N^{(0)}:=g_0=f_0-f$. Observe that from (2.9), (4.1), we have:

(7.2)\begin{equation} \sup_{t\geq t_0}\lVert \mathcal{R}_N^{(0)}(t)\rVert_{L^2}=\gamma_0\nu^2\tilde{G}. \end{equation}

Combining proposition 6.1, (5.10) for $n=1$, and (7.2) ensures that there exists a relaxation period $\rho _1>0$, for some $t_1\geq t_0$, such that

(7.3)\begin{align} \sup_{t\geq I_1}\lVert \mathcal{R}_N^{(1)}(t)\rVert_{L^2}& \leq (2C_1)^{1/2}C_0(\tilde{\sigma}_1+\tilde{G})^{1/2}\tilde{G}\left(1+\left(\frac{2C_1\nu}{\mu}\right)^{1/2}\gamma_0\tilde{G}\right)\left(\frac{\nu}{\mu}\right)^{1/2}\nonumber\\ & \quad \sup_{t\geq I_0}\lVert \mathcal{R}_N^{(0)}(t)\rVert_{L^2}, \end{align}

where $I_1=[t_0+\rho _1,\infty )$.

Further assume that $\mu$ satisfies:

(7.4)\begin{equation} \beta^2\frac{\mu}{\nu}\geq 2C_1C_0^2\left((\tilde{\sigma}_1+\tilde{G})^{1/2}+\left(\frac{2C_1}{\underline{c}_0}\right)^{1/2}\gamma_0\right)^2\tilde{G}, \end{equation}

where $C_1$ is the same constant appearing in (7.3). Then, (7.3), (6.1), and (7.4) imply:

(7.5)\begin{equation} \sup_{t\geq I_1}\lVert \mathcal{R}_N^{(1)}(t)\rVert_{L^2}\leq \beta\left(\sup_{t\geq I_0}\lVert \mathcal{R}_N^{(0)}(t)\rVert_{L^2}\right). \end{equation}

This establishes the base case.

Now, suppose that for all $k=1,\ldots, n$, there exist relaxation periods $\rho _k$, such that:

(7.6)\begin{equation} \sup_{t\geq I_k}\lVert \mathcal{R}_N^{(k)}(t)\rVert_{L^2}\leq \beta\sup_{t\geq I_{k-1}}\lVert \mathcal{R}_N^{(k-1)}(t)\rVert_{L^2}. \end{equation}

With (7.2), it follows that:

(7.7)\begin{equation} \sup_{t\geq I_k}\lVert \mathcal{R}_N^{(k)}(t)\rVert_{L^2}\leq \beta^k\gamma_0\tilde{G}\leq \gamma_0\tilde{G}, \end{equation}

for all $k=1,\dots, n$. We may thus deduce from proposition 6.1, (5.10), and (7.7) that

(7.8)\begin{align} \sup_{t\geq I_{n+1}}\lVert \mathcal{R}_N^{(n+1)}(t)\rVert_{L^2}& \leq (2C_1)^{1/2}C_0(\tilde{\sigma}_1+\tilde{G})^{1/2}\tilde{G}\left(1+\left(\frac{2C_1\nu}{\mu}\right)^{1/2}\gamma_0\tilde{G}\right)\left(\frac{\nu}{\mu}\right)^{1/2}\nonumber\\ & \quad \sup_{t\geq I_n}\lVert \mathcal{R}_N^{(n)}(t)\rVert_{L^2}. \end{align}

An application of (7.4) and (6.1) then implies

\[ \sup_{t\geq I_{n+1}}\lVert \mathcal{R}_N^{(n+1)}(t)\rVert_{L^2}\leq \beta\left(\sup_{t\geq I_n}\lVert \mathcal{R}_N^{(n)}(t)\rVert_{L^2}\right), \]

as desired. This completes the induction.

Remark 7.1 From the proof of theorem 4.1, we can obtain more informative estimates on the relaxation periods, $\rho _n$. Indeed, by (6.5) and (7.2), we see that the first relaxation period satisfies:

(7.9)\begin{equation} \rho_1\geq \frac{1}{\mu}\ln\left[\left(\frac{\mu}{C_1\nu}\right)\frac{1}{\gamma_0^2\tilde{G}^2}\left(\frac{\lVert A^{1/2}(v^0_1-u(t_0))\rVert_{L^2}}{\nu}\right)^2\right]. \end{equation}

In subsequent stages, $n\geq 1$, by (6.5) and (6.2) applied at the preceding stage, we may deduce that the relaxation periods satisfy:

(7.10)\begin{equation} \rho_n\geq \frac{1}{\mu}\ln\left(\frac{\sup_{t\in I_{n-2}}\lVert \mathcal{R}_N^{(n-2)}(t)\rVert_{L^2}}{\sup_{t\in I_{n-1}}\lVert \mathcal{R}_N^{(n-1)}(t)\rVert_{L^2}}\right)\geq{-}\frac{1}{\mu}\ln\beta. \end{equation}

One may then observe that (7.9) and (7.10) follow a consistent pattern if we allow ourselves to make special, but nevertheless natural choices for initial force ansatz, $f_0$, and the initial data, $v_1^0$, of the first nudged system. Indeed, suppose that the solution, $u$, of (2.3) possesses additional regularity, for instance, $Au(t_0)\in L^2_\sigma$. Now, initialize the nudged system (3.4) with $v_1^0=P_Nu(t_0)$ and let the initial guess $f_0$ be given by (1.2). Then, the first relaxation period satisfies:

(7.11)\begin{equation} \rho_1\geq \frac{1}{\mu}\ln\left[\left(\frac{c_1}{C_1}\right)\left(\frac{\nu^2}{\sup_{t\in I_0}\lVert \mathcal{R}_N^{(0)}(t)\rVert_{L^2}}\right)^2\right], \end{equation}

where we applied the Poincaré inequality, the upper bound in (7.1), and (2.9).

Now, we prove a variation on the time-independent case, which allows one to recycle the existing data.

Proof Proof sketch of theorem 4.5

We modify the algorithm in §3 as follows: let $J_0=I_0=[t_0,\infty )$. For convenience, we assume that $t_0=0$. Suppose that $\{P_Nu(t)\}_{t\in J_0}$ is known. At stage $n=1$, we suppose initial data $v_1^0$ is given and solve (3.4) to produce the solution $v_1$ over $J_0$. We then define (3.5) over $J_0$ and immediately generate $f_1(t)$ via (3.6) over $J_0$. To define the stage $1$ approximation to $f$, we evaluate $f_1(t)$ after a transient period of length $\rho _1>0$ to obtain $f_1:=f_1(\rho _1)$.

In subsequent stages $n\geq 1$, we generate $v_n$ over $J_0$ via (3.7), where $f_{n-1} :=f_{n-1}(\rho _{n-1})$ was generated from the preceding stage, $n-1$. We then define the stage-$n$ approximation to the true state over $J_0$ by $u_n=P_Nu+Q_Nv_n$. We generate a new force, $f_n(t)$, via the ansatz (3.8). The stage-$n$ approximation of $f$ is then given by $f_n(\rho _n)$, for some $\rho _n>0$.

To assess the error, we once again form the synchronization error, $w_n=v_n-u$, and the model error, $g_n(t)=f_n(t)-f$. For each $n\geq 1$, the identity (5.5) still holds. Consequently, (5.10) holds as well. The remaining ingredient to establish convergence is a time-independent force analogue of proposition 6.1.

Supposing that (6.1) holds with $\tilde {G}\mapsto G$, the proof of proposition 6.1 proceeds the same way, except that the analysis is carried out entirely over the interval $J_0$, to arrive at the analogue of (6.4):

(7.12)\begin{align} \left(\frac{\lVert A^{1/2} w_n(\rho_n)\rVert_{L^2}}{\nu}\right)^2 \leq {\rm e}^{-\mu\rho_n}\left(\frac{\lVert A^{1/2} w_n(0)\rVert_{L^2}}{\nu}\right)^2+ C_1\left(\frac{\nu}{\mu}\right)\left(\frac{\lVert g_{n-1}\rVert_{L^2}}{\nu^2}\right)^2. \end{align}

One chooses $\rho _n>0$ according to:

\[ \rho_n\geq \frac{1}{\mu}\left[\frac{\mu}{C_1\nu}\left(\frac{\nu\lVert A^{1/2}w_n(0)\rVert_{L^2}}{\lVert g_{n-1}\rVert_{L^2}}\right)^2\right], \]

so that (7.12) yields:

\[ \left(\frac{\lVert A^{1/2} w_n(\rho_n)\rVert_{L^2}}{\nu}\right)^2 \leq 2C_1\left(\frac{\nu}{\mu}\right)\left(\frac{\lVert g_{n-1}\rVert_{L^2}}{\nu^2}\right)^2. \]

It is at this time $t=\rho _n$, that we evaluate (3.8).

After these adjustments, it is now clear that the proof of theorem 4.5 follows an analogous manner to the proof theorem 4.1, mutatis mutandis.

Acknowledgements

The author graciously acknowledges the generosity and support of the ADAPT group, as well as efforts of the referees in reviewing this manuscript. Support for this project was also provided by the National Science Foundation through NSF-DMS 2213363 and NSF-DMS 2206491, as well as PSC-CUNY Award 65187-00 53, which is jointly funded by The Professional Staff Congress and The City University of New York.

Appendix A. Uniform-in-time estimates for palenstrophy

We now provide uniform-in-time estimates in $H^2$ for the reference flow field when the external force field is time-dependent. We begin by establishing an alternative form of the standard enstrophy balance of (2.3) as it is presented in theorem 2.1. Indeed, upon taking the $L^2$-inner product of (2.3) with $Au$, we obtain:

\[ \frac{1}2\frac{{\rm d}}{{\rm d}t}\lVert A^{1/2}u\rVert_{L^2}^2+\nu\lVert Au\rVert_{L^2}^2=\langle f,Au\rangle. \]

By the Cauchy–Schwarz inequality and (2.9), we see that:

\[ |\langle f,Au\rangle|\leq \nu^3\tilde{G}^2+\frac{\nu}{4}\lVert Au\rVert_{L^2}^2. \]

By the Poincaré inequality, it follows that:

\[ \frac{{\rm d}}{{\rm d}t}\left({\rm e}^{\nu t}\lVert A^{1/2}u\rVert_{L^2}^2\right)+\frac{\nu}2\lVert Au\rVert_{L^2}^2\leq 2\nu^3\tilde{G}^2\,{\rm e}^{\nu t}. \]

Integrating over $t\geq t_0$ yields:

\begin{align*} & \lVert A^{1/2}u(t)\rVert_{L^2}^2+\frac{\nu}2\int_{t_0}^t\,{\rm e}^{-\nu(t-s)}\lVert Au(s)\rVert_{L^2}^2\\ & \quad \leq \lVert A^{1/2}u(t_0)\rVert_{L^2}^2\,{\rm e}^{-\nu(t-t_0)}+2\nu^2\tilde{G}^2(1-{\rm e}^{-\nu(t-t_0)}), \end{align*}

which is (2.10), as desired. Thus, for $u_0\in {\tilde {B}}_1$, we may deduce:

(A.1)\begin{equation} \lVert A^{1/2}u(t)\rVert_{L^2}^2+\frac{\nu}2\int_{t_0}^t\,{\rm e}^{-\nu(t-s)}\lVert Au(s)\rVert_{L^2}^2\leq 2\nu^2\tilde{G}^2, \end{equation}

for all $t\geq t_0\geq 0$.

Proof Proof of (2.12)

Upon taking the $L^2$-inner product of (2.3) with $A^2u$, we obtain:

\[ \frac{1}2\frac{{\rm d}}{{\rm d}t}\lVert Au\rVert_{L^2}^2+\nu\lVert A^{3/2}u\rVert_{L^2}^2={-}\langle B(u,u),A^2u\rangle+\langle f,A^2u\rangle=I+II. \]

Observe that integration by parts multiple times yields:

\[ I ={-}\langle B(Au,u),Au\rangle-2\sum_{\ell=1,2}\langle B(\partial_\ell u,\partial_\ell u),Au\rangle \]

where we applied the identity $\langle B(u,v),v\rangle =0$. It then follows from Hölder's inequality, interpolation, Young's inequality, and (2.10) that:

\begin{align*} |I|& \leq 3\lVert A^{1/2}u\rVert_{L^2}\lVert Au\rVert_{L^4}^2\leq 3c_L\lVert A^{1/2} u\rVert_{L^2}\lVert Au\rVert_{L^2}\lVert A^{3/2}u\rVert_{L^2}\\ & \leq \frac{\nu}{8}\lVert A^{3/2}u\rVert_{L^2}^2+36c_L^2\nu\tilde{G}^2\lVert Au\rVert_{L^2}^2,\end{align*}

where $c_L\geq 1$ is the associated constant of interpolation, i.e. $\lVert Au \rVert _{L^4}^2\leq c_L\lVert Au \rVert _{L^2}\lVert A ^{3/2}u\rVert _{L^2}$. We treat $II$ with integration by parts and the Cauchy–Schwarz inequality to obtain:

\begin{align*} |II|& \leq|\langle A^{1/2} f,A^{3/2} u\rangle|\leq 2\nu^3\left(\frac{\lVert A^{1/2} f\rVert_{L^\infty_tL^2_x}}{\nu^2}\right)^2+\frac{\nu}{8}\lVert A^{3/2}u\rVert_{L^2}^2\\ & \leq 2\nu^3\tilde{\sigma}_1^2\tilde{G}^2+\frac{\nu}{8}\lVert A^{3/2}u\rVert_{L^2}^2. \end{align*}

We now combine the estimates $I,II$, and apply Poincaré's inequality to arrive at:

\begin{align*} \frac{{\rm d}}{{\rm d}t}\lVert Au\rVert_{L^2}^2+\nu\lVert Au\rVert_{L^2}^2+\frac{\nu}2\lVert A^{3/2}u\rVert_{L^2}^2\leq 36c_L^2\nu\tilde{G}^2\lVert Au\rVert_{L^2}^2+2\nu^3\tilde{\sigma}_1^2\tilde{G}^2. \end{align*}

An application of Gronwall's inequality yields:

\begin{align*} & \lVert Au(t)\rVert_{L^2}^2+\frac{\nu}2\int_{t_0}^t{\rm e}^{-\nu(t-s)}\lVert A^{3/2}u(s)\rVert_{L^2}^2\,{\rm d}s\\ & \quad\leq \lVert Au(t_0)\rVert_{L^2}^2\,{\rm e}^{-\nu(t-t_0)}+36c_L^2\nu\tilde{G}^2\int_{t_0}^t\,{\rm e}^{-\nu(t-s)}\lVert Au(s)\rVert_{L^2}^2\,{\rm d}s\\ & \quad +2\nu^2\tilde{\sigma}_1^2\tilde{G}^2(1-{\rm e}^{-\nu(t-t_0)}), \end{align*}

for all $t \geq t_0\geq 0$. We now apply (A.1) to bound:

\begin{align*} & \lVert Au(t)\rVert_{L^2}^2+\frac{\nu}2\int_{t_0}^t\,{\rm e}^{-\nu(t-s)}\lVert A^{3/2}u(s)\rVert_{L^2}^2\,{\rm d}s\\ & \quad \leq \lVert Au(t_0)\rVert_{L^2}^2\,{\rm e}^{-\nu(t-t_0)}+{\tilde{c}}_2^2\nu^2(\tilde{\sigma}_1+\tilde{G})^2\tilde{G}^2, \end{align*}

where ${\tilde {c}}_2=12c_L$, which holds for all $t\geq t_0\geq 0$.

Therefore, if $\lVert Au _0\rVert _{L^2}\leq \alpha {\tilde {c}}_2\nu (\tilde {\sigma }_1+\tilde {G})\tilde {G}$, for any $\alpha >0$, then

\[ \lVert Au(t)\rVert_{L^2}^2\leq {\tilde{c}}_2^2(1+\alpha^2)\nu^2(\tilde{\sigma}_1+\tilde{G})^2\tilde{G}^2, \]

for all $t\geq 0$. This completes the proof.

References

Albanez, D. A. F. and Benvenutti, M. J.. Continuous data assimilation algorithm for simplified bardina model. Evol. Equ. Control Theory 7 (2018), 3352.CrossRefGoogle Scholar
Albanez, D. A. F., Nussenzveig Lopes, H. J. and Titi, E. S.. Continuous data assimilation for the three-dimensional Navier–Stokes-$\alpha$ model. Asymptotic Anal. 97 (2016), 139164.CrossRefGoogle Scholar
Altaf, M. U., Titi, E. S., Gebrael, T., Knio, O. M., Zhao, L. and McCabe, M. F.. Downscaling the 2D Bénard convection equations using continuous data assimilation. Comput. Geosci. 21 (2017), 393410.CrossRefGoogle Scholar
Azouani, A., Olson, E. J. and Titi, E. S.. Continuous data assimilation using general interpolant observables. J. Nonlinear Sci. 24 (2014), 277304.CrossRefGoogle Scholar
Balci, N., Foias, C. and Jolly, M. S.. 2-D turbulence for forcing in all scales. J. Math. Pures Appl. 94 (2010), 132.CrossRefGoogle Scholar
Berselli, L. C., Iliescu, T. and Layton, W. J.. Mathematics of large eddy simulation of turbulent flows. Scientific Computation (Berlin: Springer-Verlag, 2006).Google Scholar
Bessaih, H., Olson, E. J. and Titi, E. S.. Continuous data assimilation with stochastically noisy data. J. Nonlinear Sci. 28 (2015), 729753.CrossRefGoogle Scholar
Biswas, A., Brown, K. R. and Martinez, V. R.. Higher-order synchronization and a refined paradigm for global interpolant observables. Ann. Appl. Math. 38 (2022), 160.Google Scholar
Biswas, A. and Hudson, J.. Determining the viscosity of the Navier–Stokes equations from observations of finitely many modes. Inverse Probl. 39 (2023), 125012.CrossRefGoogle Scholar
Biswas, A. and Martinez, V. R.. Higher-order synchronization for a data assimilation algorithm for the 2D Navier–Stokes equations. Nonlinear Anal. Real World Appl. 35 (2017), 132157.CrossRefGoogle Scholar
Biswas, A. and Price, R.. Continuous data assimilation for the three-dimensional Navier–Stokes equations. SIAM J. Math. Anal. 53 (2021), 66976723.CrossRefGoogle Scholar
Blömker, D., Law, K., Stuart, A. M. and Zygalakis, K. C.. Accuracy and stability of the continuous-time 3DVAR filter for the Navier–Stokes equation. Nonlinearity 26 (2013), 21932219.CrossRefGoogle Scholar
Blocher, J., Martinez, V. R. and Olson, E.. Data assimilation using noisy time-averaged measurements. Physica D 376-377 (2018), 4959. Special Issue: Nonlinear Partial Differential Equations in Mathematical Fluid Dynamics.CrossRefGoogle Scholar
Cao, Y., Jolly, M. S., Titi, E. S. and Whitehead, J. P.. Algebraic bounds on the Rayleigh–Bénard attractor. Nonlinearity 34 (2021), 509531.CrossRefGoogle Scholar
Carlson, E., Hudson, J. and Larios, A.. Parameter recovery for the 2 dimensional Navier–Stokes equations via continuous data assimilation. SIAM J. Sci. Comput. 42 (2020), A250A270.CrossRefGoogle Scholar
Carlson, E., Hudson, J., Larios, A., Martinez, V. R., Ng, E. and Whitehead, J. P.. Dynamically learning the parameters of a chaotic system using partial observations. Discrete Contin. Dyn. Syst. 42 (2022), 38093839.CrossRefGoogle Scholar
Carlson, E. and Larios, A.. Sensitivity analysis for the 2D Navier–Stokes equations with applications to continuous data assimilation. J. Nonlinear Sci. 31 (2021), 84.CrossRefGoogle Scholar
Celik, E., Olson, E. and Titi, E. S.. Spectral filtering of interpolant observables for a discrete-in-time downscaling data assimilation algorithm. SIAM J. Appl. Dyn. Syst. 18 (2019), 11181142.CrossRefGoogle Scholar
Chen, N., Li, Y. and Lunasin, E.. An efficient continuous data assimilation algorithm for the Sabra shell model of turbulence. Chaos 31 (2021), 103123.CrossRefGoogle ScholarPubMed
Cheskidov, A. and Dai, M.. Kolmogorov's dissipation number and the number of degrees of freedom for the 3D Navier–Stokes equations. Proc. R. Soc. Edinburgh 149 (2019), 429446.CrossRefGoogle Scholar
Cialenco, I. and Glatt-Holtz, N.. Parameter estimation for the stochastically perturbed Navier–Stokes equations. Stochastic Process. Appl. 121 (2011), 701724.CrossRefGoogle Scholar
Clark Di Leoni, P., Mazzino, A. and Biferale, L.. Inferring flow parameters and turbulent configuration with physics-informed data assimilation and spectral nudging. Phys. Rev. Fluids 3 (2018), 104604.CrossRefGoogle Scholar
Clark Di Leoni, P., Mazzino, A. and Biferale, L.. Synchronization to big data: nudging the Navier–Stokes equations for data assimilation of turbulent flows. Phys. Rev. X 10 (2020), 011023.Google Scholar
Cockburn, B., Jones, D. A. and Titi, E. S.. Determining degrees of freedom for nonlinear dissipative equations. C. R. Acad. Sci. Paris Sér. I Math. 321 (1995), 563568.Google Scholar
Cockburn, B., Jones, D. A. and Titi, E. S.. Estimating the number of asymptotic degrees of freedom for nonlinear dissipative systems. Math. Comput. 66 (1997), 10731087.CrossRefGoogle Scholar
Constantin, P. and Foias, C.. Navier–Stokes equations. Chicago Lectures in Mathematics (Chicago, IL: University of Chicago Press, 1988).CrossRefGoogle Scholar
Dascaliuc, R., Foias, C. and Jolly, M. S.. Relations between energy and enstrophy on the global attractor of the 2-D Navier–Stokes equations. J. Dyn. Differ. Equ. 17 (2005), 643736.CrossRefGoogle Scholar
Dascaliuc, R., Foias, C. and Jolly, M. S.. Some specific mathematical constraints on 2D turbulence. Physica D 237 (2008), 30203029.CrossRefGoogle Scholar
Dascaliuc, R., Foias, C. and Jolly, M. S.. On the asymptotic behavior of average energy and enstrophy in 3D turbulent flows. Physica D 238 (2009), 725736.CrossRefGoogle Scholar
Desamsetti, S., Dasari, H. P., Langodan, S., Titi, E. S., Knio, O. and Hoteit, I.. Efficient dynamical downscaling of general circulation models using continuous data assimilation. Q. J. R. Meteorol. Soc. 145 (2019), 31753194.CrossRefGoogle Scholar
Diegel, A. E. and Rebholz, L. G.. Continuous data assimilation and long-time accuracy in a $C^0$ interior penalty method for the Cahn–Hilliard equation. Appl. Math. Comput. 424 (2022), 127042.Google Scholar
Farhat, A., Glatt-Holtz, N. E., Martinez, V. R., McQuarrie, S. A. and Whitehead, J. P.. Data assimilation in large Prandtl Rayleigh–Bénard convection from thermal measurements. SIAM J. Appl. Dyn. Syst. 19 (2020), 510540.CrossRefGoogle Scholar
Farhat, A., Johnston, H., Jolly, M. S. and Titi, E. S.. Assimilation of nearly turbulent Rayleigh–Bénard flow through vorticity or local circulation measurements: a computational study. J. Sci. Comput. 77 (2018), 115.CrossRefGoogle Scholar
Farhat, A., Jolly, M. S. and Titi, E. S.. Continuous data assimilation for the 2D Bénard convection through velocity measurements alone. Physica D 303 (2015), 5966.CrossRefGoogle Scholar
Farhat, A., Lunasin, E. and Titi, E. S.. Data assimilation algorithm for 3D Bénard convection in porous media employing only temperature measurements. J. Math. Anal. Appl. 438 (2016), 492506.CrossRefGoogle Scholar
Farhat, A., Lunasin, E. and Titi, E. S.. On the Charney conjecture of data assimilation employing temperature measurements alone: the paradigm of 3D planetary geostrophic model. Math. Clim. Weather Forecast. 2 (2016), 6174.Google Scholar
Farhat, A., Lunasin, E. and Titi, E. S.. Continuous data assimilation for a 2D Bénard convection system through horizontal velocity measurements alone. J. Nonlinear Sci. 27 (2017), 10651087.CrossRefGoogle Scholar
Farhat, A., Lunasin, E. and Titi, E. S.. A data assimilation algorithm: the paradigm of the 3D Leray- $\alpha$ model of turbulence. In Nonlinear Partial Differential Equations Arising from Geometry and Physics. London Mathematical Society. Lecture Notes Series (Singapore: Cambridge University Press, 2017).Google Scholar
Foias, C., Jolly, M. S., Kravchenko, R. and Titi, E. S.. A determining form for the two-dimensional Navier–Stokes equations: the Fourier modes case. J. Math. Phys. 53 (2012), 115623.CrossRefGoogle Scholar
Foias, C., Jolly, M. S., Lithio, D. and Titi, E. S.. One-dimensional parametric determining form for the two-dimensional Navier–Stokes equations. J. Nonlinear Sci. 27 (2017), 15131529.CrossRefGoogle Scholar
Foias, C., Manley, O., Rosa, R. and Temam, R.. Navier–Stokes equations and turbulence. Encyclopedia of Mathematics and its Applications, vol. 83 (Cambridge: Cambridge University Press, 2001).CrossRefGoogle Scholar
Foias, C., Mondaini, C. and Titi, E. S.. A discrete data assimilation scheme for the solutions of the 2D Navier–Stokes equations and their statistics. SIAM J. Appl. Dyn. Syst. 15 (2016), 20192142.CrossRefGoogle Scholar
Foias, C., Nicolaenko, B., Sell, G. R. and Temam, R.. Inertial manifolds for the Kuramoto–Sivashinsky equation and an estimate of their lowest dimension. J. Math. Pures Appl. 67 (1988), 197226.Google Scholar
Foiaş, C. and Prodi, G.. Sur le comportement global des solutions non-stationnaires des équations de Navier–Stokes en dimension 2. Rend. Sem. Mat. Univ. Padova 39 (1967), 134.Google Scholar
Foias, C., Sell, G. R. and Temam, T.. Inertial manifolds for nonlinear evolutionary equations. J. Differ. Equ. 73 (1988), 309353.CrossRefGoogle Scholar
Foias, C. and Temam, R.. Determination of the solutions of the Navier–Stokes equations by a set of nodal values. Math. Comput. 43 (1984), 117133.CrossRefGoogle Scholar
Foias, C. and Temam, R.. The connection between the Navier–Stokes equations, dynamical systems, and turbulence theory. In Directions in Partial Differential Equations (Madison, WI, 1985). Publ. Math. Res. Center Univ. Wisconsin, vol. 54, pp. 55–73 (Boston, MA: Academic Press, 1987).CrossRefGoogle Scholar
García-Archilla, B. and Novo, J.. Error analysis of fully discrete mixed finite element data assimilation schemes for the Navier–Stokes equations. Adv. Comput. Math. 46 (2020), 61.CrossRefGoogle Scholar
García-Archilla, B., Novo, J. and Titi, E. S.. Uniform in time error estimates for a finite element method applied to a downscaling data assimilation algorithm for the Navier–Stokes equations. SIAM J. Numer. Anal. 58 (2020), 410429.CrossRefGoogle Scholar
Gesho, M., Olson, E. J. and Titi, E. S.. A computational study of a data assimilation algorithm for the two-dimensional Navier–Stokes equations. Commun. Comput. Phys. 19 (2016), 10941110.CrossRefGoogle Scholar
Ibdah, H. A., Mondaini, C. F. and Titi, E. S.. Fully discrete numerical schemes of a data assimilation algorithm: uniform-in-time error estimates. IMA J. Numer. Anal. 40 (2019), 25842625.CrossRefGoogle Scholar
Jolly, M. S., Martinez, V. R., Olson, E. J. and Titi, E. S.. Continuous data assimilation with blurred-in-time measurements of the surface quasi-geostrophic equation. Chin. Ann. Math. Ser. B 40 (2019), 721764.CrossRefGoogle Scholar
Jolly, M. S., Martinez, V. R. and Titi, E. S.. A data assimilation algorithm for the 2D subcritical surface quasi-geostrophic equation. Adv. Nonlinear Stud. 35 (2017), 167192.CrossRefGoogle Scholar
Jolly, M. S., Sadigov, T. and Titi, E. S.. A determining form for the damped driven nonlinear Schrödinger equation–Fourier modes case. J. Differ. Equ. 258 (2015), 27112744.CrossRefGoogle Scholar
Jolly, M. S., Sadigov, T. and Titi, E. S.. Determining form and data assimilation algorithm for weakly damped and driven Korteweg–De Vries equation – Fourier modes case. Nonlinear Anal. Real World Appl. 36 (2017), 287317.CrossRefGoogle Scholar
Jones, D. A. and Titi, E. S.. Determining finite volume elements for the 2D Navier–Stokes equations. Physica D 60 (1992), 165174.CrossRefGoogle Scholar
Jones, D. A. and Titi, E. S.. On the number of determining nodes for the 2D Navier–Stokes equations. J. Math. Anal. 168 (1992), 7288.CrossRefGoogle Scholar
Larios, A. and Pei, Y.. Approximate continuous data assimilation of the 2D Navier–Stokes equations via the Voigt-regularization with observable data. Evol. Equ. Control Theory 9 (2020), 733751.CrossRefGoogle Scholar
Larios, A., Rebholz, L. G. and Zerfas, C.. Global in time stability and accuracy of IMEX-FEM data assimilation schemes for Navier–Stokes equations. Comput. Methods Appl. Mech. Eng. 345 (2019), 10771093.CrossRefGoogle Scholar
Marchioro, C.. An example of absence of turbulence for any Reynolds number. I. Commun. Math. Phys. 105 (1986), 99106.CrossRefGoogle Scholar
Martinez, V. R.. Convergence analysis of a viscosity parameter recovery algorithm for the 2D Navier–Stokes equations. Nonlinearity 35 (2022), 22412287.CrossRefGoogle Scholar
Mondaini, C. F. and Titi, E. S.. Uniform-in-time error estimates for the postprocessing Galerkin method applied to a data assimilation algorithm. SIAM J. Numer. Anal. 56 (2018), 78110.CrossRefGoogle Scholar
Pachev, B., Whitehead, J. P. and McQuarrie, S. A.. Concurrent multi-parameter learning demonstrated on the Kuramoto–Sivashinsky equation (2021).CrossRefGoogle Scholar
Pei, Y.. Continuous data assimilation for the 3D primitive equations of the ocean. Commun. Pure Appl. Anal. 18 (2019), 643661.CrossRefGoogle Scholar
Temam, R.. Navier–Stokes Equations: Theory and Numerical Analysis, Reprint of the 1984 edn (Providence, RI: AMS Chelsea Publishing, 2001).Google Scholar
Temam, R.. Infinite-dimensional dynamical systems in mechanics and physics, 2nd edn. Applied Mathematical Sciences, vol. 68 (New York: Springer-Verlag, 1997).CrossRefGoogle Scholar
Yu, C., Giorgini, A., Jolly, M. S. and Pakzad, A.. Continuous data assimilation for the 3D Ladyzhenskaya model: analysis and computations. Nonlinear Anal.: Real World Appl. 68 (2022), 103659.Google Scholar
Zauner, M., Mons, V., Marquet, O. and Leclaire, B.. Nudging-based data assimilation of the turbulent flow around a square cylinder. J. Fluid Mech. 937 (2022), A38.CrossRefGoogle Scholar