Filtering Nonlinear Turbulent Dynamical Systems through Conditional Gaussian Statistics

Nan Chen Department of Mathematics and Center for Atmosphere Ocean Science, Courant Institute of Mathematical Sciences, New York University, New York, New York

Search for other papers by Nan Chen in
Current site
Google Scholar
PubMed
Close
and
Andrew J. Majda Department of Mathematics and Center for Atmosphere Ocean Science, Courant Institute of Mathematical Sciences, New York University, New York, and Center for Prototype Climate Modeling, NYU Abu Dhabi, Saadiyat Island, Abu Dhabi, United Arab Emirates

Search for other papers by Andrew J. Majda in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

In this paper, a general conditional Gaussian framework for filtering complex turbulent systems is introduced. Despite the conditional Gaussianity, such systems are nevertheless highly nonlinear and are able to capture the non-Gaussian features of nature. The special structure of the filter allows closed analytical formulas for updating the posterior states and is thus computationally efficient. An information-theoretic framework is developed to assess the model error in the filter estimates. Three types of applications in filtering conditional Gaussian turbulent systems with model error are illustrated. First, dyad models are utilized to illustrate that ignoring the energy-conserving nonlinear interactions in designing filters leads to significant model errors in filtering turbulent signals from nature. Then a triad (noisy Lorenz 63) model is adopted to understand the model error due to noise inflation and underdispersion. It is also utilized as a test model to demonstrate the efficiency of a novel algorithm, which exploits the conditional Gaussian structure, to recover the time-dependent probability density functions associated with the unobserved variables. Furthermore, regarding model parameters as augmented state variables, the filtering framework is applied to the study of parameter estimation with detailed mathematical analysis. A new approach with judicious model error in the equations associated with the augmented state variables is proposed, which greatly enhances the efficiency in estimating model parameters. Other examples of this framework include recovering random compressible flows from noisy Lagrangian tracers, filtering the stochastic skeleton model of the Madden–Julian oscillation (MJO), and initialization of the unobserved variables in predicting the MJO/monsoon indices.

Denotes Open Access content.

Corresponding author address: Nan Chen, Department of Mathematics and Center for Atmosphere Ocean Science, Courant Institute of Mathematical Sciences, New York University, 251 Mercer St., New York, NY 10012. E-mail: chennan@cims.nyu.edu

Abstract

In this paper, a general conditional Gaussian framework for filtering complex turbulent systems is introduced. Despite the conditional Gaussianity, such systems are nevertheless highly nonlinear and are able to capture the non-Gaussian features of nature. The special structure of the filter allows closed analytical formulas for updating the posterior states and is thus computationally efficient. An information-theoretic framework is developed to assess the model error in the filter estimates. Three types of applications in filtering conditional Gaussian turbulent systems with model error are illustrated. First, dyad models are utilized to illustrate that ignoring the energy-conserving nonlinear interactions in designing filters leads to significant model errors in filtering turbulent signals from nature. Then a triad (noisy Lorenz 63) model is adopted to understand the model error due to noise inflation and underdispersion. It is also utilized as a test model to demonstrate the efficiency of a novel algorithm, which exploits the conditional Gaussian structure, to recover the time-dependent probability density functions associated with the unobserved variables. Furthermore, regarding model parameters as augmented state variables, the filtering framework is applied to the study of parameter estimation with detailed mathematical analysis. A new approach with judicious model error in the equations associated with the augmented state variables is proposed, which greatly enhances the efficiency in estimating model parameters. Other examples of this framework include recovering random compressible flows from noisy Lagrangian tracers, filtering the stochastic skeleton model of the Madden–Julian oscillation (MJO), and initialization of the unobserved variables in predicting the MJO/monsoon indices.

Denotes Open Access content.

Corresponding author address: Nan Chen, Department of Mathematics and Center for Atmosphere Ocean Science, Courant Institute of Mathematical Sciences, New York University, 251 Mercer St., New York, NY 10012. E-mail: chennan@cims.nyu.edu

1. Introduction

Turbulent dynamical systems are ubiquitous in many disciplines of contemporary science and engineering (Hinze and Hinze 1959; Townsend 1980; Frisch 1995; Majda and Wang 2006; Vallis 2006; Salmon 1998). They are characterized by both a large dimensional phase space and a large dimensional space of instability with positive Lyapunov exponents. These linear instabilities are mitigated by energy-conserving nonlinear interactions, yielding physical constraints (Majda and Harlim 2013; Sapsis and Majda 2013b; Majda and Harlim 2012; Harlim et al. 2014), which transfer energy to the linear stable modes where it is dissipated resulting in a statistical steady state. Both understanding complex turbulent systems and improving initializations for prediction require filtering for an accurate estimation of full state variables from noisy partial observations. Since the filtering skill for turbulent signals from nature is often limited by errors due to utilizing an imperfect forecast model, coping with model errors is of wide contemporary interest (Majda and Harlim 2012; Majda 2012).

Many dynamical models of turbulent systems are summarized as conditional Gaussian systems (Majda and Harlim 2012; Majda 2003; Kalnay 2003; Majda and Gershgorin 2013; Majda et al. 1999). Despite the conditional Gaussianity, such systems nevertheless can be highly nonlinear and able to capture the non-Gaussian features of nature (Berner and Branstator 2007; Neelin et al. 2010). In this paper, we introduce a general conditional Gaussian framework for continuous-time filtering. The conditional Gaussianity means that once the trajectories of the observational variables are given, the dynamics of the unobserved variables conditioned on these highly nonlinear observed trajectories become Gaussian processes. One of the desirable features of such conditional Gaussian filter is that it allows closed analytical formulas for updating the posterior states associated with the unobserved variables (Liptser and Shiryaev 2001) and is thus computationally efficient.

The recently developed nonlinear filter for filtering the stochastic skeleton model for the Madden–Julian oscillation (MJO; Chen and Majda 2016a) belongs to the conditional Gaussian filtering framework, where equatorial waves and moisture were filtered given the observations of the highly intermittent envelope of convective activity. Another application of this exact nonlinear filter involves filtering turbulent flow fields utilizing observations from noisy Lagrangian tracer trajectories (Chen et al. 2014c, 2015), where an information barrier was shown as increasing the number of tracers (Chen et al. 2014c) and a multiscale filtering strategy was studied for the system with coupled slow vortical modes and fast gravity waves (Chen et al. 2015). In addition, a family of low-order physics-constrained nonlinear stochastic models with intermittent instability and unobserved variables, which belong to the conditional Gaussian family, was proposed for predicting the MJO and the monsoon indices (Chen et al. 2014b; Chen and Majda 2015b,a). The effective filtering scheme was adopted for the online initialization of the unobserved variables that facilitates ensemble prediction algorithm. Other applications that fit into the conditional Gaussian framework includes the cheap exactly solvable forecast models in dynamic stochastic superresolution of sparsely observed turbulent systems (Branicki and Majda 2013; Keating et al. 2012), stochastic superparameterization for geophysical turbulence (Majda and Grooms 2014), and blended particle filters for large-dimensional chaotic systems (Majda et al. 2014; Qi and Majda 2015) that captures non-Gaussian features in an adaptively evolving low-dimensional subspace through particles interacting with conditional Gaussian statistics on the remaining phase space.

In this paper, we illustrate three types of applications of the conditional Gaussian filtering framework, where the effect of model error is extensively studied. In addition to the traditional pathwise measures, an information-theoretic framework (Branicki and Majda 2014; Branicki et al. 2013; Majda and Branicki 2012; Majda and Wang 2006) is adopted to assess the lack of information and model error in filtering these turbulent systems.

The first application involves utilizing dyad models (Majda and Lee 2014; Majda 2015) to study the effect of model error due to the ignorance of energy-conserving nonlinear interactions in forecast models in filtering turbulent signals from nature. Such model error exists in many ad hoc quadratic multilevel regression models (Kravtsov et al. 2005; Kondrashov et al. 2005; Wikle and Hooten 2010; Cressie and Wikle 2011) that are utilized as data-driven statistical models for time series of partial observations of nature. However, these models were shown to suffer from a finite-time blow up of statistical solutions (Majda and Yuan 2012; Majda and Harlim 2013). To understand the effect of such model error in filtering, a physics-constrained dyad model (Majda and Lee 2014) is adopted to generate the turbulent signals of nature while a stochastic parameterized model without energy-conserving nonlinearities (Majda and Harlim 2012) is adopted as the imperfect filter. The skill of this stochastic parameterized filter is studied in different dynamical regimes and the lack of information in the filter estimates is compared with that using the perfect filter. Meanwhile, the role of observability (Gajic and Lelic 1996; Majda and Harlim 2012) is explored and its necessity in filtering turbulent systems is emphasized.

The second application of the conditional Gaussian framework is to filter a family of triad models (Majda 2003; Majda et al. 1999, 2001, 2002), which include the noisy Lorenz 63 (L-63) model (Lorenz 1963). The goal here is to explore the effect of model error due to noise inflation and underdispersion in designing filters. The motivation of studying such a kind of model error comes from the fact that many models for turbulence are underdispersed since they have too much dissipation (Palmer 2001) due to inadequate resolution and deterministic parameterization of unresolved features. On the other hand, suitably inflating the noise in imperfect forecast models are widely adopted to reduce the lack of information (Anderson 2001; Kalnay 2003; Majda and Harlim 2012) and also to suppress catastrophic filter divergence (Harlim and Majda 2010a; Tong et al. 2016). Besides filtering a single trajectory, recovering the full probability density function (PDF) associated with the unobserved variables given an ensemble of observational trajectories is also of particular interest. Combining the ensembles of the analytically solvable conditional Gaussian distribution associated with filtering each unobserved single trajectory, an effective conditional Gaussian ensemble mixture approach is proposed to approximate the time-dependent PDF associated with the unobserved variables. In this fashion, an efficient algorithm can be generated for systems with a large number of the unobserved variables, compared with applying a direct Monte Carlo method, which is extremely slow and expensive due to the “curse of dimensionality” (Majda and Harlim 2012; Daum and Huang 2003).

Parameter estimation in turbulent systems is an important issue and this is the third topic within the conditional Gaussian filtering framework of this paper. Regarding the model parameters as augmented state variables, algorithms based on particle or ensemble Kalman filters were designed for parameter estimation (Dee 1995; Smedstad and O’Brien 1991; Van Der Merwe and Wan 2001; Plett 2004; Wenzel et al. 2006; Campillo and Rossi 2009; Harlim et al. 2014; Salamon and Feyen 2009). Although many successful results utilizing these algorithms were obtained, very little mathematical analysis was provided for exploring the convergence rate and understanding the potential limitation of such algorithms. Guidelines for enhancing the efficiency of the algorithms are desirable since a short training period is preferred in many real-world applications. In the conditional Gaussian framework, the closed analytic form of the posterior state estimations facilitates the analysis of both the error and the uncertainty in the estimated parameters for a wide family of models, where detailed mathematical justifications are accessible. Here, focus is on the parameter estimation skill dependence on different factors of the model as well as the observability. In some applications, certain prior information of the parameters is available (Yeh 1986; Iglesias et al. 2014). Yet, none of the existing filtering-based parameter estimation approaches emphasizes exploiting such prior information in improving the algorithms. In this paper, stochastic parameterized equations (Majda and Harlim 2012), involving the prior knowledge of the parameters, are incorporated into the filtering algorithm as the underlying processes of the augmented state variables. This improved algorithm greatly enhances the convergence rate at the cost of only introducing a small model error and it is particularly useful when the system loses practical observability.

The remainder of this paper is as follows. The general framework of the conditional Gaussian nonlinear systems is introduced in section 2. In section 3, an information-theoretic framework for assessing the model error in filtering is proposed. The information measures compensate the insufficiency of the pathwise ones in measuring the lack of information in the filtered solutions. Section 4 deals with dyad models, where focus is on model error in filtering due to the lack of respecting the underlying physical dynamics of the partially observed system. In section 5, a general family of triad model is proposed and the noisy L-63 model is adopted as a test model for understanding the model error in noise inflation and underdispersion. In the same section, the conditional Gaussian ensemble mixture for approximating the PDF associated with unobserved variables is introduced and the model error in filtering the PDF utilizing imperfect models is studied. Section 6 involves parameter estimation, where the skill of estimating both additive and multiplicative parameters is illustrated with detailed mathematical analysis. The comparison of utilizing direct method and stochastic parameterized equations approach is shown for estimating parameters in both linear and nonlinear systems. Summary conclusions are included in section 7.

2. Conditional Gaussian nonlinear systems

The conditional Gaussian systems have the following abstract form:
e1a
e1b
where and are vector state variables; , and are vectors and matrices that depend only on time t and state variables ; and and are independent Wiener processes. Once for is given, conditioned on becomes a Gaussian process with mean and covariance :
e2
Despite the conditional Gaussianity, the coupled system in (1) remains highly nonlinear and is able to capture the non-Gaussian features such as skewed or fat-tailed distributions as observed in nature (Berner and Branstator 2007; Neelin et al. 2010).
One of the desirable features of the conditional Gaussian system in (1) is that the conditional distribution in (2) has the following closed analytic form (Liptser and Shiryaev 2001):
e3
The closed form of the exact solutions in (3) provides a general framework for studying continuous-time filtering and uncertainty quantification of the conditional Gaussian system in (1). In filtering the turbulent system (1), if is the observed process, then the posterior states of the unobserved process in (2) are updated following the analytic formulas (3) associated with the nonlinear filter in (1).
It is worthwhile remarking that the classical Kalman–Bucy filter (Kalman and Bucy 1961) is a special example within the general conditional Gaussian filtering framework in (1)(3), where the observed processes in the Kalman–Bucy filter are given by
eq1
and the unobserved variables are driven by the linear time-dependent stochastic differential equation (SDE):
eq2
Corresponding to (3), chapter 6 of Bensoussan (1992) includes rigorous mathematical derivations of the exact solutions of the Kalman–Bucy filter and some other more general conditional Gaussian filters. It is also pointed out in Bensoussan (1992) that all these filters belong to the general conditional Gaussian filtering framework in (1)(3) introduced in Liptser and Shiryaev (1977), which is an early version of Liptser and Shiryaev (2001).

3. An information-theoretic framework for assessing the model error

Assume is the true signal and is the filtered solution. The traditional measures for assessing the filtering skill in the ith dimension of and are the root-mean-square (RMS) error and anomaly pattern correlation (Hyndman and Koehler 2006; Kalnay 2003; Majda and Harlim 2012):
e4
where and represent the ith dimension of the vector fields and , respectively.

Despite their wide applications in assessing filtering and prediction skill, these pathwise measures fail to assess the lack of information in the filter estimates and the predicted states (Branicki and Majda 2014; Chen and Majda 2015b). As shown in Chen and Majda (2015b), two predicted trajectories with completely different amplitudes can have the same RMS error and anomaly pattern correlation. Undoubtedly, the solution having comparable amplitude as the truth is more skillful than the one with strongly underestimated amplitude, which misses all the extreme events (Majda et al. 2010b; Majda and Harlim 2012; Majda and Branicki 2012) that are important for the turbulent systems. Different from the indistinguishable skill utilizing the pathwise measurements, an information-theoretic framework including the measurement of the lack of information succeeds in discriminating the prediction skill of the two solutions.

In Branicki and Majda (2014), a systematic information-theoretic approach was developed to quantify the statistical accuracy of Kalman filters with model error and the optimality of the imperfect Kalman filters in terms of three information measures was presented. Another application of information theory is illustrated in Branicki and Majda (2015) for improving imperfect predictions via multimodel ensemble forecasts.

Following the general information-theoretic framework in Branicki and Majda (2014) and Chen and Majda (2015b), we consider three information measures in assessing the filtering skill.

  1. The Shannon entropy of the residual is given by the following (Majda and Wang 2006; Abramov and Majda 2004):
    e5
    where is the PDF of .
  2. The relative entropy of the PDF associated with compared with the truth π is given by the following (Majda et al. 2005; Majda and Wang 2006; Majda and Branicki 2012; Cai et al. 2002):
    e6
  3. The mutual information between the true signal and the filtered one is given by the following symmetric formula (MacKay 2003; Branicki and Majda 2014):
    e7
    where and are the PDFs of and , respectively, and is the joint distribution of and .

Each one of the three measures provides different information about the filtering skill. The mutual information measures the dependence between and . The Shannon entropy of the residual measures the uncertainty in the filtered solution compared with the truth . These two information measures are the surrogates for the anomaly pattern correlation and RMS error in the pathwise sense, respectively (Branicki and Majda 2014). Particularly, if both the truth and the filtered solution are Gaussian distributed, then the asymptotic anomaly pattern correlation and RMS error can be expressed in analytic forms by the mutual information and the Shannon’s entropy. The relative entropy quantifies the lack of information in the statistics of the filtered solution relative to that of the truth (Majda and Gershgorin 2010; Majda and Branicki 2012). Therefore, it is an indicator of assessing the disparity in the amplitudes and spread between and . Importantly, the relative entropy is able to quantify the ability of capturing the extreme events (Chen et al. 2014b; Chen and Majda 2015b; Branicki and Majda 2014), corresponding to the tails of a distribution, in the filtered solutions. The relative entropy is often interpreted as a “distance” between the two probability densities but it is not a true metric. It is nonnegative with only when and it is invariant under nonlinear changes of variables.

Because of the importance of measuring the lack of information and quantifying the ability of capturing the extreme events in the filtered solutions, the relative entropy is included in assessing the filtering skill throughout this work. Along with the relative entropy, we nevertheless show the anomaly pattern correlation and the RMS error instead of the mutual information and the Shannon’s entropy since the readers are more familiar with these traditional pathwise measures. Yet, it is important to bear in mind that the mutual information and Shannon’s entropy are the surrogates of the pathwise measures in the information-theoretic framework. Note that there are many other scoring functions that are widely utilized in meteorology in assessing the filtering and ensemble prediction skill (Gneiting and Raftery 2007) and are related to the information measures. Yet, the relative entropy can be explained as a measure of information gain, which cannot be interpreted by the traditional scoring functions, and it is also widely adopted in assessing the model error in atmosphere and ocean science (Majda and Harlim 2012).

In the study of filtering the unobserved single trajectories in sections 4 and 5a, the pathwise filtering skill is assessed. Therefore, is the single realization from perfect system and the posterior mean estimation is chosen as the filter estimate . Both the skill scores utilizing (4) and the lack of information in the time-averaged PDF of related to that of the truth via the relative entropy in (6) are assessed. In measuring the lack of information in the recovered time-dependent PDF in sections 5b and 5c, the relative entropy in (6) in the recovered PDF related to the truth π at each time instant is computed, where is obtained from the conditional Gaussian ensemble mixture approach.

It is worthwhile remarking that although most of the focus of this paper is on assessing the lack of information in the pathwise filtering solutions, the information-theoretic framework also provides a general framework for quantifying the uncertainty using imperfect models in ensemble prediction. A detailed discussion is included in appendix E.

4. Dyad models

Many turbulent dynamical systems involve dyad and triad interactions (Majda 2015; Majda and Lee 2014; Majda et al. 2009). These nontrivial nonlinear interactions between large-scale mean flow and turbulent fluctuations generate intermittent instability while the total energy from the nonlinear interactions is conserved. In this and the next sections, we study the filtering skill of dyad and triad models, where the effect of different model errors is explored.

In this section, we utilize dyad models to understand the effect of model error due to the ignorance of energy-conserving nonlinear interactions in forecast models in filtering turbulent signals from nature. As discussed in section 1, such model error exists in many ad hoc quadratic multilevel regression models (Kravtsov et al. 2005; Kondrashov et al. 2005; Wikle and Hooten 2010; Cressie and Wikle 2011) for fitting and predicting time series of partial observations of nature, which were shown to suffer from finite-time blow up of statistical solutions and also have pathological behavior of the related invariant measure (Majda and Yuan 2012; Majda and Harlim 2013). Recently, a new class of physics-constrained nonlinear regression models were developed (Majda and Harlim 2013) and the application of these physics-constrained models in ensemble Kalman filtering is shown in Harlim et al. (2014) together with other recent applications to prediction (Chen et al. 2014b; Chen and Majda 2015a,b).

The general form of the dyad models is described in Majda (2015) and Majda and Lee (2014). Here we focus on the following dyad model:
e8a
e8b
In (8), u is regarded as representing one of the resolved modes in a turbulent signal, which interacts with the unresolved mode υ through quadratic nonlinearities. The conserved energy in the quadratic nonlinear terms in (8) is seen by
eq3
Below, the physics-constrained dyad model (8) is utilized to generate true signals of nature. The goal here is to filter the unobserved process υ given one single realization of the observed process u. In addition to adopting the perfect filter, which has exactly the same dynamics of nature in (8), an imperfect filter with no energy-conserving nonlinear interactions is studied for comparison. In this imperfect filter, the nonlinear feedback in υ is dropped and the result is a stochastic parameterized filter (Majda and Harlim 2012):
e9a
e9b
In the stochastic parameterized filter (9), the parameters in the resolved variable u are assumed to be the same as nature (8). We further assume the statistics of the unobserved variable υ of nature (8) are available. Thus, the parameters and in the unresolved process υ are calibrated (Harlim and Majda 2008, 2010b; Branicki et al. 2013) by matching the mean, variance and decorrelation time of those in (8). Note that both (8) and (9) belong to the conditional Gaussian framework (1) by denoting and .
One important issue in filtering is observability (Gajic and Lelic 1996; Majda and Harlim 2012). The coupled system in (8) is said to lose its observability if the observed process u provides no information in determining the unobserved variable υ. Intuitively, this corresponds to in (8), in which case υ disappears in the observed process u. The rigorous definition of the observability is included in appendix A. To understand the role of observability in filtering, we consider the following two dynamics regimes:
e10
The fixed point associated with the deterministic part of (8) is given, respectively, by
eq4
It is clear that in dynamical regime B the system in (8) loses practical observability when the solution is around the fixed point. (As shown in Fig. 2, both models are able to generate intermittency with suitable choices of the noise in the observational process and in the filtering process.)

Below, the true signals are generated from the dyad model (8) with different noise levels and . The filtering skill scores utilizing both the physics-constrained perfect filter, which has the same form as (8), and the stochastic parameterized imperfect filter (9) are shown in Fig. 1. The first two rows show the RMS error and pattern correlation in the posterior mean estimation of υ and the third row illustrates the information model error in the time-averaged PDF of the posterior mean estimation related to that of the truth π. Here, if the model error is larger than , which is already significant, then the same color as is utilized for representation in Fig. 1.

Fig. 1.
Fig. 1.

Skill scores for filtering the unresolved process υ using the physics-constrained (perfect) filter in (8) and the stochastic parameterized (imperfect) filter in (9) as a function of and in generating the truth. The skill scores in (a),(b) dynamics regime A and (c),(d) dynamics regime B are shown. (top),(middle) The RMS error and pattern correlation in the filtered solution compared with the truth are shown. (bottom) The model error in (6) in the time-averaged PDF of the posterior mean estimation compared with that of the truth π. The parameters and in the stochastic parameterized filter in (9) is calibrated by matching the statistics with those of nature in (8).

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

The skill scores in the dynamical regime A are shown in Figs. 1a and 1b. The physics-constrained perfect filter (8) has high filtering skill when and . As contrast, the stochastic parameterized filter (9) is skillful only when , in which case the filter estimation of υ is mostly determined from the observation process u with small noise ; therefore, the two filters are expected to have comparable high skill when the system has observability. Figure 2a compares the posterior mean estimations across time with and . Clearly, both filters succeed in filtering υ provided that the practical observability is satisfied (i.e., u not approaching zero). On the other hand, as shown in Fig. 2b with and , the energy-conserving perfect filter (8) filters υ almost perfectly while the stochastic parameterized filter (9) has no skill. In fact, implies that the filter trusts more toward the dynamics of υ and the amplitude of energy feedback is much larger than the stochastic forcing in υ. Thus, the process of υ is largely driven by the nonlinear energy feedback in (8b). However, the stochastic parameterized filter (9) has no such mechanism; therefore, the posterior mean estimation is simply around the maximum likelihood state of υ (i.e., the mean ). Importantly, without the nonlinear energy feedback term , the information model error utilizing the imperfect filter (9) remains huge unless .

Fig. 2.
Fig. 2.

(Dynamical regime A: With practical observability at the attractor). Comparison of the posterior mean estimation of υ across time using the physics-constrained (perfect) filter (8) and the stochastic parameterized (imperfect) filter (9). The time-averaged PDFs of the filtered solutions compared with the truth are also illustrated. (a) The situation with small noise in the observational process and large noise in the filtering process. (b) The situation with large noise in the observational process and small noise in the filtering process. The parameters , and in the stochastic parameterized filter in (9) are calibrated by matching the statistics with those of nature in (8).

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

Next, we study the filtering skill in dynamical regime B, at the fixed point of which the system has no observability. Compared with regime A, significant deterioration of the filtering skill is found when in the truth, where the trajectory of u is around the fixed point implying no practical observability (see Fig. 3a). With the increase of , more positive values of υ are reached, which correspond to the increase of intermittent phases of u with large bursts. Clearly, the observability is regained at these intermittent phases and thus an improved skill in filtering is found (see Fig. 3b).

Fig. 3.
Fig. 3.

(Dynamical regime B: Without practical observability at the attractor). As in Fig. 2. (a) The situation with small noise in the observational process and moderate noise in the filtering process. (b) The situation with small noise in the observational process and large noise in the filtering process.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

To conclude, the energy-conserving nonlinear feedback plays a significant role in filtering the dyad model (8), especially with large noise in the observational process. Despite comparable RMS errors, the imperfect stochastic parameterized filter (9) without energy-conserving nonlinearities leads to a much larger information model error than the energy-conserving perfect filter (8) for . In addition, the observability becomes quite important when the noise in the filtering process is moderate and the noise in the observational process is small. An increase of enhances the intermittency that improves the filtering skill.

5. Triad models

The nonlinear coupling in triad systems is generic of nonlinear coupling between any three modes in larger systems with quadratic nonlinearities. Here, we introduce the general form of the triad models that belongs to the conditional Gaussian framework in (1):
e11
where and and the coefficients , and are functions of only the observed variable. In (11), either or can be regarded as the observed variable and correspondingly the other one becomes the unresolved variable that requires filtering. The triad model (11) has wide applications in atmosphere and ocean science. One example is the stochastic mode reduction model [also known as the Majda–Timofeyev–Vanden-Eijnden (MTV) model] (Majda et al. 2003, 1999, 2002, 2001), which includes both a wave–mean flow triad model and a climate scattering triad model for barotropic equations (Majda et al. 2001). Another example of (11) involves the slow–fast waves in the coupled atmosphere–ocean system (Majda and Harlim 2012), where one slow vortical mode interacts with two fast gravity modes with the same Fourier wavenumber.
With the following choice of the matrices and vectors in (11),
eq5
the triad model (11) becomes the noisy Lorenz 63 (L-63) model (Lorenz 1963):
e12
As is known, adopting the following parameters:
e13
the deterministic version of (12) has chaotic solutions, where the trajectory of the system has a butterfly profile at the attractor. Such a feature is preserved in the appearance of small or moderate noise in (12). See Fig. 4 for the trajectories of (12) with , and 10, respectively. Note that the noisy L-63 model possesses the property of energy-conserving nonlinear interactions.
Fig. 4.
Fig. 4.

Trajectories of the noisy L-63 model (12). (a) , (b) , and (c) .

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

The noisy L-63 model (12) equipped with the parameters in (13) is utilized as a test model in this section. Below, we first study filtering the unresolved trajectories given one realization of the noisy observations. Then an efficient conditional Gaussian ensemble mixture approach is designed to approximate the time-dependent PDF associated with the unresolved variables, which requires only a small ensemble of the observational trajectories. In both studies, the effect of model error due to noise inflation and underdispersion is studied. The underdispersion occurs in many models for turbulence since they have too much dissipation (Palmer 2001) due to inadequate resolution and deterministic parameterization of unresolved features while noise inflation is adopted in many imperfect forecast models to reduce the lack of information (Anderson 2001; Kalnay 2003; Majda and Harlim 2012) and suppress the catastrophic filter divergence (Harlim and Majda 2010a; Tong et al. 2016).

a. Model error in filtering the unresolved processes

We explore filtering the unresolved single trajectories in the L-63 model utilizing imperfect filters, where model error comes from the noise , and in both the observational and filtering processes. Here, the noisy L-63 model (12) is adopted to generate true signals. The model utilized for filtering differs from (12) by the noise amplitudes:
e14
Note that although the noise in the filtering processes in (14) can be arbitrary, the noise amplitude in the observational processes must be nonzero to avoid the singularity in solving the posterior estimations in (3).

1) Filtering the deterministic L-63 system utilizing the imperfect forecast model with noise

The first test involves the situation that the true signal is generated from the L-63 model which has no stochastic noise [i.e., in (12)]:
e15
The filtering skill utilizing the imperfect model in (14) with nonzero noise is studied. This demonstrates the role of noise inflation in the forecast model. In the situation of filtering x with observations from y and z, we have the following results.
Proposition 1: Assume the true signal is generated from the system (15). In the situation of filtering x with observations from y and z, the posterior variance of and the error in the posterior mean utilizing the forecast model (14) with nonzero and , are bounded by
e16a
e16b
where and are the mean and uncertainty of variable x at initial time, respectively.

The detailed derivation of Proposition 1 is shown in appendix B. The results in (16) imply that the error in the posterior mean estimation decays to zero in an exponentially fast rate regardless of the noise level , and in the imperfect filter (14). The uncertainty after the initial period is essentially bounded by the noise variance associated with the filtering process over the known parameter . This indicates if the noise is zero in the imperfect forecast model (14), then the posterior estimation will converge to the truth with an uncertainty that decays exponentially to zero. Figure 5a validates Proposition 1, where the statistics are averaged across time . The nearly zero RMS error and nearly one pattern correlation reveal that the posterior mean converges to the truth. The posterior variance increases as the noise in the observational process increases.

Fig. 5.
Fig. 5.

Filtering the L-63 model (15) utilizing the imperfect noisy L-63 model (14). (a) Observing y and z and filtering x. The three noise levels , and in the imperfect forecast model (14) are set to be the same, and the RMS error and pattern correlation in the posterior mean estimation compared with the truth as a function of these noise levels are shown in the first two rows, where the statistics are computed across time and the dotted line shows the standard deviation of each variable at equilibrium. The third row shows the posterior variance as a function of the noise levels, where the statistics are averaged across time and the dotted line shows the equilibrium covariance of each variable. (b) Observing x and filtering y and z. (c) The observation in x across time and the filtering estimators in y and z with and nonzero noise in the imperfect forecast model (14). (d) As in (c), but with zero noise .

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

The qualitative conclusions are the same in the situation of observing x and filtering y and z (see Fig. 5b). The uncertainty in filtering z is larger than that in filtering y, since y is directly related in the observational process x in (14). The trajectories as a function of time are shown in Figs. 5c and 5d, with nonzero and zero noise in the filtering processes, respectively. In both cases the posterior mean converges to the truth. If the noise is nonzero, then the posterior variance for both y and z remains nonzero but is bounded.

These results indicate that noise inflation in the imperfect forecast model brings no error regarding the posterior mean estimation and a bounded posterior uncertainty after a short relaxation time provided that the signal is generated from the system with no stochastic noise.

2) Filtering the noisy L-63 system utilizing the imperfect forecast model with no noise in the filtering processes

Next, we reverse the setup in section 5a(1). We assume the true signal is generated from the noisy L-63 model in (12) but the imperfect forecast model (14) contains no noise in the filtering process. This illustrates the effect of utilizing an underdispersive imperfect forecast model in filtering. Note that the noise in the observational process in (14) must be nonzero to avoid the singularity in solving the posterior states in (3). Since the two situations that observing either x or y and z lead to qualitatively the same results, we focus on the situation that only x is observed. Thus, the imperfect filter has the following form:
e17
Below, we assume the noise in the observational process in (17) is the same as in the model in (12) that generates the true signal. Then model error in filtering simply comes from the ignorance of the noise and in the filtering processes in (12). Figures 6a–c show the dependence of the statistics on the noise and in (12), where we set for simplicity. Clearly, with the increase of and , the filtering skill regarding the RMS error and the pattern correlation in the posterior mean of both y and z becomes worse while these posterior states are quite certain, both of which indicate the negative effect of underdispersion in the imperfect forecast model. In addition, the model error in (6) increases as a function of the noise and in (12) and is larger in filtering variable z than y. The comparable statistics in the three columns of Fig. 6 reveal that increasing the noise in the observational process in the true model (12) has little effect in filtering the unresolved variables provided that the noise in the observational process in the imperfect forecast model (17) equals .
Fig. 6.
Fig. 6.

Filtering the noisy L-63 model (12) utilizing the imperfect forecast L-63 model (17) with no noise in the filtering processes, where x is the observed variable and y and z are the variables for filtering. The noise in the observational process in (17) is set to be the same in (12). (a)–(c) The RMS error and pattern correlation in the posterior mean compared with the truth, the posterior covariance, and model error in (6) with different as a function of the noise and in the true model in (12) where . The dotted line in the first row shows the equilibrium standard deviation and that in the third row shows the equilibrium variance of each variable. The statistics are averaged across time . (d) The filtering estimation across time, where .

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

In Fig. 6d, we show the trajectories with and . Therefore, a severe underdispersion exists in the imperfect forecast model (17). A larger skewness is found in time-averaged PDF of the filter estimation of z than that of the truth, which explains the model error.

3) Filtering the noisy L-63 system utilizing the imperfect L-63 forecast model with different noise amplitudes

Finally, we study the general situation that both the system that generates the true signal in (12) and the imperfect filter (14) contain nonzero noise. Again, we illustrate the situation with observing x and filtering y and z. The other case has the same qualitative results.

Figure 7 shows the filtering skill utilizing the imperfect filter (14), where the model error comes from either the noise in the observational process or those in the filtering processes and the noise levels in the truth , and are set to equal with each other. In Figs. 7a, 7c, and 7e, the noise level in the true dynamics in (12) is gradually increased to , and 10, respectively. In the imperfect filter, the noise and are taken to be the same as and and the filtering skill with different noise in the observational process is studied. Clearly, inflating the noise in the observational process in the imperfect forecast model (14) leads to only small model errors (Figs. 7a,c with ). On the other hand, underestimating corresponds to a rapid increase of the RMS error and a quick decrease of the pattern correlation (Fig. 7e). At the same time, the posterior variance in the underdispersion case becomes smaller, implying these inaccurate estimations are quite certain. It is worthwhile noting that the model error in the time-averaged PDF of the posterior mean estimation associated with variable z shoots up in the underdispersive case (Fig. 7e), indicating a significant lack of information in the filter estimates. Similar conclusions are found in with imperfect noise levels and . Large errors are found when and are underdispersed (Fig. 7f) while noise inflation has little negative effect on the model error (Figs. 7b,d).

Fig. 7.
Fig. 7.

Filtering the noisy L-63 model (12) utilizing an imperfect forecast model (14) with model error in the noise, where the observational variable is x and the variables for filtering are y and z. (a),(c),(e) Filtering skill as a function of the noise in the observational process. (b),(d),(f) Filtering skill as a function of the noise and , where . (a),(b) Small; (c),(d) moderate; and (e),(f) large noise , and 10 in the true system are shown. The dotted line in the first row shows the equilibrium standard deviation and that in the third row shows the equilibrium variance of each variable. The statistics are averaged across time .

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

Figure 8 illustrates the posterior mean estimation as a function of time compared with the truth in two underdispersive situations. In the case that the noise in the observational process is underestimated (Figs. 8a–c), the filtered trajectories of both y and z are quite noisy. In addition, the mean of the PDF associated with the filtered variable z has a positive bias, which explains the large model error in Fig. 7e. Looking at the filtered trajectory of z in Fig. 8c, the filtered solution misses many negative extreme events such as those around time , and 18. At these time instants, the corresponding values of the observed variable x are all near zero, which implies the system loses practical observability (see appendix A for details). In fact, when the process of z is completely decoupled from x and y in (14) and observing x plays no role in filtering z. On the other hand, as shown in Figs. 8d–f, even though the model errors in the unresolved variables y and z are small with the underestimated noise and , the RMS error in the filter estimation remains significant.

Fig. 8.
Fig. 8.

Comparison of the true signal (blue) and posterior mean estimation (red) in the underdispersion cases, where in the model that generates the true signal in (12). (a)–(c) Filtering skill with underdispersed noise in the observational process. (d)–(f) Filtering skill with underdispersed noise in the filtering processes.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

Therefore, we conclude that underdispersion in the imperfect filter deteriorates the filtering skill while noise inflation within certain range introduces little error.

b. Recovering the time-dependent PDF of the unresolved variables utilizing conditional Gaussian mixture

One important issue in uncertainty quantification for turbulent systems is to recover the time-dependent PDF associated with the unobserved processes. In a typical scenario, the phase space of the unobserved variables is quite large while that of the observed ones remains moderate or small. The classical approaches involve solving the Fokker–Planck equation or adopting the Monte Carlo simulation, both of which are quite expensive with the increase of the dimension, known as the curse of dimensionality (Majda and Harlim 2012; Daum and Huang 2003). For conditional Gaussian systems, the PDF associated with the unobserved processes can be approximated by an efficient conditional Gaussian ensemble mixture with high accuracy, where only a small ensemble of observed trajectories is needed due to its relatively low dimension and is thus computationally affordable. Here, an ensemble of independent observed trajectories is usually obtained from repeated experiments, which is not uncommon in real applications. One typical example involves recovering the PDF of random turbulent flows utilizing Lagrangian tracers (Chen et al. 2014c, 2015), where each observed tracer trajectory evolves independently. Note that the idea here is similar to that of the blended method for filtering high-dimensional turbulent systems (Majda et al. 2014; Qi and Majda 2015; Slivinski et al. 2015; Sapsis and Majda 2013a).

Below, we provide a general framework of utilizing conditional Gaussian mixtures in approximating the time-dependent PDF associated with the unobserved processes. Although the test examples of this approach below are based on the 3D noisy L-63 system, this method can be easily generalized to systems with a large number of unobserved variables. This section deals with the situation with no model error. In section 5c, the skill of recovering the PDF in the appearance of the model error due to noise inflation or underdispersion is explored.

Let us recall the observed variables and the unobserved variables in the conditional Gaussian system (1). Their joint distribution is denoted by
eq6
Assume we have L independent observational trajectories and, therefore, they are equally weighted. The marginal distribution of is approximated by
e18
The marginal distribution of at time t is expressed by
e19
where for each observation , according to the analytically closed form in (3):
e20
Thus, the PDF associated with the unobserved variable is approximated utilizing (19) and (20). Note that in many practical issues associated with turbulent systems, the dimension of the observed variables is much lower than that of the unobserved ones. Thus, only a small number of L is needed in approximating the low-dimensional marginal distribution in (18) to recover the marginal distribution associated with the unobserved process with this conditional Gaussian ensemble mixture approach.

Now we utilize the noisy L-63 model (12) as a test model for the conditional Gaussian ensemble mixture idea in (19). Here we assume x is the observed process, while y and z are the unobserved ones. The tests with different noise in the observational process and different noise and in the filtering processes ranging from 1 to 10 reach similar qualitative conclusions and thus we only show the situation where . The initial distribution is assumed to be Gaussian with mean following Majda and Harlim (2012) and a small covariance .

Figure 9 shows the recovery of the first four central moments (i.e., mean, variance, skewness, and kurtosis) associated with the unobserved variable z with different L. For comparison, we also show the results by adopting the Monte Carlo simulation with a large ensemble number N = 50 000, which is regarded as the truth. Even with in (19), the short-term transitions in the mean, variance, and skewness are captured quite well. With ensembles, the leading four moments have already been recovered with high accuracy. If ensembles are adopted, then these statistics are recovered almost perfectly. The same results are found in variable y and thus they are omitted here.

Fig. 9.
Fig. 9.

(a)–(d) Recovering of the mean, variance, skewness, and kurtosis associated with the marginal PDF associated with the unobserved variable z in the noisy L-63 model (12) utilizing the conditional Gaussian ensemble mixture approach (19) with different number of ensembles L in a perfect model setting. As comparison, the recovered statistics utilizing Monte Carlo simulation with N = 50 000 ensemble members are also included. The green dot in (c) indicates the largest skewness in the transition phase, which will be utilized in Fig. 10.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

In Fig. 10a, we show the model error in (6) utilizing the conditional Gaussian ensemble mixture in recovering the marginal PDF of y at a short-term transition time , where the skewness arrives at its maximum. Clearly, the model error decays as L and it is already negligible with . The comparison of the marginal PDFs is shown in Fig. 10e, which validates the results in Fig. 10a. Figures 10b and 10f are the analogy to Figs. 10a and 10e for recovering the marginal PDF of z at the most skewed transition phase . Figures 10c and 10d show the model error dependence of L at the essentially statistical equilibrium phase (). Again, is a sufficient number for approximating the marginal PDFs with high skill.

Fig. 10.
Fig. 10.

The model error in (6) in recovering the marginal PDF associated with the unobserved variables y and z in the noisy L-63 model (12) utilizing the conditional Gaussian ensemble mixture approach in (19) with different L in a perfect model setting. (a) The model error in the PDF of y as a function of L at a short-term transition phase with largest skewness. (e) Comparing the PDFs of y at utilizing the conditional Gaussian mixture in (19) and Monte Carlo with N = 50 000. (c),(g) As in (a),(e), but at time , at which the system reaches statistical equilibrium state. (b),(d),(f),(h) The marginal distribution associated with the unobserved variable z.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

We have also tested the model error dependence on the ensemble number N utilizing Monte Carlo simulations. To reach a comparable skill with utilizing a conditional Gaussian ensemble mixture, the ensemble size utilizing the Monte Carlo simulation is around , which is much larger than L. Note that N increases dramatically with the dimension of the unobserved processes.

c. Recovering the time-dependent PDF of the unresolved variables with model error

Now we study recovering the time-dependent PDF of the unresolved variables in the noisy L-63 model with model error. The true signal associated with the observed variable x is generated from model (12) and the imperfect model with model error in the noise levels in (14) is utilized to recover the unobserved PDFs. Below, the effects of the model error due to both noise inflation and underdispersion are explored.

First, we take in the noisy L-63 model (12) to generate the true signal while the imperfect model for recovering the hidden PDFs in (14) are equipped with noise . Therefore, the noise is inflated in the imperfect forecast model (see Fig. 11). The recovered statistics associated with y are quite accurate utilizing the conditional Gaussian mixture approach in (19) with . On the other hand, there is an information barrier in the recovered PDF of z at a short-term transition time due to the underestimation of the skewness (see Figs. 11c,e). Despite the failure of capturing this non-Gaussian feature at the short transition time, the time-dependent mean and variance of z are recovered with high accuracy with and the equilibrium marginal distributions (Figs. 11f,g) associated with both y and z are estimated with almost no model error.

Fig. 11.
Fig. 11.

Recovery of the marginal PDFs associated with the unobserved variables y and z in the presence of model error from noise inflation. The noisy L-63 model (12) with is utilized in generating the true signal. The imperfect model (14) with is adopted for recovering the hidden PDFs. (a)–(d) The recovered mean, variance, skewness, and kurtosis compared with the truth that is computed by Monte Carlo simulation with N = 50 000 samples. (e) The recovered PDFs at short-term transition time with maximum skewness, where for y and for z. (f) The recovered PDFs at statistical equilibrium state for both y and z and (g) the PDFs at in logarithmic scale.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

Next, we take in the noisy L-63 model (12) to generate the true signal, while the imperfect model for recovering the hidden PDFs in (14) are equipped with noise . Therefore, model error comes from underdispersion in the imperfect forecast model. As shown in Fig. 12b, the marginal variance of both y and z is always underestimated. Therefore, the recovered marginal PDFs have smaller spreads than the truth, especially at a short-term transition phase for variable z (Fig. 12e). Moreover, even at the essentially statistical equilibrium state , obvious errors are found in the tails of the recovered PDFs (Fig. 12g), which implies that the probability of the extreme events is severely underestimated.

Fig. 12.
Fig. 12.

Recovery of the marginal PDFs associated with the unobserved variables y and z in the presence of model error from underdispersion of noise. The noisy L-63 model (12) with is utilized in generating the true signal. The imperfect model (14) with is adopted for recovering the hidden PDFs. (a)–(d) The recovered mean, variance, skewness, and kurtosis compared with the truth that is computed by Monte Carlo simulation with N = 50 000 samples. (e) The recovered PDFs at short-term transition time with maximum skewness, where for y and for z. (f) The recovered PDFs at statistical equilibrium state for both y and z and (g) the PDFs at in logarithmic scale.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

6. Parameter estimation

One of the important issues in many scientific and engineering areas is to estimate model parameters given noisy observations. Classical ways of estimating parameters includes maximum likelihood (Snijders 2011; Sowell 1992), Bayesian inference (Bretthorst 2013; Golightly and Wilkinson 2008; Chen et al. 2014a), and least squares methods (Marquardt 1963). One promising way for the real-time estimation of the parameters in turbulent systems is via filtering/data assimilation, in which the parameters are regarded as augmented state variables (Doucet and Tadić 2003; Beskos et al. 2006). Ensemble Kalman filter or particle filters are applied to estimate parameters in both drift and diffusion terms in stochastic processes (Golightly 2009; Doucet et al. 2000; Harlim et al. 2014). Here, we study the skill of estimating the parameters in the dynamics that have the following general form:
e21
where are the observed state variables and are the parameters to be estimated that are assumed to be constants. We also assume the system contains random noise, the amplitude of which is known. Evidently, the drift function can consist of polynomials or trigonometric polynomials, where the coefficient of each monomial is to be estimated. Note that both the dyad and triad systems in (8) and (11) belong to the model family in (21) provided that all the state variables are observed.
Since these parameters are constants, it is natural to augment the system (21) by an n-dimensional trivial equations for (Harlim et al. 2014; Smedstad and O’Brien 1991; Van Der Merwe and Wan 2001; Plett 2004; Wenzel et al. 2006). This forms the framework of the parameter estimation with direct approach:
e22a
e22b
Throughout this section, (with asterisk) always represents the true value of the parameters while stands for the variables in the parameter estimation framework.
In some applications, prior information about the possible range of the parameters is available. To incorporate such information into the parameter estimation framework, we augment the system (21) by a group of stochastic equations of (Majda and Harlim 2012), where the equilibrium distributions of these stochastic processes represent the prior information for the range of the parameters . This framework is called parameter estimation with stochastic parameterized equations:
e23a
e23b
Note that in addition to including the prior information into the parameter estimation framework, adding random perturbations into the evolution of the parameters also helps reduce sample attrition (Liu and West 2001).

Given an initial value and an initial uncertainty of each component of , both the augmented systems (22) and (23) belong to the conditional Gaussian framework in (1)(3), where and . Therefore, the time evolution of is solved via closed analytic formulas.

Below we aim at studying the dependence of the error and uncertainty on different factors, such as the noise in the system, the initial uncertainty, and the model structure, utilizing both the frameworks in (22) and (23). The important role of observability in parameter estimation will be emphasized. In addition, the difference of the skill in parameter estimation in linear and nonlinear problems will be explored. The detailed derivations associated with the propositions shown below are all included in appendix C.

a. Estimating one additive parameter in a linear scalar model

We start with estimating one additive parameter in the following linear scalar model:
e24
Given the initial guess of the parameter with initial uncertainty , the simple structure of model (24) allows for the analytic expression of the error in the posterior mean estimation and the posterior uncertainty as a function of time.
We start with estimating the additive parameter in (24) within the framework utilizing direct approach (22):
e25a
e25b
Proposition 2: In estimating the additive parameter in (24) within the framework utilizing direct approach (25), the posterior variance and the error in the posterior mean have the following closed analytical form:
e26a
e26b
According to (26), both the posterior uncertainty and the deterministic part of the error in posterior mean converge to zero asymptotically at an algebraic rate of time . The second term on its right-hand side of (26b) represents the stochastic fluctuation of the error that comes from the system noise. The variance of this fluctuation at time t is given by
eq7
where the asymptotic convergence rate of which is as well.

It is clear from (26) that decreasing the noise and increasing the prefactor helps accelerate the reduction of both the error and the uncertainty for long-term behavior. In fact, a nearly zero implies the system loses practical observability, which corresponds to a slow convergence rate. On the other hand, although increasing the initial uncertainty enhances the convergence rate of the deterministic part of , it has no effect on the long-term behavior of reducing either the uncertainty and the error in the fluctuation part of . In addition, a large leads to a large error in the fluctuation part of at the initial period.

Next, we study estimating in (24) within the framework utilizing the stochastic parameterized equations (23). To this end, we form the augmented system:
e27a
e27b
where the equilibrium distribution of γ in (27b) is Gaussian with mean and variance .
Proposition 3: In estimating the additive parameter in (24) within the framework utilizing stochastic parameterized equations (27), the posterior variance and the error in the posterior mean have the following closed analytical form:
e28a
e28b
where is assumed to be larger than in (28a) and are the two roots of the following algebraic equation:
eq8
In (28b), the variance is replaced by its equilibrium value for the conciseness of the expression due to its exponentially fast convergence rate.
Unlike (26b) where the error in the posterior mean estimation converges to zero eventually, the error utilizing the stochastic parameterized equation (28b) converges to
eq9
which is nonzero unless the mean of the stochastic parameterized equation (27b) (i.e., ) equals the true value of the parameter . Similarly, the posterior uncertainty converges to a nonzero value unless the right-hand side of (27b) disappears.

Yet, comparing the formulas in (26) and (28), it is obvious that the parameter estimation framework utilizing stochastic parameterized equations (27) leads to an exponential convergence rate for both the reduction of the posterior uncertainty and the error in the posterior mean, which implies a much shorter training phase is needed in the framework in (27). The convergence rate is controlled by the tuning factors in the stochastic parameterized equations. Thus, with a suitable choice of (27b), the convergence rate is greatly improved at the cost of only introducing a small bias in parameter estimation.

b. Estimating one multiplicative parameter in a linear model

Many applications require estimating parameters that appear as the multiplicative factors of the state variables. Here we study a simple situation where only one multiplicative parameter appears in the dynamics. Consider the following system:
e29
where we assume the parameter to guarantee the mean stability of the system. Given the initial guess of the parameter with initial uncertainty , the analytic expressions of the error and the uncertainty are still available in the framework utilizing direct approach (22). There is no simple closed expression for the error estimation in the framework utilizing stochastic parameterized equations (23) but numerical results will be provided for comparing the skill utilizing the two approaches.
The augmented system utilizing direct approach (22) has the following form:
e30
Proposition 4: In estimating the multiplicative parameter in (29) within the parameter estimation framework utilizing direct approach (30), the posterior variance and the error in the posterior mean have the following closed analytical form:
e31a
e31b
The long-term behavior of (31) can be further simplified. Apply the Reynold’s decomposition:
e32
where represents the ensemble mean of a random variable u at a fixed time t. Thus,
e33
Utilizing ergodicity, the two integrals on the right-hand side of (33) are given by
e34
respectively, where is the equilibrium Gaussian distribution associated with the system (29). Thus, the long-term behavior of (31) simplifies to
e35a
e35b
Similar to the situation in estimating one additive parameter in (26), the convergence of both the error and uncertainty in (35) is at an algebraic rate . However, the convergence strongly depends on the prefactor . When is zero, the denominator of the terms on the right-hand side of (35) becomes , which is independent of the noise amplitude . On the other hand, when is highly nonzero, decreasing the noise level accelerates the convergence. In fact, a nearly zero implies that the mean state of u is nearly zero and the system (30) has no practical observability. With a small noise, losing practical observability implies a much slower convergence.
Alternately, the augmented system utilizing stochastic parameterized equations (23) has the following form:
e36a
e36b
Since there are no simple closed formulas for the error and uncertainty in the posterior estimation, we show the numerical results utilizing the equations in (3) for estimating utilizing (36) and compare with those from (30).

In Fig. 13, we show the posterior mean and the posterior uncertainty in estimating the multiplicative parameter in (29) utilizing both the direct approach (30) and the stochastic parameterized equation (36). Here, the truth is . The constant factor in (29) is set to be such that the system has no practical observability. Different noise and initial uncertainty are chosen. To introduce an initial error, the initial value of γ in both (30) and (36) is set to be . When estimating γ utilizing the stochastic parameterized equation (36), the ratio is assumed such that there exists a bias in the equilibrium mean in (36b) and the equilibrium variance is also fixed. Thus, there is only one freedom in (36b), the inverse of which is the decorrelation time.

Fig. 13.
Fig. 13.

Comparison of estimating the multiplicative parameter in (29). (a)–(d) Estimation skill utilizing the direct approach in (30). (e)–(h) Estimation skill utilizing the stochastic parameterized equation (36). Here, the truth and are adopted. In the stochastic parameterized equation (36), the equilibrium mean and equilibrium are fixed. The black dotted line represents the truth ; the red curve is the posterior mean or posterior variance ; the two green curves around the posterior mean differs from the mean by one standard deviation (i.e., ).

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

We first look at the parameter estimation skill utilizing the direct approach (30). Since , the system has no practical observability and the convergence rate has no dependence on according to (35), which is validated by Figs. 13a–c. Clearly, the posterior uncertainty goes to zero but the error in the posterior mean is still above 0.5 even after nondimensional units. When the initial uncertainty decreases from (Fig. 13a) to (Fig. 13d), the convergence becomes slower as expected from (35). On the other hand, the convergence utilizing the stochastic parameterized equation (36b) (Figs. 13e–h) is much faster and it is almost unchanged by reducing the initial uncertainty (Fig. 13h). Although the equilibrium mean of the stochastic parameterized equation (36b) has a bias of 0.5 unit in γ, with the help of observations the averaged posterior mean at the equilibrium differs from the truth by only 0.1 to 0.2 unit. In addition, the posterior mean estimation is quite robust with respect to the choice of the coefficients , and in the stochastic parameterized equations (36b) as seen in Figs. 13e–g. Yet, overestimating (Fig. 13e) and underestimating (Fig. 13g) lead to the increase of fluctuations and the decrease of convergence, respectively. The optimal choice here is as shown in Fig. 13f.

We have so far focused on the parameter estimation skill in the appearance of one observational trajectory. In some applications, repeated experiments are available and, therefore, it is worthwhile studying the parameter estimation skill given an ensemble of independent observations. Assume the number of the independent observed trajectory is L. Corresponding to (30), the parameter estimation framework utilizing direct approach is given by
e37
where is a column vector, representing L-independent observations. All the entries in the column vector are equal to . Both and are diagonal matrices, where each diagonal entry of is .
Proposition 5: In estimating the multiplicative parameter in (29) within the parameter estimation framework utilizing direct approach (37) with L-independent observed trajectories, the posterior variance and the error in the posterior mean have the following closed analytical form:
e38a
e38b
Comparing (31) and (38), the asymptotic convergence with L-independent trajectories within the direct approach framework is enhanced by a multiplier L in front of t. Thus, increasing the number of independent observations accelerates the convergence, but the convergence rate remains algebraic.

c. Estimating parameters in cubic nonlinear models

From now on, we study the parameter estimation issue in nonlinear models. Our focus is on a model with cubic nonlinearity:
e39
where to guarantee the mean stability. The cubic model (39) is a special case of the normal form for the reduced stochastic climate model (Majda et al. 2009) and it is utilized as a test model for fluctuation-dissipation theorems in Majda et al. (2010a). The goal is to estimate the four parameters: .
To understand the underlying difference of estimating parameters in nonlinear and linear dynamics, we begin with a simplified version of (39):
e40
where the analytic formulas of the posterior uncertainty and the error in the posterior mean are available in the framework utilizing the direct approach (22):
e41a
e41b
Proposition 6: For any odd k, the framework utilizing direct approach (22) to estimate the parameter in
eq10
is given by
e42a
e42b
The posterior variance and the error in the posterior mean associated with system (42) have the following closed analytical form:
e43
Applying Reynold’s decomposition in (32), the integral can be rewritten as
e44
Regarding the cubic model (41), the index k in (42) and (44) is set to be . Further consider the situation with , which implies that the system loses practical observability with at the equilibrium. Clearly, the only nonzero term on the right-hand side of (44) at a long-term range is . Since , we utilize u to replace for notation simplicity. In light of the ergodicity of u:
eq11
where the analytic expression of the equilibrium PDF is given by Majda et al. (2009):
eq12
Direct calculation shows that
e45
where is the Gamma function (Abramowitz et al. 1965). Therefore, the posterior variance and the error in the posterior mean for the long-term behavior utilizing the direct approach (41) with have the following closed analytical form:
e46a
e46b
where the constant .

We compare the results in (46) of the cubic nonlinear system (40) with those in (35) of the linear system (29). The most significant difference is the role of the noise . In the linear model, without practical observability (i.e., ), the convergence rate has no dependence on . On the other hand, in the cubic nonlinear model, increasing the noise accelerates the convergence! This seems to be counterintuitive. However, the cubic nonlinearity, serving as the cubic damping in (40), indicates that the state variable u is trapped to the region around its attractor more severely than that in the linear model. Since the system has no practical observability around , an enhanced is preferred for increasing the amplitude of u and thus improves the parameter estimation skill.

Now we focus on the full cubic system (39) and estimate the four parameters in different dynamical regimes. Phase portrait analysis indicates that the deterministic part of (39) can have either 1) one stable equilibrium or 2) two stable equilibria and one unstable equilibrium. Here we fix the parameter and and consider the free parameters and . The phase space is divided into two separate regions with different dynamical behaviors, where the dividing curve between these two regimes can be written down analytically (Majda et al. 2009):
eq13
(see Fig. 14).
Fig. 14.
Fig. 14.

Phase portrait of for the deterministic part of cubic model with and fixed. The three red dots are the three examples utilized in Fig. 15 to study the parameter estimation skill.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

Below, we study the parameter estimation skill within the framework in (22) utilizing the direct approach in three dynamical regimes as shown in Fig. 14, where regime I and regime III correspond to one and three equilibria in phase portrait, respectively, and regime II has one equilibrium but the parameter values are near the dividing curve. Here a moderate noise is chosen. Figures 15a and 15b show the observed trajectories and equilibrium PDFs of u for the three regimes. The bimodal and nearly Gaussian PDFs for regimes I and III are due to the number of stable equilibria. The PDF for regime II is skewed where the one-sided extreme events in the trajectory increase the probability at the tail of the PDF. In the parameter estimation framework in (22), the initial value of each parameter is chosen to be two units smaller than the truth and the initial uncertainty is set to be .

Fig. 15.
Fig. 15.

Parameter estimation of the cubic model (39) in different dynamical regimes. (a) The observational trajectories with , and as shown in Fig. 14. The other parameters are , and . (b) The corresponding PDF. (c),(d) The estimated parameters and the associated estimation uncertainty of parameter c.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

The posterior mean and posterior uncertainty associated with the parameter c corresponding to three regimes are shown in Figs. 15c and 15d. A rapid convergence in both posterior mean and posterior uncertainty is found in regime I, where the two distinct states in the trajectory of u clearly indicate the dynamical behavior. In contrast, the convergence of the posterior uncertainty in regime III is quite slow and the posterior mean remains far from the truth even after nondimensional units. Such unskillful behavior is due to the fact that the dynamical structure is hard to be recovered from the noisy trajectory with a short memory. An interesting phenomenon is found in regime II. The convergence remains slow at short and medium lead times while a sudden uncertainty reduction occurs around , at which time an extreme event occurs in u. Such extreme events, despite having small probability, are important in conveying information of the dynamical structure.

It is worthwhile remarking that if the noise is too small in regime I, then the trajectory of u will be trapped in one attractor, which leads to an extremely slow convergence of the posterior uncertainty and a significant error in the posterior mean estimation. Thus, a moderately large noise helps enhance the parameter estimation skill in the model with cubic nonlinearity, which is consistent with the conclusions from the special case in (46).

Finally, to overcome the slow convergence of parameter estimation utilizing the direct approach (22) in regime III, we turn to the framework utilizing stochastic parameterized equation (23), which is given by
e47a
e47b
For each parameter, we set the mean of the stochastic parameterized equations (47b) to be 0.5 units larger than the truth, representing model error, and the equilibrium variance is assumed to be . The damping coefficient is set to be for all the four parameters. The comparison of estimating the four parameters utilizing the direct approach (22) and the stochastic parameterized equation (23) is shown in Fig. 16. Estimating the parameters utilizing the stochastic parameterized equations has a much faster convergence and the model error in the stochastic parameterized equation is alleviated with observations.
Fig. 16.
Fig. 16.

Parameter estimation of the cubic nonlinear model (39) in regime III with with (a)–(d) the direct approach and (e)–(h) the stochastic parameterized equation method. The black dotted line shows the truth of each parameter and the red dashed line shows the averaged value of the estimation of each parameter utilizing stochastic parameterized equation method at equilibrium. Here the equilibrium mean of stochastic parameterized equation has 0.5 unit bias from the truth. The equilibrium variance of each stochastic parameterized equation is . The damping coefficient in the stochastic parameterized equation is set to be .

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0437.1

7. Summary, conclusions, and discussion

In this paper, we study filtering the nonlinear turbulent dynamical system (1) through conditional Gaussian statistics. The special structure of the system allows closed the analytic form for the updates of the posterior states (section 2). Information measures (section 3) are adopted for assessing the model error and lack of information in filtering.

The role of energy-conserving nonlinear interactions in filtering the turbulent systems is studied in section 4 based on a dyad model (8). The lack of information in the stochastic parameterized filter (9) is large and the energy-conserving nonlinear feedback is found to be particularly important when the stochastic noise amplitude in the observed process is not negligible. The observability plays a key role with moderate and small in generating the true signal. Intermittency increases the signal-to-noise ratio, which helps improve the filtering skill.

The model error in the stochastic forcing amplitudes is studied in section 5 where the L-63 model (a triad model) is adopted as a test model. Both mathematical analysis (Proposition 1) and numerical experiments (Fig. 7) support that noise inflation leads to little error in filtering the unobserved trajectory while significant model errors are found in the imperfect filter due to underdispersion (Figs. 68). An efficient conditional Gaussian ensemble mixture method (19) is proposed in approximating the time-dependent PDF of the unobserved processes, which requires only small ensembles (Fig. 10) and can be generalized to systems with a large number of unobserved variables. This is an alternative to the traditional finite ensemble filter, which is usually quite expensive and sometimes suffers from sampling issues. Again, noise inflation in the imperfect model leads to only a small model error (Fig. 11) while underdispersion results in an obvious gap in estimating the PDF, where a severe underestimation of the variance implies the failure of capturing extreme events (Fig. 12).

The conditional Gaussian framework also allows for systematical study of parameter estimation skill, where the parameters are regarded as the augmented state variables. The convergence rate of the estimated parameters depends largely on the observability. Without practical observability, a slow convergence rate is found utilizing the direct parameter estimation approach (22) (Proposition 2, Proposition 4). On the other hand, a suitable choice of the stochastic parameterized equations for the augmented state variables in (23) leads to an exponentially fast convergence rate at the cost of only introducing a small error (Proposition 3; Fig. 13). In estimating parameters in a cubic nonlinear system, the convergence rate varies in different dynamical regimes utilizing the direct approach (22). The solutions converge to the truth very quickly in a bimodal regime while an extremely slow convergence is found in the nearly Gaussian regime (Fig. 15). Adopting the stochastic parameterized equations (23) again improves the skill of parameter estimation significantly (Fig. 16).

In addition to the three examples extensively discussed in sections 46 and those mentioned in section 1, there are many other potential applications of the conditional Gaussian filtering framework in (1)(3) in atmosphere and ocean sciences. One example is to adopt a simple linear ocean model and filter the ocean flows by fully observing a nonlinear noisy atmosphere, where the ocean–atmosphere coupling is via the noisy temperature flux at the surface with quadratic surface flux exchange. Another example involves recovering the temperature in the full Boussinesq system, where the temperature variable enters linearly into the observed momentum equation with noisy forcing. Furthermore, the conditional Gaussian filtering framework naturally applies to the equations of chemical reaction. In many of these topics, the unobserved dynamics are usually the reduced versions of their fully complex nonlinear physical systems. Yet, this is the typical strategy in dealing with the complicated atmosphere–ocean problems. In fact, the judicious choice of the simplified equations, which offsets part of the model error and the measurements error, has been justified to retain a high skill in many filtering issues (Kalnay 2003; Majda and Harlim 2012).

Finally, we comment on the three underlying assumptions for directly applying the conditional Gaussian filtering framework in (1)(3): (i) observations are taken continuously in time, (ii) there are no measurement errors beyond the observational uncertainty, and (iii) the state variables collected in are fully observed. In real applications, since the up-to-date satellites, radars, and drifters are all able to collect data with high frequency, it is reasonable to regard the observations from these sources be continuous in time for long-range forecasting. Therefore, the continuous-time conditional Gaussian filtering framework introduced and discussed in this work in (1)(3) fits the requirements of many practical data assimilation problems. Note that it is also straightforward to generalize this framework to the situation with discrete observations and state variables, which is actually included in (Liptser and Shiryaev 2001). In addition, although the conditional Gaussian filtering framework in (1)(3) does not explicitly take into consideration the observational noise or measurement errors, the filtering results by formally applying the framework in (1)(3) to the noisy observational data are captured in many practical applications (Chen et al. 2015; Chen and Majda 2016a) despite the fact that the filters become suboptimal. The constraint that the state variables are fully observed requires a careful choice of the observational variables. The velocity fields, the geopotential height, or the precipitation are the typical choices of in many real applications.

Developing a systematic framework for optimizing the stochastic parameterized equations will be useful for estimating parameters in more complex systems. The information-theoretic framework described in section 3 is a good candidate for this optimization, which remains as a future work. Other future works involve designing computationally affordable filters based on (3), following the guidelines provided in this work, for more complex turbulent systems. Noticeably, the conditional Gaussian framework in (1)(3) is also quite useful in studying the ensemble prediction skill and quantifying the uncertainty with model error.

Acknowledgments

This research of A.J.M is partially supported by the Office of Naval Research Grant ONR MURI N00014-12-1-0912. N.C. is supported as a graduate research assistant on this grant. A.J.M also gratefully acknowledges the generous support of the Center for Prototype Climate Modeling of NYU Abu Dhabi Research Institute grant.

APPENDIX A

Observability of Continuous Systems

Observability plays an important role in filtering the hidden variables from observations. Let us consider the linearized coupled observation-filtering system:
ea1
ea2
where and are the observational and filtering processes, respectively.
The observability (Gajic and Lelic 1996) of the system (A1) and (A2) can be derived as follows. Taking one more derivative with respect to (A1), with the help of (A2), yields
ea3
A similar argument applies for higher-order derivative of . Therefore, the augmented system is given by
ea4
A system is said to be observable if, for any possible sequence of the state (unobserved variable) and control quantities , the current state can be determined using only the observations . Therefore, the condition of the observability is that the rank of matrix equals the dimension of . In practice, due to the noise and numerical errors, the system is said to have no practical observability if the matrix is nearly singular.

a. Observability of the dyad model (8)

Let us linearize both u and υ around the mean states and :
eq14
The associated equations of (8) for the perturbed variables and are given by
ea5
Since in (A5) is a scalar, the observability matrix in (A4) becomes . Clearly, the dyad system in (8) loses its observability when . This implies the unobserved variable υ is decoupled from the observational process. In dynamical regime B with , the fixed point is , around which the system has no practical observability.

b. Observability of the L-63 model (12)

Again, we linearize , and z around their mean states , , and ,
eq15
The associated equations of the L-63 model (12) for the perturbed variables are given by
ea6
If x is the observed variables and y and z are the filtering variables, then corresponding to (A1) and (A2), , , and
eq16
According to (A4), the observability matrix is given by
eq17
Since is given and is nonzero, the system loses observability when . It is also clear that when the system loses observability, the second column of the observability matrix becomes zero and therefore observations provide no information in filtering the variable z.

On the other hand, if the observed variables are y and z and the filtering variable is x, direct calculations show that the system has no observability when and .

APPENDIX B

Detailed Derivation of Proposition 1 in Section 5a of Triad Models

In the situation of observing y and z while filtering x, the posterior variance utilizing the forecast model (14) is given by, according to (3),
eb1
Note that the variance remains nonnegative in (B1). Clearly,
eq18
and therefore
eq19
If we formally write (B1) as
eb2
where
eq20
then it is obvious that the convergence rate of the posterior covariance to the equilibrium is faster than . Actually, the solution of is bounded by
eb3
where the right-hand side of (B3) is the solution of the following equation:
eq21
If the noise in the filter model (14) is zero, then the posterior variance converges to zero in the exponential rate.
The posterior mean evolution can be written down explicitly as
eb4
Recall y and z equation in the perfect model (15):
eb5
Therefore, inserting (B5) into (B4) leads to
eb6
In addition, note the x equation of the perfect model (15) is given by
eb7
Subtracting (B7) from (B6) leads to
eb8
The error equation is given by
eb9
Since both are are nonnegative, the error is bounded by
eb10
which decays to zero in an exponential rate.

APPENDIX C

Detailed Derivations for Propositions 2–4 in Sections 6a and 6b of Parameter Estimation in Linear Models

a. Detailed derivations of Proposition 2

We aim at estimating the additive parameter in the linear system:
ec1
The augmented system for estimating utilizing direct approach (22) is given by
eca2a
ecb2b
where the initial guess and the initial uncertainty are assigned. In light of (3), the evolutions of posterior mean and posterior variance of γ have the following form:
eca3a
ecb3b
The solution of in (C3b) is reached by separation of variables:
ec4
To calculate the error in the posterior mean compared with the constant truth , we first rewrite (C3b) as
ec5
Since u in (C5) is from the true observations, we insert (C1) into (C5):
ec6
With the expression of the variance in (C4), we have
ec7
For the simplicity of notation, we define
eq22
Then (C7) becomes
ec8
Applying the method of integrating factor, we have
ec9
Changing back to the original notations leads to
ec10

b. Detailed derivations of Proposition 3

Now we estimate the additive parameter in (C1) utilizing stochastic parameterized equation method (23):
eca11a
ecb11b
where is to guarantee the mean stability of (C11b). The evolutions of the posterior mean and variance of γ have the closed form, according to (3):
eca12a
ecb12b
Clearly, for and , the algebraic equation
ec13
always having two real roots with different signs. Let us assume and initial value . Utilizing separation of variables, the posterior variance is solved,
ec14
This implies will converge to the equilibrium state in an exponential way.
To solve the error in the posterior mean, we use the equilibrium variance to replace in (C12a) due to the fact that converges exponentially fast to . The qualitative conclusion does not change if we keep in (C12a) but the expression will becomes extremely complicated. Again, noticing the true value is a constant and making use of the true dynamics in (C1), the error in the posterior mean in (C12a) becomes
ec15
Utilizing the integrating factor method, we arrive at the following solution:
ec16

c. Detailed derivations of Proposition 4

Now we estimate the multiplicative parameter in the linear system
ec17
The augmented system by utilizing the direct method in (22) yields
eca18a
ecb18b
The evolutions of mean and variance of γ are given by, according to (3),
eca19a
ecb19b
In light of the method of separation of variables, (C19b) leads to the solution for the posterior variance:
ec20
To solve the error in the posterior mean, we make use of (C17), (C19a), and (C20):
ec21
For the simplicity of notation, we again define
eq23
Then (C21) becomes
ec22
The solution of (C22) is given by
ec23
Note that
eq24
Therefore, (C23) reduces to
ec24
Changing back to the original notations, (C24) leads to
ec25

APPENDIX D

Detailed Derivations of Proposition 6 in Section 6c of Estimating the Multiplicative Parameter in the Special Cubic System Utilizing the Direct Approach

The derivation of (43) in Proposition 6c follows those in (C20) and (C25). Here, we derive (45).

Recall the analytic expression of the equilibrium PDF is given by Majda et al. (2009):
eq25
The integral factor is given by
ed1
Let
ed2
and correspondingly
ed3
Inserting (D3) into (D1) results in
ed4
Recall the definition of function (Abramowitz et al. 1965):
ed5
Then (D4) becomes
eq26
This leads to
ed6
With the formula of in (D6), we are able to solve :
eq27

APPENDIX E

Information-Theoretic Framework in Quantifying the Model Error in Ensemble Prediction

As mentioned in section 3, the information-theoretic framework also provides a general framework in quantifying the uncertainty using imperfect models in ensemble prediction. This appendix includes the details.

Assume the joint distributions regarding and in (1) for perfect and imperfect models are given by
eq28
where due to the incomplete knowledge or the coarse-grained effect the distribution associated with the imperfect model is assumed to be formed only by the conditional moments up to L. According to Branicki et al. (2013), the lack of information in the imperfect model related to the perfect one is given by
ee1
where is the PDF reconstructed using the L moments of the perfect model. The first term on the right-hand side of (E1) is called the intrinsic barrier, which measures the lack of information in the perfect model due to the coarse-grained effect from the insufficient measurement. This intrinsic barrier cannot be overcome by choosing the models in the family that contains only by the conditional moments up to L. The second term on the right-hand side of (E1) is the model error using the imperfect model, which is aimed at being minimized in order to find the optimal imperfect model. Direct calculation shows that
ee2
ee3
In the conditional Gaussian framework, and the relative entropy for the conditional Gaussian distributions in (E3) is assessed in light of the closed analytic formulas in (3) for both the distributions. Note that in filtering complex turbulent systems, if the observations in the imperfect filter are assumed to be the same as in the perfect filter, then the lack of information in the imperfect filter related to the perfect one is simply assessed by
ee4
The information measurement in (E4) provides a guideline in designing practical imperfect filters. An example of applying (E4) to assess the information model error in different imperfect filters is shown in Chen and Majda (2016b) for filtering a turbulent flow field using noisy Lagrangian tracers.

REFERENCES

  • Abramov, R. V., and A. J. Majda, 2004: Quantifying uncertainty for non-Gaussian ensembles in complex systems. SIAM J. Sci. Comput., 26, 411447, doi:10.1137/S1064827503426310.

    • Search Google Scholar
    • Export Citation
  • Abramowitz, M., and I. A. Stegun, Eds., 1965: Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables. Dover, 1046 pp.

  • Anderson, J. L., 2001: An ensemble adjustment Kalman filter for data assimilation. Mon. Wea. Rev., 129, 28842903, doi:10.1175/1520-0493(2001)129<2884:AEAKFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bensoussan, A., 1992: Stochastic Control of Partially Observed Systems. Cambridge University Press, 353 pp.

  • Berner, J., and G. Branstator, 2007: Linear and nonlinear signatures in the planetary wave dynamics of an AGCM: Probability density functions. J. Atmos. Sci., 64, 117136, doi:10.1175/JAS3822.1.

    • Search Google Scholar
    • Export Citation
  • Beskos, A., O. Papaspiliopoulos, G. O. Roberts, and P. Fearnhead, 2006: Exact and computationally efficient likelihood-based estimation for discretely observed diffusion processes (with discussion). J. Roy. Stat. Soc. Series B Stat. Methodol., 68, 333382, doi:10.1111/j.1467-9868.2006.00552.x.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2013: Dynamic stochastic superresolution of sparsely observed turbulent systems. J. Comput. Phys., 241, 333363, doi:10.1016/j.jcp.2012.11.037.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2014: Quantifying Bayesian filter performance for turbulent dynamical systems through information theory. Commun. Math. Sci., 12, 901978, doi:10.4310/CMS.2014.v12.n5.a6.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2015: An information-theoretic framework for improving imperfect predictions via Multi-Model Ensemble forecasts. J. Nonlinear Sci., 25, 489538, doi:10.1007/s00332-015-9233-1.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., N. Chen, and A. J. Majda, 2013: Non-Gaussian test models for prediction and state estimation with model errors. Chin. Ann. Math. Ser. B, 34, 2964, doi:10.1007/s11401-012-0759-3.

    • Search Google Scholar
    • Export Citation
  • Bretthorst, G. L., 2013: Bayesian Spectrum Analysis and Parameter Estimation.Vol. 48, Springer Science & Business Media, 209 pp.

  • Cai, D., R. Kleeman, and A. Majda, 2002: A mathematical framework for quantifying predictability through relative entropy. Methods Appl. Anal., 9, 425444, doi:10.4310/MAA.2002.v9.n3.a8.

    • Search Google Scholar
    • Export Citation
  • Campillo, F., and V. Rossi, 2009: Convolution particle filter for parameter estimation in general state-space models. IEEE Trans. Aerospace Electron. Syst., 45, 10631072.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2015a: Predicting the cloud patterns for the boreal summer intraseasonal oscillation through a low-order stochastic model. Math. Climate Wea. Forecasting, 1, 120.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2015b: Predicting the real-time multivariate Madden–Julian oscillation index through a low-order nonlinear stochastic model. Mon. Wea. Rev., 143, 21482169, doi:10.1175/MWR-D-14-00378.1.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2016a: Filtering the stochastic skeleton model for the Madden–Julian oscillation. Mon. Wea. Rev., 144, 501527, doi:10.1175/MWR-D-15-0261.1.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2016b: Model error in filtering random compressible flows utilizing noisy Lagrangian tracers. Mon. Wea. Rev., doi:10.1175/MWR-D-15-0438.1, in press.

    • Search Google Scholar
    • Export Citation
  • Chen, N., D. Giannakis, R. Herbei, and A. J. Majda, 2014a: An MCMC algorithm for parameter estimation in signals with hidden intermittent instability. SIAM/ASA J. Uncertainty Quant., 2, 647669, doi:10.1137/130944977.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and D. Giannakis, 2014b: Predicting the cloud patterns of the Madden-Julian oscillation through a low-order nonlinear stochastic model. Geophys. Res. Lett., 41, 56125619, doi:10.1002/2014GL060876.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and X. T. Tong, 2014c: Information barriers for noisy Lagrangian tracers in filtering random incompressible flows. Nonlinearity, 27, 2133, doi:10.1088/0951-7715/27/9/2133.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and X. T. Tong, 2015: Noisy Lagrangian tracers for filtering random rotating compressible flows. J. Nonlinear Sci., 25, 451488, doi:10.1007/s00332-014-9226-5.

    • Search Google Scholar
    • Export Citation
  • Cressie, N., and C. K. Wikle, 2011: Statistics for Spatio-Temporal Data. John Wiley & Sons, 624 pp.

  • Daum, F., and J. Huang, 2003: Curse of dimensionality and particle filters. Proc. IEEE Aerospace Conf., Vol. 4, Big Sky, MT, Institute of Electrical and Electronics Engineers, doi:10.1109/AERO.2003.1235126.

  • Dee, D. P., 1995: On-line estimation of error covariance parameters for atmospheric data assimilation. Mon. Wea. Rev., 123, 11281145, doi:10.1175/1520-0493(1995)123<1128:OLEOEC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., and V. B. Tadić, 2003: Parameter estimation in general state-space models using particle methods. Ann. Inst. Stat. Math., 55 (2), 409422.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., S. Godsill, and C. Andrieu, 2000: On sequential Monte Carlo sampling methods for Bayesian filtering. Stat. Comput., 10, 197208, doi:10.1023/A:1008935410038.

    • Search Google Scholar
    • Export Citation
  • Frisch, U., 1995: Turbulence: The Legacy of A. N. Kolmogorov. Cambridge University Press, 296 pp.

  • Gajic, Z., and M. Lelic, 1996: Modern Control Systems Engineering. Prentice-Hall, Inc., 448 pp.

  • Gneiting, T., and A. E. Raftery, 2007: Strictly proper scoring rules, prediction, and estimation. J. Amer. Stat. Assoc., 102, 359378, doi:10.1198/016214506000001437.

    • Search Google Scholar
    • Export Citation
  • Golightly, A., 2009: Bayesian filtering for jump-diffusions with application to stochastic volatility. J. Comput. Graph. Stat., 18 (2), 384400.

    • Search Google Scholar
    • Export Citation
  • Golightly, A., and D. J. Wilkinson, 2008: Bayesian inference for nonlinear multivariate diffusion models observed with error. Comput. Stat. Data Anal., 52, 16741693, doi:10.1016/j.csda.2007.05.019.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. Majda, 2008: Filtering nonlinear dynamical systems with linear stochastic models. Nonlinearity, 21, 1281, doi:10.1088/0951-7715/21/6/008.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. J. Majda, 2010a: Catastrophic filter divergence in filtering nonlinear dissipative systems. Commun. Math. Sci., 8, 2743, doi:10.4310/CMS.2010.v8.n1.a3.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. J. Majda, 2010b: Filtering turbulent sparsely observed geophysical flows. Mon. Wea. Rev., 138, 10501083, doi:10.1175/2009MWR3113.1.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., A. Mahdi, and A. J. Majda, 2014: An ensemble Kalman filter for statistical estimation of physics constrained nonlinear regression models. J. Comput. Phys., 257, 782812, doi:10.1016/j.jcp.2013.10.025.

    • Search Google Scholar
    • Export Citation
  • Hinze, J. O., and J. Hinze, 1959: Turbulence: An Introduction to Its Mechanism and Theory. Vol. 162, McGraw-Hill, 586 pp.

  • Hyndman, R. J., and A. B. Koehler, 2006: Another look at measures of forecast accuracy. Int. J. Forecasting, 22, 679688, doi:10.1016/j.ijforecast.2006.03.001.

    • Search Google Scholar
    • Export Citation
  • Iglesias, M. A., K. Lin, and A. M. Stuart, 2014: Well-posed Bayesian geometric inverse problems arising in subsurface flow. Inverse Prob., 30 (11), 114001, doi:10.1088/0266-5611/30/11/114001.

    • Search Google Scholar
    • Export Citation
  • Kalman, R. E., and R. S. Bucy, 1961: New results in linear filtering and prediction theory. J. Basic Eng., 83, 95108, doi:10.1115/1.3658902.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., 2003: Atmospheric Modeling, Data Assimilation, and Predictability. Cambridge University Press, 368 pp.

  • Keating, S. R., A. J. Majda, and K. S. Smith, 2012: New methods for estimating ocean eddy heat transport using satellite altimetry. Mon. Wea. Rev., 140, 17031722, doi:10.1175/MWR-D-11-00145.1.

    • Search Google Scholar
    • Export Citation
  • Kondrashov, D., S. Kravtsov, A. W. Robertson, and M. Ghil, 2005: A hierarchy of data-based ENSO models. J. Climate, 18, 44254444, doi:10.1175/JCLI3567.1.

    • Search Google Scholar
    • Export Citation
  • Kravtsov, S., D. Kondrashov, and M. Ghil, 2005: Multilevel regression modeling of nonlinear processes: Derivation and applications to climatic variability. J. Climate, 18, 44044424, doi:10.1175/JCLI3544.1.

    • Search Google Scholar
    • Export Citation
  • Liptser, R. S., and A. N. Shiryaev, 1977: Statistics of Random Processes I: General Theory. Springer. 402 pp.

  • Liptser, R. S., and A. N. Shiryaev, 2001: Statistics of Random Processes II: Applications. Springer, 402 pp.

  • Liu, J., and M. West, 2001: Combined parameter and state estimation in simulation-based filtering. Sequential Monte Carlo Methods in Practice, A. Doucet, N. de Freitas, and N. Gordon, Eds., Springer, 197–223.

  • Lorenz, E. N., 1963: Deterministic nonperiodic flow. J. Atmos. Sci., 20, 130141, doi:10.1175/1520-0469(1963)020<0130:DNF>2.0.CO;2.

  • MacKay, D. J., 2003: Information Theory, Inference and Learning Algorithms. Cambridge University Press, 640 pp.

  • Majda, A. J., 2003: Introduction to PDEs and Waves for the Atmosphere and Ocean. Courant Lecture Notes Series, Vol. 9, American Mathematical Society, 234 pp.

  • Majda, A. J., 2012: Challenges in climate science and contemporary applied mathematics. Commun. Pure Appl. Math., 65, 920948, doi:10.1002/cpa.21401.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., 2015: Statistical energy conservation principle for inhomogeneous turbulent dynamical systems. Proc. Natl. Acad. Sci. USA, 112, 89378941, doi:10.1073/pnas.1510465112.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and X. Wang, 2006: Nonlinear Dynamics and Statistical Theories for Basic Geophysical Flows. Cambridge University Press, 564 pp.

  • Majda, A. J., and B. Gershgorin, 2010: Quantifying uncertainty in climate change science through empirical information theory. Proc. Natl. Acad. Sci. USA, 107, 14 95814 963, doi:10.1073/pnas.1007009107.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and M. Branicki, 2012: Lessons in uncertainty quantification for turbulent dynamical systems. Discrete Cont. Dyn. Syst., 32, 31333221, doi:10.3934/dcds.2012.32.3133.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and J. Harlim, 2012: Filtering Complex Turbulent Systems. Cambridge University Press, 357 pp.

  • Majda, A. J., and Y. Yuan, 2012: Fundamental limitations of ad hoc linear and quadratic multi-level regression models for physical systems. Discrete Cont. Dyn. Syst., 17, 13331363, doi:10.3934/dcdsb.2012.17.1333.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and B. Gershgorin, 2013: Elementary models for turbulent diffusion with complex physical features: Eddy diffusivity, spectrum and intermittency. Philos. Trans. A Math. Phys. Eng. Sci., 371, 20120184, doi:10.1098/rsta.2012.0184.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and J. Harlim, 2013: Physics constrained nonlinear regression models for time series. Nonlinearity, 26, 201271, doi:10.1088/0951-7715/26/1/201.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and I. Grooms, 2014: New perspectives on superparameterization for geophysical turbulence. J. Comput. Phys., 271, 6077, doi:10.1016/j.jcp.2013.09.014.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and Y. Lee, 2014: Conceptual dynamical models for turbulence. Proc. Natl. Acad. Sci. USA, 111, 65486553, doi:10.1073/pnas.1404914111.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden Eijnden, 1999: Models for stochastic climate prediction. Proc. Natl. Acad. Sci. USA, 96, 14 68714 691, doi:10.1073/pnas.96.26.14687.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden Eijnden, 2001: A mathematical framework for stochastic climate models. Commun. Pure Appl. Math., 54, 891974, doi:10.1002/cpa.1014.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden-Eijnden, 2002: A priori tests of a stochastic mode reduction strategy. Physica D, 170, 206252, doi:10.1016/S0167-2789(02)00578-X.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden-Eijnden, 2003: Systematic strategies for stochastic mode reduction in climate. J. Atmos. Sci., 60, 17051722, doi:10.1175/1520-0469(2003)060<1705:SSFSMR>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., R. V. Abramov, and M. J. Grote, 2005: Information Theory and Stochastics for Multiscale Nonlinear Systems. CRM Monogr. Series, Vol. 25, American Mathematical Society, 133 pp.

  • Majda, A. J., C. Franzke, and D. Crommelin, 2009: Normal forms for reduced stochastic climate models. Proc. Natl. Acad. Sci. USA, 106, 36493653, doi:10.1073/pnas.0900173106.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., R. Abramov, and B. Gershgorin, 2010a: High skill in low-frequency climate response through fluctuation dissipation theorems despite structural instability. Proc. Natl. Acad. Sci. USA, 107, 581586, doi:10.1073/pnas.0912997107.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., J. Harlim, and B. Gershgorin, 2010b: Mathematical strategies for filtering turbulent dynamical systems. Discrete Cont. Dyn. Syst., 27, 441486, doi:10.3934/dcds.2010.27.441.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., D. Qi, and T. P. Sapsis, 2014: Blended particle filters for large-dimensional chaotic dynamical systems. Proc. Natl. Acad. Sci. USA, 111, 75117516, doi:10.1073/pnas.1405675111.

    • Search Google Scholar
    • Export Citation
  • Marquardt, D. W., 1963: An algorithm for least-squares estimation of nonlinear parameters. J. Soc. Ind. Appl. Math., 11, 431441, doi:10.1137/0111030.

    • Search Google Scholar
    • Export Citation
  • Neelin, J. D., B. R. Lintner, B. Tian, Q. Li, L. Zhang, P. K. Patra, M. T. Chahine, and S. N. Stechmann, 2010: Long tails in deep columns of natural and anthropogenic tropospheric tracers. Geophys. Res. Lett., 37, L05804, doi:10.1029/2009GL041726.

    • Search Google Scholar
    • Export Citation
  • Palmer, T. N., 2001: A nonlinear dynamical perspective on model error: A proposal for non-local stochastic-dynamic parametrization in weather and climate prediction models. Quart. J. Roy. Meteor. Soc., 127, 279304, doi:10.1002/qj.49712757202.

    • Search Google Scholar
    • Export Citation
  • Plett, G. L., 2004: Extended Kalman filtering for battery management systems of LiPB-based HEV battery packs: Part 3. State and parameter estimation. J. Power Sources, 134, 277292, doi:10.1016/j.jpowsour.2004.02.033.

    • Search Google Scholar
    • Export Citation
  • Qi, D., and A. J. Majda, 2015: Blended particle methods with adaptive subspaces for filtering turbulent dynamical systems. Physica D, 298–299, 2141, doi:10.1016/j.physd.2015.02.002.

    • Search Google Scholar
    • Export Citation
  • Salamon, P., and L. Feyen, 2009: Assessing parameter, precipitation, and predictive uncertainty in a distributed hydrological model using sequential data assimilation with the particle filter. J. Hydrol., 376, 428442, doi:10.1016/j.jhydrol.2009.07.051.

    • Search Google Scholar
    • Export Citation
  • Salmon, R., 1998: Lectures on Geophysical Fluid Dynamics. Oxford University Press, 400 pp.

  • Sapsis, T. P., and A. J. Majda, 2013a: Blending modified Gaussian closure and non-Gaussian reduced subspace methods for turbulent dynamical systems. J. Nonlinear Sci., 23, 10391071, doi:10.1007/s00332-013-9178-1.

    • Search Google Scholar
    • Export Citation
  • Sapsis, T. P., and A. J. Majda, 2013b: Statistically accurate low-order models for uncertainty quantification in turbulent dynamical systems. Proc. Natl. Acad. Sci. USA, 110, 13 70513 710, doi:10.1073/pnas.1313065110.

    • Search Google Scholar
    • Export Citation
  • Slivinski, L., E. Spiller, A. Apte, and B. Sandstede, 2015: A hybrid particle–ensemble Kalman filter for Lagrangian data assimilation. Mon. Wea. Rev., 143, 195211, doi:10.1175/MWR-D-14-00051.1.

    • Search Google Scholar
    • Export Citation
  • Smedstad, O. M., and J. J. O’Brien, 1991: Variational data assimilation and parameter estimation in an equatorial Pacific Ocean model. Prog. Oceanogr., 26, 179241, doi:10.1016/0079-6611(91)90002-4.

    • Search Google Scholar
    • Export Citation
  • Snijders, T. A. B., 2011: Multilevel analysis. International Encyclopedia of Statistical Science, M. Lovric, Ed., Springer, 879–882, doi:10.1007/978-3-642-04898-2_387.

  • Sowell, F., 1992: Maximum likelihood estimation of stationary univariate fractionally integrated time series models. J. Econometrics, 53, 165188, doi:10.1016/0304-4076(92)90084-5.

    • Search Google Scholar
    • Export Citation
  • Tong, X. T., A. J. Majda, and D. Kelly, 2016: Nonlinear stability of the ensemble Kalman filter with adaptive covariance inflation. Commun. Math. Sci., 14, 12831313, doi:10.4310/CMS.2016.v14.n5.a5.

    • Search Google Scholar
    • Export Citation
  • Townsend, A. A., 1980: The Structure of Turbulent Shear Flow. Cambridge University Press, 429 pp.

  • Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. Cambridge University Press, 745 pp.

  • Van Der Merwe, R., and E. Wan, 2001: The square-root unscented Kalman filter for state and parameter-estimation. IEEE Int. Conf. on Acoustics, Speech, and Signal Processing, 2001 (ICASSP’01). Vol. 6, Salt Lake City, UT, Institute of Electrical and Electronics Engineers, 3461–3464, doi:10.1109/ICASSP.2001.940586.

  • Wenzel, T. A., K. Burnham, M. Blundell, and R. Williams, 2006: Dual extended Kalman filter for vehicle state and parameter estimation. Veh. Syst. Dyn., 44, 153171, doi:10.1080/00423110500385949.

    • Search Google Scholar
    • Export Citation
  • Wikle, C. K., and M. B. Hooten, 2010: A general science-based framework for dynamical spatio-temporal models. TEST, 19, 417451, doi:10.1007/s11749-010-0209-z.

    • Search Google Scholar
    • Export Citation
  • Yeh, W. W.-G., 1986: Review of parameter identification procedures in groundwater hydrology: The inverse problem. Water Resour. Res., 22, 95108, doi:10.1029/WR022i002p00095.

    • Search Google Scholar
    • Export Citation
Save
  • Abramov, R. V., and A. J. Majda, 2004: Quantifying uncertainty for non-Gaussian ensembles in complex systems. SIAM J. Sci. Comput., 26, 411447, doi:10.1137/S1064827503426310.

    • Search Google Scholar
    • Export Citation
  • Abramowitz, M., and I. A. Stegun, Eds., 1965: Handbook of Mathematical Functions: With Formulas, Graphs, and Mathematical Tables. Dover, 1046 pp.

  • Anderson, J. L., 2001: An ensemble adjustment Kalman filter for data assimilation. Mon. Wea. Rev., 129, 28842903, doi:10.1175/1520-0493(2001)129<2884:AEAKFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bensoussan, A., 1992: Stochastic Control of Partially Observed Systems. Cambridge University Press, 353 pp.

  • Berner, J., and G. Branstator, 2007: Linear and nonlinear signatures in the planetary wave dynamics of an AGCM: Probability density functions. J. Atmos. Sci., 64, 117136, doi:10.1175/JAS3822.1.

    • Search Google Scholar
    • Export Citation
  • Beskos, A., O. Papaspiliopoulos, G. O. Roberts, and P. Fearnhead, 2006: Exact and computationally efficient likelihood-based estimation for discretely observed diffusion processes (with discussion). J. Roy. Stat. Soc. Series B Stat. Methodol., 68, 333382, doi:10.1111/j.1467-9868.2006.00552.x.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2013: Dynamic stochastic superresolution of sparsely observed turbulent systems. J. Comput. Phys., 241, 333363, doi:10.1016/j.jcp.2012.11.037.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2014: Quantifying Bayesian filter performance for turbulent dynamical systems through information theory. Commun. Math. Sci., 12, 901978, doi:10.4310/CMS.2014.v12.n5.a6.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., and A. J. Majda, 2015: An information-theoretic framework for improving imperfect predictions via Multi-Model Ensemble forecasts. J. Nonlinear Sci., 25, 489538, doi:10.1007/s00332-015-9233-1.

    • Search Google Scholar
    • Export Citation
  • Branicki, M., N. Chen, and A. J. Majda, 2013: Non-Gaussian test models for prediction and state estimation with model errors. Chin. Ann. Math. Ser. B, 34, 2964, doi:10.1007/s11401-012-0759-3.

    • Search Google Scholar
    • Export Citation
  • Bretthorst, G. L., 2013: Bayesian Spectrum Analysis and Parameter Estimation.Vol. 48, Springer Science & Business Media, 209 pp.

  • Cai, D., R. Kleeman, and A. Majda, 2002: A mathematical framework for quantifying predictability through relative entropy. Methods Appl. Anal., 9, 425444, doi:10.4310/MAA.2002.v9.n3.a8.

    • Search Google Scholar
    • Export Citation
  • Campillo, F., and V. Rossi, 2009: Convolution particle filter for parameter estimation in general state-space models. IEEE Trans. Aerospace Electron. Syst., 45, 10631072.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2015a: Predicting the cloud patterns for the boreal summer intraseasonal oscillation through a low-order stochastic model. Math. Climate Wea. Forecasting, 1, 120.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2015b: Predicting the real-time multivariate Madden–Julian oscillation index through a low-order nonlinear stochastic model. Mon. Wea. Rev., 143, 21482169, doi:10.1175/MWR-D-14-00378.1.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2016a: Filtering the stochastic skeleton model for the Madden–Julian oscillation. Mon. Wea. Rev., 144, 501527, doi:10.1175/MWR-D-15-0261.1.

    • Search Google Scholar
    • Export Citation
  • Chen, N., and A. J. Majda, 2016b: Model error in filtering random compressible flows utilizing noisy Lagrangian tracers. Mon. Wea. Rev., doi:10.1175/MWR-D-15-0438.1, in press.

    • Search Google Scholar
    • Export Citation
  • Chen, N., D. Giannakis, R. Herbei, and A. J. Majda, 2014a: An MCMC algorithm for parameter estimation in signals with hidden intermittent instability. SIAM/ASA J. Uncertainty Quant., 2, 647669, doi:10.1137/130944977.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and D. Giannakis, 2014b: Predicting the cloud patterns of the Madden-Julian oscillation through a low-order nonlinear stochastic model. Geophys. Res. Lett., 41, 56125619, doi:10.1002/2014GL060876.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and X. T. Tong, 2014c: Information barriers for noisy Lagrangian tracers in filtering random incompressible flows. Nonlinearity, 27, 2133, doi:10.1088/0951-7715/27/9/2133.

    • Search Google Scholar
    • Export Citation
  • Chen, N., A. J. Majda, and X. T. Tong, 2015: Noisy Lagrangian tracers for filtering random rotating compressible flows. J. Nonlinear Sci., 25, 451488, doi:10.1007/s00332-014-9226-5.

    • Search Google Scholar
    • Export Citation
  • Cressie, N., and C. K. Wikle, 2011: Statistics for Spatio-Temporal Data. John Wiley & Sons, 624 pp.

  • Daum, F., and J. Huang, 2003: Curse of dimensionality and particle filters. Proc. IEEE Aerospace Conf., Vol. 4, Big Sky, MT, Institute of Electrical and Electronics Engineers, doi:10.1109/AERO.2003.1235126.

  • Dee, D. P., 1995: On-line estimation of error covariance parameters for atmospheric data assimilation. Mon. Wea. Rev., 123, 11281145, doi:10.1175/1520-0493(1995)123<1128:OLEOEC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., and V. B. Tadić, 2003: Parameter estimation in general state-space models using particle methods. Ann. Inst. Stat. Math., 55 (2), 409422.

    • Search Google Scholar
    • Export Citation
  • Doucet, A., S. Godsill, and C. Andrieu, 2000: On sequential Monte Carlo sampling methods for Bayesian filtering. Stat. Comput., 10, 197208, doi:10.1023/A:1008935410038.

    • Search Google Scholar
    • Export Citation
  • Frisch, U., 1995: Turbulence: The Legacy of A. N. Kolmogorov. Cambridge University Press, 296 pp.

  • Gajic, Z., and M. Lelic, 1996: Modern Control Systems Engineering. Prentice-Hall, Inc., 448 pp.

  • Gneiting, T., and A. E. Raftery, 2007: Strictly proper scoring rules, prediction, and estimation. J. Amer. Stat. Assoc., 102, 359378, doi:10.1198/016214506000001437.

    • Search Google Scholar
    • Export Citation
  • Golightly, A., 2009: Bayesian filtering for jump-diffusions with application to stochastic volatility. J. Comput. Graph. Stat., 18 (2), 384400.

    • Search Google Scholar
    • Export Citation
  • Golightly, A., and D. J. Wilkinson, 2008: Bayesian inference for nonlinear multivariate diffusion models observed with error. Comput. Stat. Data Anal., 52, 16741693, doi:10.1016/j.csda.2007.05.019.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. Majda, 2008: Filtering nonlinear dynamical systems with linear stochastic models. Nonlinearity, 21, 1281, doi:10.1088/0951-7715/21/6/008.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. J. Majda, 2010a: Catastrophic filter divergence in filtering nonlinear dissipative systems. Commun. Math. Sci., 8, 2743, doi:10.4310/CMS.2010.v8.n1.a3.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., and A. J. Majda, 2010b: Filtering turbulent sparsely observed geophysical flows. Mon. Wea. Rev., 138, 10501083, doi:10.1175/2009MWR3113.1.

    • Search Google Scholar
    • Export Citation
  • Harlim, J., A. Mahdi, and A. J. Majda, 2014: An ensemble Kalman filter for statistical estimation of physics constrained nonlinear regression models. J. Comput. Phys., 257, 782812, doi:10.1016/j.jcp.2013.10.025.

    • Search Google Scholar
    • Export Citation
  • Hinze, J. O., and J. Hinze, 1959: Turbulence: An Introduction to Its Mechanism and Theory. Vol. 162, McGraw-Hill, 586 pp.

  • Hyndman, R. J., and A. B. Koehler, 2006: Another look at measures of forecast accuracy. Int. J. Forecasting, 22, 679688, doi:10.1016/j.ijforecast.2006.03.001.

    • Search Google Scholar
    • Export Citation
  • Iglesias, M. A., K. Lin, and A. M. Stuart, 2014: Well-posed Bayesian geometric inverse problems arising in subsurface flow. Inverse Prob., 30 (11), 114001, doi:10.1088/0266-5611/30/11/114001.

    • Search Google Scholar
    • Export Citation
  • Kalman, R. E., and R. S. Bucy, 1961: New results in linear filtering and prediction theory. J. Basic Eng., 83, 95108, doi:10.1115/1.3658902.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., 2003: Atmospheric Modeling, Data Assimilation, and Predictability. Cambridge University Press, 368 pp.

  • Keating, S. R., A. J. Majda, and K. S. Smith, 2012: New methods for estimating ocean eddy heat transport using satellite altimetry. Mon. Wea. Rev., 140, 17031722, doi:10.1175/MWR-D-11-00145.1.

    • Search Google Scholar
    • Export Citation
  • Kondrashov, D., S. Kravtsov, A. W. Robertson, and M. Ghil, 2005: A hierarchy of data-based ENSO models. J. Climate, 18, 44254444, doi:10.1175/JCLI3567.1.

    • Search Google Scholar
    • Export Citation
  • Kravtsov, S., D. Kondrashov, and M. Ghil, 2005: Multilevel regression modeling of nonlinear processes: Derivation and applications to climatic variability. J. Climate, 18, 44044424, doi:10.1175/JCLI3544.1.

    • Search Google Scholar
    • Export Citation
  • Liptser, R. S., and A. N. Shiryaev, 1977: Statistics of Random Processes I: General Theory. Springer. 402 pp.

  • Liptser, R. S., and A. N. Shiryaev, 2001: Statistics of Random Processes II: Applications. Springer, 402 pp.

  • Liu, J., and M. West, 2001: Combined parameter and state estimation in simulation-based filtering. Sequential Monte Carlo Methods in Practice, A. Doucet, N. de Freitas, and N. Gordon, Eds., Springer, 197–223.

  • Lorenz, E. N., 1963: Deterministic nonperiodic flow. J. Atmos. Sci., 20, 130141, doi:10.1175/1520-0469(1963)020<0130:DNF>2.0.CO;2.

  • MacKay, D. J., 2003: Information Theory, Inference and Learning Algorithms. Cambridge University Press, 640 pp.

  • Majda, A. J., 2003: Introduction to PDEs and Waves for the Atmosphere and Ocean. Courant Lecture Notes Series, Vol. 9, American Mathematical Society, 234 pp.

  • Majda, A. J., 2012: Challenges in climate science and contemporary applied mathematics. Commun. Pure Appl. Math., 65, 920948, doi:10.1002/cpa.21401.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., 2015: Statistical energy conservation principle for inhomogeneous turbulent dynamical systems. Proc. Natl. Acad. Sci. USA, 112, 89378941, doi:10.1073/pnas.1510465112.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and X. Wang, 2006: Nonlinear Dynamics and Statistical Theories for Basic Geophysical Flows. Cambridge University Press, 564 pp.

  • Majda, A. J., and B. Gershgorin, 2010: Quantifying uncertainty in climate change science through empirical information theory. Proc. Natl. Acad. Sci. USA, 107, 14 95814 963, doi:10.1073/pnas.1007009107.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and M. Branicki, 2012: Lessons in uncertainty quantification for turbulent dynamical systems. Discrete Cont. Dyn. Syst., 32, 31333221, doi:10.3934/dcds.2012.32.3133.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and J. Harlim, 2012: Filtering Complex Turbulent Systems. Cambridge University Press, 357 pp.

  • Majda, A. J., and Y. Yuan, 2012: Fundamental limitations of ad hoc linear and quadratic multi-level regression models for physical systems. Discrete Cont. Dyn. Syst., 17, 13331363, doi:10.3934/dcdsb.2012.17.1333.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and B. Gershgorin, 2013: Elementary models for turbulent diffusion with complex physical features: Eddy diffusivity, spectrum and intermittency. Philos. Trans. A Math. Phys. Eng. Sci., 371, 20120184, doi:10.1098/rsta.2012.0184.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and J. Harlim, 2013: Physics constrained nonlinear regression models for time series. Nonlinearity, 26, 201271, doi:10.1088/0951-7715/26/1/201.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and I. Grooms, 2014: New perspectives on superparameterization for geophysical turbulence. J. Comput. Phys., 271, 6077, doi:10.1016/j.jcp.2013.09.014.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., and Y. Lee, 2014: Conceptual dynamical models for turbulence. Proc. Natl. Acad. Sci. USA, 111, 65486553, doi:10.1073/pnas.1404914111.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden Eijnden, 1999: Models for stochastic climate prediction. Proc. Natl. Acad. Sci. USA, 96, 14 68714 691, doi:10.1073/pnas.96.26.14687.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden Eijnden, 2001: A mathematical framework for stochastic climate models. Commun. Pure Appl. Math., 54, 891974, doi:10.1002/cpa.1014.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden-Eijnden, 2002: A priori tests of a stochastic mode reduction strategy. Physica D, 170, 206252, doi:10.1016/S0167-2789(02)00578-X.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., I. Timofeyev, and E. Vanden-Eijnden, 2003: Systematic strategies for stochastic mode reduction in climate. J. Atmos. Sci., 60, 17051722, doi:10.1175/1520-0469(2003)060<1705:SSFSMR>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., R. V. Abramov, and M. J. Grote, 2005: Information Theory and Stochastics for Multiscale Nonlinear Systems. CRM Monogr. Series, Vol. 25, American Mathematical Society, 133 pp.

  • Majda, A. J., C. Franzke, and D. Crommelin, 2009: Normal forms for reduced stochastic climate models. Proc. Natl. Acad. Sci. USA, 106, 36493653, doi:10.1073/pnas.0900173106.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., R. Abramov, and B. Gershgorin, 2010a: High skill in low-frequency climate response through fluctuation dissipation theorems despite structural instability. Proc. Natl. Acad. Sci. USA, 107, 581586, doi:10.1073/pnas.0912997107.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., J. Harlim, and B. Gershgorin, 2010b: Mathematical strategies for filtering turbulent dynamical systems. Discrete Cont. Dyn. Syst., 27, 441486, doi:10.3934/dcds.2010.27.441.

    • Search Google Scholar
    • Export Citation
  • Majda, A. J., D. Qi, and T. P. Sapsis, 2014: Blended particle filters for large-dimensional chaotic dynamical systems. Proc. Natl. Acad. Sci. USA, 111, 75117516, doi:10.1073/pnas.1405675111.

    • Search Google Scholar
    • Export Citation
  • Marquardt, D. W., 1963: An algorithm for least-squares estimation of nonlinear parameters. J. Soc. Ind. Appl. Math., 11, 431441, doi:10.1137/0111030.

    • Search Google Scholar
    • Export Citation
  • Neelin, J. D., B. R. Lintner, B. Tian, Q. Li, L. Zhang, P. K. Patra, M. T. Chahine, and S. N. Stechmann, 2010: Long tails in deep columns of natural and anthropogenic tropospheric tracers. Geophys. Res. Lett., 37, L05804, doi:10.1029/2009GL041726.

    • Search Google Scholar
    • Export Citation
  • Palmer, T. N., 2001: A nonlinear dynamical perspective on model error: A proposal for non-local stochastic-dynamic parametrization in weather and climate prediction models. Quart. J. Roy. Meteor. Soc., 127, 279304, doi:10.1002/qj.49712757202.

    • Search Google Scholar
    • Export Citation
  • Plett, G. L., 2004: Extended Kalman filtering for battery management systems of LiPB-based HEV battery packs: Part 3. State and parameter estimation. J. Power Sources, 134, 277292, doi:10.1016/j.jpowsour.2004.02.033.

    • Search Google Scholar
    • Export Citation
  • Qi, D., and A. J. Majda, 2015: Blended particle methods with adaptive subspaces for filtering turbulent dynamical systems. Physica D, 298–299, 2141, doi:10.1016/j.physd.2015.02.002.

    • Search Google Scholar
    • Export Citation
  • Salamon, P., and L. Feyen, 2009: Assessing parameter, precipitation, and predictive uncertainty in a distributed hydrological model using sequential data assimilation with the particle filter. J. Hydrol., 376, 428442, doi:10.1016/j.jhydrol.2009.07.051.

    • Search Google Scholar
    • Export Citation
  • Salmon, R., 1998: Lectures on Geophysical Fluid Dynamics. Oxford University Press, 400 pp.

  • Sapsis, T. P., and A. J. Majda, 2013a: Blending modified Gaussian closure and non-Gaussian reduced subspace methods for turbulent dynamical systems. J. Nonlinear Sci., 23, 10391071, doi:10.1007/s00332-013-9178-1.

    • Search Google Scholar
    • Export Citation
  • Sapsis, T. P., and A. J. Majda, 2013b: Statistically accurate low-order models for uncertainty quantification in turbulent dynamical systems. Proc. Natl. Acad. Sci. USA, 110, 13 70513 710, doi:10.1073/pnas.1313065110.

    • Search Google Scholar
    • Export Citation
  • Slivinski, L., E. Spiller, A. Apte, and B. Sandstede, 2015: A hybrid particle–ensemble Kalman filter for Lagrangian data assimilation. Mon. Wea. Rev., 143, 195211, doi:10.1175/MWR-D-14-00051.1.

    • Search Google Scholar
    • Export Citation
  • Smedstad, O. M., and J. J. O’Brien, 1991: Variational data assimilation and parameter estimation in an equatorial Pacific Ocean model. Prog. Oceanogr., 26, 179241, doi:10.1016/0079-6611(91)90002-4.

    • Search Google Scholar
    • Export Citation
  • Snijders, T. A. B., 2011: Multilevel analysis. International Encyclopedia of Statistical Science, M. Lovric, Ed., Springer, 879–882, doi:10.1007/978-3-642-04898-2_387.

  • Sowell, F., 1992: Maximum likelihood estimation of stationary univariate fractionally integrated time series models. J. Econometrics, 53, 165188, doi:10.1016/0304-4076(92)90084-5.

    • Search Google Scholar
    • Export Citation
  • Tong, X. T., A. J. Majda, and D. Kelly, 2016: Nonlinear stability of the ensemble Kalman filter with adaptive covariance inflation. Commun. Math. Sci., 14, 12831313, doi:10.4310/CMS.2016.v14.n5.a5.

    • Search Google Scholar
    • Export Citation
  • Townsend, A. A., 1980: The Structure of Turbulent Shear Flow. Cambridge University Press, 429 pp.

  • Vallis, G. K., 2006: Atmospheric and Oceanic Fluid Dynamics: Fundamentals and Large-Scale Circulation. Cambridge University Press, 745 pp.

  • Van Der Merwe, R., and E. Wan, 2001: The square-root unscented Kalman filter for state and parameter-estimation. IEEE Int. Conf. on Acoustics, Speech, and Signal Processing, 2001 (ICASSP’01). Vol. 6, Salt Lake City, UT, Institute of Electrical and Electronics Engineers, 3461–3464, doi:10.1109/ICASSP.2001.940586.

  • Wenzel, T. A., K. Burnham, M. Blundell, and R. Williams, 2006: Dual extended Kalman filter for vehicle state and parameter estimation. Veh. Syst. Dyn., 44, 153171, doi:10.1080/00423110500385949.

    • Search Google Scholar
    • Export Citation
  • Wikle, C. K., and M. B. Hooten, 2010: A general science-based framework for dynamical spatio-temporal models. TEST, 19, 417451, doi:10.1007/s11749-010-0209-z.

    • Search Google Scholar
    • Export Citation
  • Yeh, W. W.-G., 1986: Review of parameter identification procedures in groundwater hydrology: The inverse problem. Water Resour. Res., 22, 95108, doi:10.1029/WR022i002p00095.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    Skill scores for filtering the unresolved process υ using the physics-constrained (perfect) filter in (8) and the stochastic parameterized (imperfect) filter in (9) as a function of and in generating the truth. The skill scores in (a),(b) dynamics regime A and (c),(d) dynamics regime B are shown. (top),(middle) The RMS error and pattern correlation in the filtered solution compared with the truth are shown. (bottom) The model error in (6) in the time-averaged PDF of the posterior mean estimation compared with that of the truth π. The parameters and in the stochastic parameterized filter in (9) is calibrated by matching the statistics with those of nature in (8).

  • Fig. 2.

    (Dynamical regime A: With practical observability at the attractor). Comparison of the posterior mean estimation of υ across time using the physics-constrained (perfect) filter (8) and the stochastic parameterized (imperfect) filter (9). The time-averaged PDFs of the filtered solutions compared with the truth are also illustrated. (a) The situation with small noise in the observational process and large noise in the filtering process. (b) The situation with large noise in the observational process and small noise in the filtering process. The parameters , and in the stochastic parameterized filter in (9) are calibrated by matching the statistics with those of nature in (8).

  • Fig. 3.

    (Dynamical regime B: Without practical observability at the attractor). As in Fig. 2. (a) The situation with small noise in the observational process and moderate noise in the filtering process. (b) The situation with small noise in the observational process and large noise in the filtering process.

  • Fig. 4.

    Trajectories of the noisy L-63 model (12). (a) , (b) , and (c) .

  • Fig. 5.

    Filtering the L-63 model (15) utilizing the imperfect noisy L-63 model (14). (a) Observing y and z and filtering x. The three noise levels , and in the imperfect forecast model (14) are set to be the same, and the RMS error and pattern correlation in the posterior mean estimation compared with the truth as a function of these noise levels are shown in the first two rows, where the statistics are computed across time and the dotted line shows the standard deviation of each variable at equilibrium. The third row shows the posterior variance as a function of the noise levels, where the statistics are averaged across time and the dotted line shows the equilibrium covariance of each variable. (b) Observing x and filtering y and z. (c) The observation in x across time and the filtering estimators in y and z with and nonzero noise in the imperfect forecast model (14). (d) As in (c), but with zero noise .

  • Fig. 6.

    Filtering the noisy L-63 model (12) utilizing the imperfect forecast L-63 model (17) with no noise in the filtering processes, where x is the observed variable and y and z are the variables for filtering. The noise in the observational process in (17) is set to be the same in (12). (a)–(c) The RMS error and pattern correlation in the posterior mean compared with the truth, the posterior covariance, and model error in (6) with different as a function of the noise and in the true model in (12) where . The dotted line in the first row shows the equilibrium standard deviation and that in the third row shows the equilibrium variance of each variable. The statistics are averaged across time . (d) The filtering estimation across time, where .

  • Fig. 7.

    Filtering the noisy L-63 model (12) utilizing an imperfect forecast model (14) with model error in the noise, where the observational variable is x and the variables for filtering are y and z. (a),(c),(e) Filtering skill as a function of the noise in the observational process. (b),(d),(f) Filtering skill as a function of the noise and , where . (a),(b) Small; (c),(d) moderate; and (e),(f) large noise , and 10 in the true system are shown. The dotted line in the first row shows the equilibrium standard deviation and that in the third row shows the equilibrium variance of each variable. The statistics are averaged across time .

  • Fig. 8.

    Comparison of the true signal (blue) and posterior mean estimation (red) in the underdispersion cases, where in the model that generates the true signal in (12). (a)–(c) Filtering skill with underdispersed noise in the observational process. (d)–(f) Filtering skill with underdispersed noise in the filtering processes.

  • Fig. 9.

    (a)–(d) Recovering of the mean, variance, skewness, and kurtosis associated with the marginal PDF associated with the unobserved variable z in the noisy L-63 model (12) utilizing the conditional Gaussian ensemble mixture approach (19) with different number of ensembles L in a perfect model setting. As comparison, the recovered statistics utilizing Monte Carlo simulation with N = 50 000 ensemble members are also included. The green dot in (c) indicates the largest skewness in the transition phase, which will be utilized in Fig. 10.

  • Fig. 10.

    The model error in (6) in recovering the marginal PDF associated with the unobserved variables y and z in the noisy L-63 model (12) utilizing the conditional Gaussian ensemble mixture approach in (19) with different L in a perfect model setting. (a) The model error in the PDF of y as a function of L at a short-term transition phase with largest skewness. (e) Comparing the PDFs of y at utilizing the conditional Gaussian mixture in (19) and Monte Carlo with N = 50 000. (c),(g) As in (a),(e), but at time , at which the system reaches statistical equilibrium state. (b),(d),(f),(h) The marginal distribution associated with the unobserved variable z.

  • Fig. 11.

    Recovery of the marginal PDFs associated with the unobserved variables y and z in the presence of model error from noise inflation. The noisy L-63 model (12) with is utilized in generating the true signal. The imperfect model (14) with is adopted for recovering the hidden PDFs. (a)–(d) The recovered mean, variance, skewness, and kurtosis compared with the truth that is computed by Monte Carlo simulation with N = 50 000 samples. (e) The recovered PDFs at short-term transition time with maximum skewness, where for y and for z. (f) The recovered PDFs at statistical equilibrium state for both y and z and (g) the PDFs at in logarithmic scale.

  • Fig. 12.

    Recovery of the marginal PDFs associated with the unobserved variables y and z in the presence of model error from underdispersion of noise. The noisy L-63 model (12) with is utilized in generating the true signal. The imperfect model (14) with is adopted for recovering the hidden PDFs. (a)–(d) The recovered mean, variance, skewness, and kurtosis compared with the truth that is computed by Monte Carlo simulation with N = 50 000 samples. (e) The recovered PDFs at short-term transition time with maximum skewness, where for y and for z. (f) The recovered PDFs at statistical equilibrium state for both y and z and (g) the PDFs at in logarithmic scale.

  • Fig. 13.

    Comparison of estimating the multiplicative parameter in (29). (a)–(d) Estimation skill utilizing the direct approach in (30). (e)–(h) Estimation skill utilizing the stochastic parameterized equation (36). Here, the truth and are adopted. In the stochastic parameterized equation (36), the equilibrium mean and equilibrium are fixed. The black dotted line represents the truth ; the red curve is the posterior mean or posterior variance ; the two green curves around the posterior mean differs from the mean by one standard deviation (i.e., ).

  • Fig. 14.

    Phase portrait of for the deterministic part of cubic model with and fixed. The three red dots are the three examples utilized in Fig. 15 to study the parameter estimation skill.

  • Fig. 15.

    Parameter estimation of the cubic model (39) in different dynamical regimes. (a) The observational trajectories with , and as shown in Fig. 14. The other parameters are , and . (b) The corresponding PDF. (c),(d) The estimated parameters and the associated estimation uncertainty of parameter c.

  • Fig. 16.

    Parameter estimation of the cubic nonlinear model (39) in regime III with with (a)–(d) the direct approach and (e)–(h) the stochastic parameterized equation method. The black dotted line shows the truth of each parameter and the red dashed line shows the averaged value of the estimation of each parameter utilizing stochastic parameterized equation method at equilibrium. Here the equilibrium mean of stochastic parameterized equation has 0.5 unit bias from the truth. The equilibrium variance of each stochastic parameterized equation is . The damping coefficient in the stochastic parameterized equation is set to be .

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 1400 679 67
PDF Downloads 623 136 8