Review of the Ensemble Kalman Filter for Atmospheric Data Assimilation

P. L. Houtekamer Meteorology Research Division, Environment and Climate Change Canada, Dorval, Québec, Canada

Search for other papers by P. L. Houtekamer in
Current site
Google Scholar
PubMed
Close
and
Fuqing Zhang Department of Meteorology, The Pennsylvania State University, University Park, Pennsylvania

Search for other papers by Fuqing Zhang in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

This paper reviews the development of the ensemble Kalman filter (EnKF) for atmospheric data assimilation. Particular attention is devoted to recent advances and current challenges. The distinguishing properties of three well-established variations of the EnKF algorithm are first discussed. Given the limited size of the ensemble and the unavoidable existence of errors whose origin is unknown (i.e., system error), various approaches to localizing the impact of observations and to accounting for these errors have been proposed. However, challenges remain; for example, with regard to localization of multiscale phenomena (both in time and space). For the EnKF in general, but higher-resolution applications in particular, it is desirable to use a short assimilation window. This motivates a focus on approaches for maintaining balance during the EnKF update. Also discussed are limited-area EnKF systems, in particular with regard to the assimilation of radar data and applications to tracking severe storms and tropical cyclones. It seems that relatively less attention has been paid to optimizing EnKF assimilation of satellite radiance observations, the growing volume of which has been instrumental in improving global weather predictions. There is also a tendency at various centers to investigate and implement hybrid systems that take advantage of both the ensemble and the variational data assimilation approaches; this poses additional challenges and it is not clear how it will evolve. It is concluded that, despite more than 10 years of operational experience, there are still many unresolved issues that could benefit from further research.

Contents

  • Introduction...4490

  • Popular flavors of the EnKF algorithm...4491

    1. General description...4491

    2. Stochastic and deterministic filters...4492

      1. The stochastic filter...4492

      2. The deterministic filter...4492

    3. Sequential or local filters...4493

      1. Sequential ensemble Kalman filters...4493

      2. The local ensemble transform Kalman filter...4494

    4. Extended state vector...4494

    5. Issues for the development of algorithms...4495

  • Use of small ensembles...4495

    1. Monte Carlo methods...4495

    2. Validation of reliability...4497

    3. Use of group filters with no inbreeding...4498

    4. Sampling error due to limited ensemble size: The rank problem...4498

    5. Covariance localization...4499

      1. Localization in the sequential filter...4499

      2. Localization in the LETKF...4499

      3. Issues with localization...4500

    6. Summary...4501

  • Methods to increase ensemble spread...4501

    1. Covariance inflation...4501

      1. Additive inflation...4501

      2. Multiplicative inflation...4502

      3. Relaxation to prior ensemble information...4502

      4. Issues with inflation...4503

    2. Diffusion and truncation...4503

    3. Error in physical parameterizations...4504

      1. Physical tendency perturbations...4504

      2. Multimodel, multiphysics, and multiparameter approaches...4505

      3. Future directions...4505

    4. Realism of error sources...4506

  • Balance and length of the assimilation window...4506

    1. The need for balancing methods...4506

    2. Time-filtering methods...4506

    3. Toward shorter assimilation windows...4507

    4. Reduction of sources of imbalance...4507

  • Regional data assimilation...4508

    1. Boundary conditions and consistency across multiple domains...4509

    2. Initialization of the starting ensemble...4510

    3. Preprocessing steps for radar observations...4510

    4. Use of radar observations for convective-scale analyses...4511

    5. Use of radar observations for tropical cyclone analyses...4511

    6. Other issues with respect to LAM data assimilation...4511

  • The assimilation of satellite observations...4512

    1. Covariance localization...4512

    2. Data density...4513

    3. Bias-correction procedures...4513

    4. Impact of covariance cycling...4514

    5. Assumptions regarding observational error...4514

    6. Recommendations regarding satellite observations...4515

  • Computational aspects...4515

    1. Parameters with an impact on quality...4515

    2. Overview of current parallel algorithms...4516

    3. Evolution of computer architecture...4516

    4. Practical issues...4517

    5. Approaching the gray zone...4518

    6. Summary...4518

  • Hybrids with variational and EnKF components...4519

    1. Hybrid background error covariances...4519

    2. E4DVar with the α control variable...4519

    3. Not using linearized models with 4DEnVar...4520

    4. The hybrid gain algorithm...4521

    5. Open issues and recommendations...4521

  • Summary and discussion...4521

    1. Stochastic or deterministic filters...4522

    2. The nature of system error...4522

    3. Going beyond the synoptic scales...4522

    4. Satellite observations...4523

    5. Hybrid systems...4523

    6. Future of the EnKF...4523

APPENDIX A...4524

Types of Filter Divergence...4524

  1. Classical filter divergence...4524

  2. Catastrophic filter divergence...4524

    APPENDIX B...4524

    Systems Available for Download...4524

    References...4525

Denotes Open Access content.

Corresponding author address: P. L. Houtekamer, Section de la Recherche en Assimilation des Données et en Météorologie Satellitaire, 2121 Route Trans-Canadienne, Dorval, QC H9P 1J3, Canada. E-mail: peter.houtekamer@canada.ca

Abstract

This paper reviews the development of the ensemble Kalman filter (EnKF) for atmospheric data assimilation. Particular attention is devoted to recent advances and current challenges. The distinguishing properties of three well-established variations of the EnKF algorithm are first discussed. Given the limited size of the ensemble and the unavoidable existence of errors whose origin is unknown (i.e., system error), various approaches to localizing the impact of observations and to accounting for these errors have been proposed. However, challenges remain; for example, with regard to localization of multiscale phenomena (both in time and space). For the EnKF in general, but higher-resolution applications in particular, it is desirable to use a short assimilation window. This motivates a focus on approaches for maintaining balance during the EnKF update. Also discussed are limited-area EnKF systems, in particular with regard to the assimilation of radar data and applications to tracking severe storms and tropical cyclones. It seems that relatively less attention has been paid to optimizing EnKF assimilation of satellite radiance observations, the growing volume of which has been instrumental in improving global weather predictions. There is also a tendency at various centers to investigate and implement hybrid systems that take advantage of both the ensemble and the variational data assimilation approaches; this poses additional challenges and it is not clear how it will evolve. It is concluded that, despite more than 10 years of operational experience, there are still many unresolved issues that could benefit from further research.

Contents

  • Introduction...4490

  • Popular flavors of the EnKF algorithm...4491

    1. General description...4491

    2. Stochastic and deterministic filters...4492

      1. The stochastic filter...4492

      2. The deterministic filter...4492

    3. Sequential or local filters...4493

      1. Sequential ensemble Kalman filters...4493

      2. The local ensemble transform Kalman filter...4494

    4. Extended state vector...4494

    5. Issues for the development of algorithms...4495

  • Use of small ensembles...4495

    1. Monte Carlo methods...4495

    2. Validation of reliability...4497

    3. Use of group filters with no inbreeding...4498

    4. Sampling error due to limited ensemble size: The rank problem...4498

    5. Covariance localization...4499

      1. Localization in the sequential filter...4499

      2. Localization in the LETKF...4499

      3. Issues with localization...4500

    6. Summary...4501

  • Methods to increase ensemble spread...4501

    1. Covariance inflation...4501

      1. Additive inflation...4501

      2. Multiplicative inflation...4502

      3. Relaxation to prior ensemble information...4502

      4. Issues with inflation...4503

    2. Diffusion and truncation...4503

    3. Error in physical parameterizations...4504

      1. Physical tendency perturbations...4504

      2. Multimodel, multiphysics, and multiparameter approaches...4505

      3. Future directions...4505

    4. Realism of error sources...4506

  • Balance and length of the assimilation window...4506

    1. The need for balancing methods...4506

    2. Time-filtering methods...4506

    3. Toward shorter assimilation windows...4507

    4. Reduction of sources of imbalance...4507

  • Regional data assimilation...4508

    1. Boundary conditions and consistency across multiple domains...4509

    2. Initialization of the starting ensemble...4510

    3. Preprocessing steps for radar observations...4510

    4. Use of radar observations for convective-scale analyses...4511

    5. Use of radar observations for tropical cyclone analyses...4511

    6. Other issues with respect to LAM data assimilation...4511

  • The assimilation of satellite observations...4512

    1. Covariance localization...4512

    2. Data density...4513

    3. Bias-correction procedures...4513

    4. Impact of covariance cycling...4514

    5. Assumptions regarding observational error...4514

    6. Recommendations regarding satellite observations...4515

  • Computational aspects...4515

    1. Parameters with an impact on quality...4515

    2. Overview of current parallel algorithms...4516

    3. Evolution of computer architecture...4516

    4. Practical issues...4517

    5. Approaching the gray zone...4518

    6. Summary...4518

  • Hybrids with variational and EnKF components...4519

    1. Hybrid background error covariances...4519

    2. E4DVar with the α control variable...4519

    3. Not using linearized models with 4DEnVar...4520

    4. The hybrid gain algorithm...4521

    5. Open issues and recommendations...4521

  • Summary and discussion...4521

    1. Stochastic or deterministic filters...4522

    2. The nature of system error...4522

    3. Going beyond the synoptic scales...4522

    4. Satellite observations...4523

    5. Hybrid systems...4523

    6. Future of the EnKF...4523

APPENDIX A...4524

Types of Filter Divergence...4524

  1. Classical filter divergence...4524

  2. Catastrophic filter divergence...4524

    APPENDIX B...4524

    Systems Available for Download...4524

    References...4525

Denotes Open Access content.

Corresponding author address: P. L. Houtekamer, Section de la Recherche en Assimilation des Données et en Météorologie Satellitaire, 2121 Route Trans-Canadienne, Dorval, QC H9P 1J3, Canada. E-mail: peter.houtekamer@canada.ca

1. Introduction

The ensemble Kalman filter (EnKF; Evensen 1994) originated from the merger of Kalman filter theory and Monte Carlo estimation methods. The Kalman–Bucy filter (Kalman 1960; Kalman and Bucy 1961) provides the mathematical framework for the four-dimensional (4D) assimilation of observations into a state vector. Precursor references in the meteorological literature appear in the late 1960s (Jones 1965; Petersen 1968). The use of the Kalman filter in meteorology was further investigated in the 1980s and early 1990s (e.g., Ghil et al. 1981; Cohn and Parrish 1991; Daley 1995). One unsolved problem, aimed at an application with a realistic high-dimensional atmospheric forecast model, was how to obtain an appropriate low-dimensional approximation of the background error covariance matrix for a feasible implementation on a computational platform. The use of random ensembles currently seems to be the most practical way to address the issue.

The use of Monte Carlo experiments and ensembles also has long roots in NWP, in particular in the fields of ensemble forecasting (Lorenz 1965; Leith 1974; Kalnay and Dalcher 1987) and observing system simulation experiments (OSSEs) (Newton 1954; Daley and Mayer 1986). The OSSEs form a special category where the ensemble is composed of only two members. Here, one NWP center uses its model to provide a long integration, called a “nature run,” which serves as a proxy for the true atmospheric state. This center will also apply the forward operators of its data assimilation system to generate simulated observations from the nature run. An independent NWP center will subsequently use its own data assimilation system to assimilate these simulated observations with its own forward operators and forecast model. Because of the collaboration of two independent NWP centers, the errors, such as the error due to having an imperfect forecast model, are being sampled in a realistic manner. Unfortunately, with only two participating centers, only a single realization of the error is obtained and spatial or temporal averaging will be required to estimate characteristics of the error (e.g., Errico et al. 2013). The Monte Carlo method provides a general framework for the sampling of errors that can be due to a large variety of sources (Houtekamer et al. 1996). Bonavita et al. (2012) describe the use of an ensemble of data assimilations (EDA) to provide estimates of background error variances to the operational 4D-Var system at ECMWF.

The EnKF uses Monte Carlo methods to estimate the error covariances of the background error. In combination with covariance localization (Hamill et al. 2001), it provides an approximation to the Kalman–Bucy filter that is feasible for operational atmospheric data assimilation problems (Houtekamer et al. 2005, 2014a); it also provides an ensemble of initial conditions that can be used in an ensemble prediction system.

In the first review of the EnKF, Evensen (2003) gives a comprehensive description of the then already rapidly developing field with references to many applications in the earth sciences. An excellent review by Hamill (2006) relates the EnKF to Bayesian methods, to the Kalman filter, and to the extended Kalman filter. It also discusses different properties of stochastic and deterministic update algorithms, and stresses the need for model error parameterization and covariance localization. Hamill (2006) mostly speculates on prospects for having operational applications in the future, but a lot of work has been accomplished since. At the Canadian Meteorological Centre (CMC), a global EnKF has been used operationally since January 2005 to provide the initial conditions for a global ensemble prediction system (Houtekamer et al. 2005). In this system, observation preprocessing is done by a higher-resolution variational analysis system. At the time of this writing, the global EnKF is also used to provide initial and lateral boundary conditions to a regional ensemble (Lavaysse et al. 2013) and it provides the background error covariances for global (Buehner et al. 2015) and regional (Caron et al. 2015) ensemble-variational analysis systems. At NCEP, a “hybrid�? global EnKF is used operationally, in combination with a variational solver, for both the deterministic high-resolution global (X. Wang et al. 2013) and the regional (Pan et al. 2014) analyses. A regional EnKF system is in operational use at the Italian National Meteorological Center (CNMCA) (Bonavita et al. 2010) and high-resolution convection-permitting EnKF systems are used experimentally (e.g., Zhang et al. 2011; Aksoy et al. 2013; Xue et al. 2013; Putnam et al. 2014; Schwartz et al. 2015; Zhang and Weng 2015).

In the current review, the focus is on issues directly related to improving the quality of operational, quasi-operational, and experimental EnKF systems in atmospheric applications. Many of these issues, such as how to best account for model error, have themselves been the subject of workshops and review papers. Here, we will try to highlight the relationships between different aspects of the EnKF algorithm. One may choose, for instance, a certain algorithm that parallelizes well. This, in turn, may impose a certain choice for the covariance localization method, which ultimately may facilitate (or not) the assimilation of satellite or radar observations. We hope the reader will bear with us and arrive at a better understanding of the complex issues associated with the development of a high-quality EnKF system.

In section 2, the basic EnKF algorithm, as well as some of the most popular variations, are presented briefly. Section 3 describes how cross validation and covariance localization allow a small ensemble to be used in an EnKF that behaves well for high-dimensional systems. In truly realistic environments, there are sources of error that are not fully understood. Methods to account for such error sources are presented in section 4. One of the side effects of localization is imbalance; methods to control imbalance are the subject of section 5. This issue is important because well-balanced initial conditions allow for more frequent analyses in high-resolution models. Issues specific to high-resolution EnKF systems, such as how to best use observations for modern radar systems, are reviewed in section 6. An open question, of particular importance for operational centers, is whether EnKF systems can use satellite observations as effectively as variational systems. Various issues that could play a role are discussed in section 7. EnKF systems will reputedly scale well on modern and future computer systems with O(10 000)–O(100 000) cores; the current status with respect to computational issues is described in section 8. At operational centers, as a result of a variety of scientific and practical considerations, there is a lot of interest in the combination of variational and EnKF systems in a manner that leads to the highest quality forecasts. This is discussed in section 9. Finally, in section 10, we summarize what has been achieved and discuss where we think progress can be made.

2. Popular flavors of the EnKF algorithm

After the introduction of the EnKF (Evensen 1994), many variations on the original algorithm have been developed. Depending on the particular application, different aspects of the system may be judged more or less important. In this section, we try to give an overview of some of the main EnKF families that now coexist. In section 2a, we essentially follow Houtekamer and Mitchell (2005, their sections 2a,b) with a general description of the ensemble approximation to the Kalman filter. In section 2b, we discuss the dichotomy between stochastic and deterministic filters. In section 2c, we discuss how sequential and local filters can be used toward a computationally feasible algorithm. The commonly used extended state vector technique is referred to in section 2d. Finally, in section 2e, we discuss why it is too early to favor one EnKF algorithm over another.

a. General description

Any EnKF implementation updates a prior estimate of the atmosphere valid at some time t with the information in new observations to arrive at an updated estimate of the atmosphere as in Eq. (1). To this end, a Kalman gain matrix can be used to give an appropriate weight to the observations, which have error covariance , and the background, which has error covariance , as in Eq. (2). The forward operator performs the mapping from model space to observation space. Finally, a forecast model is needed to transport the new estimate to the next analysis time as in Eq. (3):
e1
e2
e3
In a pure Monte Carlo implementation, the ith member of an -member analysis ensemble is obtained by evaluating Eq. (1) using a randomly perturbed vector of observations and using a member of a corresponding ensemble of background estimates:
e4
Similarly, to obtain a member of the background ensemble valid at time , Eq. (3) can be used with a corresponding member of the analysis ensemble and a realization of the forecast model :
e5

The ensembles, generated by evaluating Eqs. (4) and (5), can be used to approximate the analysis error covariance matrix and the background error covariance matrix .

In an EnKF, one never needs a full covariance matrix like in model state space. Instead, for instance to compute the Kalman gain in Eq. (2), one uses ensemble-based approximations of and [Houtekamer and Mitchell (2001), their Eqs. (2) and (3)]:
e6
e7
where
eq1
Here a possibly nonlinear forward operator is used on the right-hand side of Eqs. (6) and (7). We thus obtain a nonlinear ensemble-based approximation of the terms and that appear in the standard Kalman filter equations [Ghil and Malanotte-Rizzoli (1991), their Eq. (4.17c)]. Similarly, the ensemble of nonlinear equations, Eqs. (4) and (5), replaces the corresponding matrix equations of the standard Kalman filter [Ghil and Malanotte-Rizzoli (1991), their Eqs. (4.13a) and (4.13b)]:
e8
e9
Here is the tangent linear approximation to the forecast model and the matrix contains the covariances of the forecast-model error. In the special case that the optimal Kalman gain is used, Eq. (9) can be rewritten as follows [Ghil and Malanotte-Rizzoli (1991), their Eq. (4.17d)]:
e10

b. Stochastic and deterministic filters

1) The stochastic filter

After the first implementation of an EnKF (Evensen 1994), it was realized that, to arrive at a consistent analysis scheme, the observations also needed to be treated as random variables (Burgers et al. 1998; Houtekamer and Mitchell 1998).

Lacking other information, it is commonly assumed that observation errors have a Gaussian distribution. The addition of random Gaussian noise at each analysis time, via Eq. (4), tends to erase the non-Gaussian higher moments nonlinear error growth may have generated (Lawson and Hansen 2004, their section 5). This can be understood as a manifestation of the Central Limit Theorem (Fishman 1996, his section 1.1.1). Since higher moments are not explicitly considered in an EnKF, maintaining Gaussianity likely has a positive impact on analysis quality.

A stochastic EnKF has been in operational use at CMC since January 2005 (Houtekamer and Mitchell 2005). As of November 2014, it uses 256 ensemble members (Table 1).

Table 1.

Information from global EnKF configurations at operational centers. Information provided in February 2015 by M. Bonavita for ECMWF and J. Whitaker for NCEP (M. Bonavita and J. Whitaker 2015, personal communications). The terms , , , and are the number of observations that serve to compute analysis increments, the number of ensemble members, the number of model coordinates in the control vector of the analysis, and the number of computer cores available to the analysis algorithm, respectively.

Table 1.

2) The deterministic filter

After the introduction of the stochastic EnKF, it was realized that the small, but spurious, correlations between the ensembles of backgrounds and observations could lead to a degradation of analysis quality. In a short period of time, the ensemble square root filter (EnSRF; Whitaker and Hamill 2002), the ensemble adjustment Kalman filter (EAKF; Anderson 2001), and the ensemble transform Kalman filter (ETKF; Bishop et al. 2001) were proposed. In these filters, which are rightly called deterministic, observations are not perturbed randomly. Instead, it is assumed that the optimal gain is available such that the analysis-error covariance can be obtained from Eq. (10). Without covariance localization, the ETKF, the EAKF, and the EnSRF lead to analysis ensembles that can be different, but that span the same subspace and have the same covariance (Tippett et al. 2003).

In a first step of an EnSRF, the ensemble mean analysis is obtained using the regular gain matrix [Eq. (2)], the ensemble mean trial field , and unperturbed observations . Subsequently, a modified gain matrix is used to obtain the ensemble of differences between the ensemble of analyses and the ensemble mean analysis (Potter 1964; Whitaker and Hamill 2002):
e11
e12
This modification is such that the observational uncertainty is properly accounted for in the resulting analysis-error covariances. Note that the observations are assimilated one at a time [also see section 2c(1)], and consequently α, , and are scalars in Eqs. (11) and (12).

Deterministic filters use Eq. (10) to relate background and analysis error covariances [Tippett et al. 2003, their Eq. (2)]. It follows that the computed analysis error covariances are smaller than the computed background error covariances . Unfortunately, when small ensembles are used, the Kalman gain will be affected by sampling error and the true analysis error may well exceed the background error. Since cross validation (section 3c) is not typically used in the deterministic filter, the ensemble may thus become underdispersive. To compensate for this particular problem, it is common to use a covariance inflation or relaxation procedure [section 4a(3)]. Because it avoids the impact of spurious correlations, the deterministic filter can obtain a similar quality as a corresponding stochastic filter with a smaller total ensemble size (Mitchell and Houtekamer 2009, lhs of their Fig. 4). It has also been shown (Lawson and Hansen 2004; Mitchell and Houtekamer 2009, their section 7b; Evensen 2009, their Fig. 7) that, without the regular introduction of random forcings, the deterministic filter can develop highly non-Gaussian distributions. In deterministic filters, the regular addition of model error fields, via Eq. (22) [section 4a(1)], can serve as a forcing toward Gaussian distributions (Lawson and Hansen 2004, their section 5).

Whereas, as mentioned, the EAKF and EnSRF algorithms are very similar, actual implementations for complex assimilation systems can be quite different. A deterministic EnSRF has been operational at NCEP using members since May 2012 (Whitaker et al. 2008) to provide flow-dependent covariances to a high-resolution variational analysis (X. Wang et al. 2013). The EAKF is conveniently available from the Data Assimilation Research Testbed (DART; Anderson et al. 2009) and is perhaps the most widely used EnKF algorithm. For oceanographic applications, there is a flavor of the deterministic filter in which the observation error covariances in the denominator of the gain matrix are obtained from an ensemble (Evensen 2004). This algorithm can be used to handle correlated observation errors.

c. Sequential or local filters

Two different methods have been proposed to reduce the numerical cost associated with the matrix inversion in Eq. (2). One can either use a sequential algorithm, in which observations are assimilated in a sequence of small batches, or one can split the spatial domain into a number of local areas where the analysis is solved independently as in the local ensemble transform Kalman filter (LETKF).

1) Sequential ensemble Kalman filters

If the observations have independent errors, they can be assimilated one at a time (e.g., Cohn and Parrish 1991; Anderson 2001) (i.e., serially) or in batches (Houtekamer and Mitchell 2001). In the context of the Kalman filter, both of these assimilation procedures lead to the same result as assimilating all observations simultaneously. Beyond the conceptual simplification, the computational cost of the inversion of is avoided or rendered insignificant, and it is possible to use different parameters at different stages of the algorithm. As will be discussed later, the latter feature can be exploited to have different localization length scales for different sets of observations (e.g., Zhang et al. 2009a; Snook et al. 2015).

In the sequential algorithm, a different order of the observations, which can result from trivial changes in observation preprocessing, can lead to moderate amplitude changes in the resulting analysis (e.g., Houtekamer and Mitchell 2001). This can make it difficult to interpret results of sensitivity experiments. In some cases, the combination of sequential processing and localization can make the analysis process unstable (Nerger 2015).

2) The local ensemble transform Kalman filter

There are various ways to derive and write the Kalman filter equations (Snyder 2015). Instead of Eq. (10), it is possible to write the following [Ghil and Malanotte-Rizzoli (1991), their Eq. (4.16a)]:
e13
and for the Kalman gain [Eq. (2)], using a linear operator , we have the following [Ghil and Malanotte-Rizzoli (1991), their Eq. (4.16b)]:
e14
The equivalence of the two sets of equations can be shown using the Sherman–Morrison–Woodbury identity (Golub and Van Loan 1996).
Continuing from the rhs of Eq. (14), the equations for the LETKF [Hunt et al. 2007, their Eqs. (21), (23), and (24)] can be written as
e15
e16
e17
where consists of local background perturbations interpolated to observations. The column , corresponding to ensemble member i, is formed as the difference of and its ensemble mean value . Similarly matrices are the differences between and their ensemble mean values. The matrix , for the analysis error covariance in ensemble space, has dimension . The LETKF first solves for the ensemble mean using Eq. (15), and subsequently the ensemble of background perturbations is transformed into an ensemble of analysis perturbations using the weights given by Eq. (17) (hence, the “ensemble transform�? in the name LETKF). Because each weight corresponds with a model trajectory valid in the data assimilation window, the time of validity of observations can be effectively accounted for at low computational cost (Hunt et al. 2004). The weights can also be interpolated in space (Yang et al. 2009), leading to additional computational efficiency. Finally, the weights could be reused when additional variables are added in the model state vector. Note that, since no randomly perturbed observations are used in Eqs. (15)(17), the LETKF is a deterministic filter. Because of the use of Eq. (17), it is also a square root filter.

The LETKF evolved from the earlier local ensemble Kalman filter (Ott et al. 2004). It has been developed with a prime focus on computational efficiency on massively parallel computers (Hunt et al. 2007; Szunyogh et al. 2008). Good scalability is achieved by the decomposition of the global analysis domain into a number of independent domains. In each such domain all nearby observations are used, and the analysis equations are solved in the space spanned by the ensemble perturbations. For large ensemble sizes, the main computational cost for computing the analysis is in finding the eigensolution of local matrices (Miyoshi et al. 2014).

A regional configuration of the LETKF is used operationally by the CNMCA (Bonavita et al. 2010) and a configuration of the KENDA-LETKF (Schraff et al. 2016) is used operationally by MeteoSwiss since May 2016. An LETKF has been developed for research purposes at the Japan Meteorological Agency (JMA; Miyoshi et al. 2010) as well as at the ECMWF (Hamrud et al. 2015). At the time of this writing, the Deutscher Wetterdienst (DWD) (Schraff et al. 2016) and Argentina (Dillion et al. 2016) were testing a regional LETKF for an operational implementation.

d. Extended state vector

A very powerful and useful technique is the use of joint state-observation space (Tarantola 1987; Anderson 2003). Here the joint space state vector is defined and computed as
e18
and subsequently the analysis equations are solved for . With this change, the values of will be updated by the analysis algorithm with no new evaluations of the nonlinear operator . For instance, if the observed quantity is the amount of precipitation, the forward operator can consist of the parameterization for deep convection as applied during a model integration—with no need to also implement this operator in the analysis code. The technique is also used to arrive at an efficient implementation of time interpolation (Houtekamer and Mitchell 2005, their section 4e). The same technique can also be applied to extend the model state with additional parameters as in the case of bias estimation for satellite observations (section 7c) and the estimation of surface fluxes (Kang et al. 2012).

e. Issues for the development of algorithms

There have been various intercomparisons of the EnSRF and LETKF deterministic algorithms (e.g., Greybush et al. 2011; Hamrud et al. 2015; Thompson et al. 2015). It has generally been concluded that the two methods provide comparable quality. At operational centers there is evidence of high (i.e., operational or near operational) quality results for the stochastic EnKF at CMC, for the deterministic EnSRF at NCEP, and for the LETKF at CNMCA, at DWD, and at ECMWF and JMA. The choice of algorithm tends to depend on lower-level algorithmic choices as summarized in Table 2.

Table 2.

Algorithmic choices made in popular EnKF algorithms. The algorithms are discussed in section 2. Cross validation is discussed in section 3c. The - and -localization methods are discussed in section 3e. The Stochastic EnKF, EnSRF, and EAKF use sequential algorithms with localization. The EnSRF, EAKF, and LETKF use a deterministic algorithm without cross validation.

Table 2.

For instance, when a stochastic algorithm is selected, cross validation can be used to obtain a reliable (i.e., sufficiently large) ensemble spread. With a deterministic algorithm, however, an additional algorithm will need to be used to maintain a sufficient spread in the analysis ensemble in areas where observations have been assimilated. For this, it is common to use relaxation methods that relax the analysis spread back to the original (bigger) spread in the background ensemble [section 4a(3)]. An advantage of deterministic algorithms is that, since they avoid perturbing the observations (a source of noise) in the update equations, they are likely to provide more accurate posterior mean estimates when the ensemble size is small. Similarly, the selection of a particular sequential or local algorithm, to make the computation of the analysis increments feasible, will have implications for the covariance localization algorithm (section 3e).

Because of the large uncertainty associated with all these important factors, it is not clear which of the three algorithms—if any—will eventually turn out to be the best choice for atmospheric data assimilation.

3. Use of small ensembles

In realistic atmospheric applications, the affordable ensemble size is limited by the cost of integrating the forecast model [Eq. (3)]. In such systems (Table 1), one often finds to be O(100), which is much smaller than the number of model variables in the control vector of the analysis , which is O(). Unfortunately, the use of such a relatively small ensemble has a profound impact on properties of the algorithm and special algorithmic measures are necessary to obtain good filter behavior.

In section 3a, we give an overview of the use of Monte Carlo methods in data assimilation. In section 3b, the difficulty of validating the reliability of such systems is discussed. Cross validation is presented in section 3c as a method to arrive at reliable ensembles. The sampling error due to having a small ensemble is discussed in section 3d. It can be alleviated using some flavor of covariance localization (section 3e). A summary is provided in section 3f.

a. Monte Carlo methods

An example of a Monte Carlo application is shown in Fig. 1. Here, we have a best estimate of the input and output state and we want to know how the uncertainty in the input state translates into uncertainty in an output state. In NWP, we have a numerical tool, like a forecast model, to generate an output from an input. By running the tool once for each input, we can obtain an ensemble of outputs. The only new utility required for running a Monte Carlo experiment is the tool to generate an ensemble of inputs. Note, in addition, that Monte Carlo experiments can be performed in sequence, with the output ensemble of one experiment serving as input ensemble for the next experiment. In this case, instead of specifically following a best estimate and its uncertainty, one can decide to always estimate a required probability distribution from the available ensemble of estimates. When a best estimate is also required, the ensemble mean can be used for this purpose and the ensemble spread should provide a reliable estimate of the error in the mean. The basic assumption justifying the use of the Monte Carlo method is that the uncertainty with respect to our best estimate evolves in almost the same way as the uncertainty with respect to the true state (Press et al. 1992, their section 15.6). This is similar to the assumption underlying the extended Kalman filter that a linearized model can provide a useful description of the evolution of small perturbations with respect to a reference trajectory (e.g., Ghil et al. 1981). The use of the nonlinear forecast model [Eq. (5)] in the Monte Carlo procedure does, however, permit a natural saturation of error for unstable, rapidly growing, modes. The EnKF generally features robust behavior when used for nonlinear problems (Verlaan and Heemink 2001).

Fig. 1.
Fig. 1.

The general recipe for the application of the Monte Carlo procedure. The best estimates are plotted with the large open circles. The black box is a procedure to transform input into output. To obtain information on the uncertainty in the output, the procedure is executed once for each member in the input ensemble. The individual estimates in each ensemble are indicated by the small filled circles.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0440.1

The increasing popularity of Monte Carlo methods is partly due to the simplicity of the design and partly due to the availability of powerful computers that permit executing tasks many times. For the solution of high-dimensional problems, Monte Carlo methods are also often more efficient than other methods (e.g., Fishman 1996, his chapter 1; Houtekamer 1993, his section 4). In meteorology, the use of Monte Carlo methods was suggested originally in the context of predictability research and ensemble forecasting (Lorenz 1965; Leith 1974). Here, it is necessary to sample the uncertainty in the initial conditions a number of times to arrive at an ensemble of forecasts [the Ensemble Prediction System (EPS) item in Table 3]. This is a perfect implementation of the Monte Carlo method to the extent that the forecast model is indeed independent of the ensemble of input perturbations. Since the model can be integrated with no knowledge or analysis of its inner logic, it can be considered a black box. Going toward an operational implementation of such an ensemble prediction system, the main difficulty is how to obtain a realistic multivariate covariance matrix from which to sample the initial conditions.

Table 3.

Some examples of possible applications of the Monte Carlo method. As input, only quantities of which random realizations are needed are listed. The “first initial state�? is an initial state valid at the time when the experiment started and “all observations�? refers to all observations that have been received since that time. The blackness is perfect when the system can be considered a black box, with all ensemble members being independent (such that changing the input for one member would have no impact on the output for the other members).

Table 3.

In EDA systems (Houtekamer et al. 1996, their Fig. 1), the system, that generates output from an input, is expanded to include an assimilation component (the EDA item in Table 3). At the minimum, the uncertain input of a data assimilation cycle consists of the observed values that arrive for each new assimilation window. Another input, that theoretically should also be considered, is the background field that served for the very first assimilation window. In practice, it can be assumed that the details of how the very first set of initial perturbations were obtained gradually cease to be important. The random perturbations that are being added continuously to the incoming observations, and, possibly also, to the forecast model, are what determines the spread in the output ensemble. In an experiment with a global EnKF that included model error sampling, Houtekamer et al. (2005, their Fig. 3) observe stable error levels after an initial spinup of about 4–7 days. In experiments without model error (Houtekamer et al. 2005, their Fig. 4), it appears to take longer to achieve convergence. In a regional EnKF, convergence can likely be obtained earlier (Cohn and Parrish 1991, their Fig. 5). An EDA system, with the assimilation of sets of randomly perturbed observations into an ensemble of background fields using an optimal interpolation methodology, served for many years in the operational medium-range EPS of CMC (Houtekamer et al. 1996, their section 4a). At ECMWF, the EDA method is used to estimate flow-dependent background error variances for a 4D-Var system (Bonavita et al. 2012). An EDA system is a clean implementation of the Monte Carlo method as long as the covariances of the EDA ensemble of background fields are not used to optimize the analysis system that serves in the EDA. Therefore, it can be expected that the ensemble spread will be representative of the error in the ensemble mean.

In stochastic and deterministic EnKF systems, we go one step further and estimate the Kalman gain matrix using the ensemble of background fields [Eqs. (2), (6), and (7)]. Such EnKF systems deviate from the Monte Carlo methodology, which requires that an existing assimilation system be tested with an independent test ensemble. One should not use the same ensemble both to obtain the gain matrix, which defines the data assimilation system, and subsequently to estimate the quality of that matrix from the spread in the resulting ensemble of analyses. As the gain matrix has been optimized for the specific directions that are present in the test ensemble, the gain matrix is going to be particularly effective for these directions and, consequently, the spread in the test ensemble of analyses is going to be relatively small. Unfortunately, the true error in the ensemble mean background and observations is not known, and will not be exactly in the space spanned by the simulated perturbations. A typical consequence is a lack of reliability, also known as inbreeding, in which the ensemble spread becomes an underestimate of the error of the ensemble mean (Houtekamer and Mitchell 1998, their Fig. 3).

b. Validation of reliability

In the EnKF, ensembles are used to estimate flow-dependent background error covariances. It would be desirable to verify if the generated ensembles do provide a reliable estimate of the ensemble mean error. It is, however, not evident how to verify the reliability of high-dimensional ensemble-based covariances. A simple verification is to check that the sum of the spread in the background ensemble and the observation-error variance match with the variance of the innovations [e.g., Houtekamer et al. (2005), their Eq. (4) and Fig. 5; Desroziers et al. (2005), their Eq. (1)]. Ensemble reliability for scalar quantities can also be verified using, notably, rank histograms and the reliability/resolution decomposition of the continuous ranked probability score (Hersbach 2000; Hamill 2001; Candille and Talagrand 2005, 2008). A difficulty in verifying the reliability of data assimilation systems is that unknown properties of the observational error will have an impact on verification results. The interpretation of results will also become problematic if innovation amplitudes, which supposedly provide independent information for the verification, are also being used in the tuning of additive and multiplicative inflation procedures [e.g., Houtekamer et al. (2005), their Eq. (3)].

Controlled experiments in which only the observations are considered to be a source of error are known as “perfect-model�? experiments. Here, one will normally generate random observation error using the observation-error covariance matrix . For example, Houtekamer et al. (2009, their Fig. 3) were able to obtain a good match between ensemble spread and error, thus validating the Monte Carlo technique in a perfect-model experiment with a stochastic EnKF. The simulated error levels in this experiment were, however, estimated to be only about half the size of actual error levels in operational systems. Similarly, using an EDA system, Bonavita et al. (2012) also found that globally averaged EDA spread values were underestimated by approximately a factor of 2. In other words, in data assimilation systems the origin of approximately half of the actual error amplitude is not simply a consequence of observational errors of known amplitude. Taking squares, it follows that only a quarter of the error variance has a known origin. Efforts to account for additional error sources will be reviewed in section 4.

A fairly high natural error level in perfect-model experiments would suggest that Eqs. (1) and (3) provide a fairly reasonable set of equations for the evolution of errors in the system, and could thus be seen as a justification for the use of a costly 4D assimilation system, such as 4D-Var and the EnKF. Given the modest fraction of error sources of known and quantifiable origin, it is perhaps surprising that operationally interesting results can already be obtained with the current generation of 4D assimilation systems.

c. Use of group filters with no inbreeding

As mentioned in section 3a, the use of a Monte Carlo ensemble to estimate the Kalman gain can lead to inbreeding in which the ensemble spread becomes a systematic underestimate of the error of the ensemble mean. In the context of a stochastic EnKF, it is possible to counter this problem using k-fold cross validation (Houtekamer and Mitchell 1998; Mitchell and Houtekamer 2009). Here, the original ensemble is divided into k subensembles of equal size. To assimilate observations into one subensemble of background fields one uses a gain matrix computed from other, independent, members. In the limiting case of leave-one-out cross validation, the subensembles consist of only one member (Hamill and Snyder 2000). As a result of cross validation, it is possible, at least in perfect-model environments, to maintain a representative ensemble in which the ensemble spread is an unbiased estimate of the ensemble mean error (e.g., Houtekamer et al. 2009, their Fig. 3). Whitaker and Hamill (2002, their appendix) give an example of twofold cross validation for a deterministic filter. Although comparatively good results were obtained, the algorithm had a very substantial computational cost and has not been pursued further.

When cross validation is used to obtain a reliable ensemble, not all members are used in the computation of the gain for a subensemble. Alternatively, one can use a single ensemble configuration in which all available members are used to compute a single gain matrix. This matrix will likely be more accurate, since the use of more members leads to less sampling error. Comparing the single ensemble and cross-validation approaches, it has been found that the single ensemble, while being less reliable, may well have an ensemble mean of similar or better quality, when the number k of subensembles that is used in the cross validation is small, but tends to have an ensemble mean of poorer quality when the number k of subensembles is large (Mitchell and Houtekamer 2009, their Figs. 2 and 3). In the case of dense observations that are only weakly correlated with the model state, cross validation can signal degraded quality via an increase of the ensemble spread. This can even lead to catastrophic filter divergence in which the ensemble spread becomes excessively large (Houtekamer and Mitchell 2005, their section 3b). See appendix A for a more comprehensive discussion of filter divergence in EnKF systems.

Related to cross validation is the hierarchical filter (Anderson 2007). Here the observations are assimilated sequentially, one at a time. At each step of the sequential algorithm, an ensemble of EnKFs is run to obtain a confidence factor for the regression coefficient in the scalar Kalman gain . These factors can subsequently serve toward the generation of more optimal localization functions.

d. Sampling error due to limited ensemble size: The rank problem

In the original Kalman filter equations, the matrix is full rank. To obtain this matrix, one would need to integrate the tangent linear model once for each of the NmodelO(100 000 000) coordinates of the numerical model . While this would be an expensive proposition, it would provide us with different directions in phase space (i.e., the eigenvectors of ), for which observations could be used toward a reduction of the uncertainty. In addition, the eigenvectors with significant eigenvalues can be expected to locally follow the attractor of the model, since they have been obtained using model integrations.

An EnKF system with members only provides directions in phase space. Thus, the information in the observations, with typically O(1 000 000), must be projected onto a small number of directions (Lorenc 2003, his section 3b). The fact that and is commonly known as the rank problem; it is the most important approximation/difference with respect to the Kalman filter. How exactly this issue is dealt with has a major impact on the characteristics of an EnKF implementation.

It may be noted that the rank problem is not unique to EnKF systems. For variational data assimilation systems, the NMC method (Parrish and Derber 1992) or the EDA method (Bonavita et al. 2012) typically provides an ensemble of error directions that can be used in the estimate of a matrix . In this case, covariance modeling is used to make full rank. Such modeling is obtained by means of a sequence of well-designed coordinate transformations. Fisher (2004) showed how a wavelet-based covariance model can be used to obtain spatially varying vertical and horizontal correlations. Flow-dependent dynamical balances could also be included. The authors are not aware of reports on rank issues encountered in the subsequent assimilation of O(10 000 000) observations. How to similarly extend the low-rank background ensemble of an EnKF into a flow-dependent high-rank matrix for subsequent use in an EnKF data assimilation system is a major problem that, as we will see, is still not well resolved.

e. Covariance localization

To address the rank problem (section 3d), an intuitive solution is to split the global data assimilation problem into a number of quasi-independent local problems. For each of the local problems, the local directions of the ensemble may permit a fairly relevant approximation of the true uncertainty (Oczkowski et al. 2005). Houtekamer and Mitchell (2005, their section 5b) estimated that, with localization, a 96-member ensemble of global fields can provide an effective dimensionality of over 10 000. In fact, localization is the critical approximation that makes Monte Carlo methods a feasible proposition for the solution of the Kalman filter equations in large problems. There is a consensus in the atmospheric data assimilation community that localization is an essential component of an EnKF algorithm (Hamill et al. 2001; Anderson 2012). As larger ensembles become available, it will likely be optimal to use a less severe covariance localization (Houtekamer and Mitchell 1998, their Fig. 5; Miyoshi et al. 2014).

How localization is implemented exactly is often fairly ad hoc and depends on the particular EnKF algorithm that is used. In the implementation by Houtekamer and Mitchell (1998), observations are not used if their distance from an analysis grid point exceeds some critical value. Similarly, in an LETKF, it is natural to exclude observations beyond a certain distance from the local analysis domain.

1) Localization in the sequential filter

In a sequential filter, localization can be implemented using the product of the ensemble-based covariances with a smooth correlation function �? [Houtekamer and Mitchell (2001), their Eq. (6)]:
e19
where the symbol denotes the Schur (element-wise) product of two matrices and �? is a function of the distance d between two items. For the localizing term in the numerator of Eq. (19), which multiplies , it is thus necessary to define the distance between observations and model coordinates. For the term in the denominator, which multiplies , only the distance between different observations is needed. Note that there can be both integral model variables, such as surface pressure, and integral observation variables, such as satellite radiances, for which it is not clear how to define a location in space.
For the localizing correlation function �?, it is common to use the compactly supported fifth-order piecewise rational function proposed by Gaspari and Cohn [(1999), their Eq. (4.10)]. This function looks like a Gaussian and depends on a single length scale parameter. The compact support leads to zero impact of an observation beyond a certain distance. This property can be exploited to reduce computational cost. Note, however, that without a localization step, it would be possible to perform matrix operations in a different order for a more efficient algorithm [Mandel (2006), his Eq. (4.1); Houtekamer et al. (2014b), their section 4b(6)]. With the significant vertical extent of the model domain, it is advantageous to use localization in both horizontal and vertical directions [Houtekamer et al. (2005), their Eq. (2)]:
e20
where and are the correlation functions for the vertical and horizontal localizations, respectively.

Multiplication with a localizing function is not part of Kalman filter theory; therefore, some of the properties of the Kalman filter do not carry over to an EnKF with localization. One of these properties is the equivalence between processing observations either serially or all at once [section 2c(1)]. The analysis increments will also not be exactly in the space spanned by the background ensemble. The associated imbalance, and return to the model attractor, may lead to rapid adjustment in the forecasts initiated from the analyses.

2) Localization in the LETKF

In an LETKF, calculations are done in ensemble space and consequently the matrix is not represented in physical space. Thus, it is not possible to use Eq. (19) to obtain localized analysis increments. Hunt et al. (2007) proposed to localize instead by gradually increasing observation-error variances for remote observations using the positive exponential function:
e21
where is the distance between observation i and model grid point j, and L is the length scale parameter for the localization.

Greybush et al. (2011) compared the localization of Eq. (21) with the localization of Eq. (19). They found the localization to be more severe than the localization when the same length scales are used in the localizing functions, and consequently, the optimal localization length was found to be longer for the localization. When each scheme was used with its optimal localization length, they found the two techniques to be comparable in performance with respect to both rms analysis error and balance. Sakov and Bertino (2011) also find that the two localization methods should provide very similar results in practice. They note, however, more robust behavior for the localization in the special case of “strong�? data assimilation (i.e., when the analysis causes a big reduction in the uncertainty).

3) Issues with localization

Localization has been presented as a way to implement an approximation to the optimal Kalman filter with a small [] number of ensemble members. It would appear that the severity of the localization is also determined by the desire to maintain physical balances, such as geostrophy or hydrostatic balance, in the analyses (Cohn et al. 1998; Lorenc 2003). To maintain geostrophic balance in a global data assimilation system, Mitchell et al. (2002) suggested limiting the impact of observations to a distance not smaller than about 3000 km. Within a region of this size, the ensemble can only provide directions in phase space. The a priori expectation is that with higher-resolution models and with increasing numbers of observations, it will be necessary to increase to permit balanced analysis increments that cover both large and small scales. This trend is illustrated by the evolution of the operational EnKF at CMC which, as of November 2014, has a horizontal resolution of 50 km as well as a fairly sizable ensemble with members. In scaling experiments for hypothetical future EnKF configurations (Houtekamer et al. 2014b), the maximum ensemble size was . Ensemble sizes of have not historically been considered to be reasonable and may indicate a need for major algorithmic changes. In the currently operational CMC EnKF, the cutoff distance for observations is, depending on the height in the atmosphere, between 2100 and 3000 km (Houtekamer et al. 2014a, their Table 3). For the NCEP EnKF, with 80 members, a shorter distance of 1600 km is used (X. Wang et al. 2013, their section 2). This reflects that optimal localizaton distances are found after experimentation considering a number of factors, of which balance is only one.

In the successive covariance localization (SCL; Zhang et al. 2009a) algorithm, a sequential EnKF is used to first assimilate large-scale information from a small subset of observations using broad localization functions and subsequently smaller-scale features are obtained from larger sets of observations using tighter correlation functions. Proceeding in three steps, the procedure permits obtaining a detailed hurricane analysis including mesoscale vortices in a reasonably balanced large-scale environment. The SCL technique requires some prior knowledge of the characteristic and dynamic scales of the phenomena of interest. It remains a challenge, for example, how to use convective area observations to update environmental (cloud free) state variables for convective storm analysis and prediction.

A similar pragmatic procedure can be applied in the vertical, using severe localization for conventional observations, moderate localization for surface pressure observations, and broad localization for satellite observations reflecting the nonlocal nature of the latter (X. Wang et al. 2013). Also, because horizontal length scales generally increase with height in the atmosphere, horizontal localization lengths can be made to increase with height as in Houtekamer et al. (2014a, their Table 3) and Kleist and Ide (2015a, their Fig. 3).

A more dramatic algorithmic change is provided by the ensemble multiscale filter of Zhou et al. (2008), which effectively replaces, at each analysis time, the prior sample covariance with a multiscale tree. This permits tracking changing features over long distances without spatial localization. For large problems, this method provides a hybrid localization in both space and scale. With a variational solver, Buehner (2012) compared spatial localization, spatial/spectral localization, and wavelet-diagonal approaches. The comparative performance was found to depend on ensemble size. With a 48-member ensemble, spatial/spectral localization gave the best results. With a 12-member ensemble, the wavelet-diagonal approach performed best.

It would of course be desirable (if possible) to have some guiding principle in the choice of a localization method. It could, for instance, be defined as the operation that minimizes the analysis error given a certain limited ensemble size (Zhen and Zhang 2014; Flowerdew 2015). There are, for now, no known analytical methods to derive such an operation for complex situations with multiple observations. Anderson (2007) shows, with his hierarchical filter, that the optimal localization for spatially averaged observations can be quite different from a smooth Gaspari–Cohn function. The method could also be used to determine the significance of correlations between variables of different type. Note that, in variational methods, it is common to specify a zero covariance between background errors of temperature and specific humidity. Localization between different variables was first applied in an EnKF in the context of a carbon cycle assimilation system (Kang et al. 2011). Here, nonzero error covariances were allowed between and the wind field, which affects transport of , but not between and other meteorological variables. This procedure effectively filters the noisy ensemble-based correlations between physically unrelated variables. Lei and Anderson (2014) propose using empirical localization functions (ELFs) determined from OSSEs to give appropriate localization for any potential observation type and kind of state variable. A perhaps more practical approach is to use, as an assimilation cycle proceeds, improving prior estimates of the correlation between classes of variables (Anderson 2016). These prior estimates and the estimated flow-dependent correlations are subsequently combined to obtain a posterior estimate that is to be used to compute analysis increments.

Lange and Craig (2014) present high-resolution data assimilation experiments for convective storms. Here, with a 5-min assimilation window and a localization cutoff length of 8 km, the scales are more than an order of magnitude different from global data assimilation scales and geostrophic balance is perhaps no longer a relevant concept. Nevertheless, similar problems are noted in that severe localization causes imbalance and some sort of adjustment that may degrade subsequent forecasts.

f. Summary

In perfect-model environments, using a stochastic algorithm and cross validation, it is possible to reliably simulate error evolution. In deterministic filters, reliable ensembles can also be obtained, but it is necessary to add an algorithm to relax the underdispersive analysis spread back to the prior spread [section 4a(3)]. How to best verify reliability of realistic EnKF systems in a direct and informative manner is still an open issue. Nevertheless, reliability is a highly desirable feature of ensemble systems. Using a small ensemble, the large spread in a reliable EnKF would reliably show that the analysis quality is low. With increased ensemble size, the reduced spread would reflect the reduced error in the ensemble mean. It does greatly facilitate development work if such changes to the EnKF are reliably reflected in changed ensemble spread (Houtekamer et al. 2014a, their section 6). Whereas such reliability of an EnKF is desirable, it does not in itself guarantee high analysis quality. It is, in principle, possible to have an unreliable EnKF algorithm that, for a given ensemble size, provides smaller ensemble mean errors than a corresponding reliable algorithm (Houtekamer and Mitchell 1998, the middle panels of their Fig. 3).

To obtain high-quality analyses with a small ensemble, it is necessary to localize the analysis increments. We have seen various ways to localize the ensemble covariances using tapering functions that reduce analysis increments associated with distant observations to zero. Unfortunately, a certain amount of experimentation will often be necessary to arrive at an optimal length scale for the tapering function. With grids covering a range of scales that are being observed with a heterogeneous observational network, one would like to have a more flexible and general algorithm for covariance localization. How to best proceed is still an unresolved issue. Ideally, of course, one would be able use an ensemble that is so large that details of the localization method (if any) do not matter (Miyoshi et al. 2014).

4. Methods to increase ensemble spread

As discussed in section 3b, assuming that the forecast model is perfect, and that errors are solely due to propagation of observational error in the assimilation cycle, both EnKF and EDA studies show that the resulting ensembles explain only about a quarter of the error variance of the ensemble mean. In this section, we will describe methods that can be used to arrive at more realistic levels of spread.

We refer (see Table 4) to error sources associated with the forecast model [i.e., Eq. (3)] as “model error�? and to error sources associated with the computation of the analysis increment [Eq. (1)] as “data assimilation error.�? Errors whose origin is unknown, but that could originate either in the forecast model or the data assimilation step will be referred to as “system error�? (Houtekamer and Mitchell 2005).

Table 4.

Schematic overview of sources of system error in an assimilation cycle. Issues only on the left-hand side are specific for the assimilation system and issues on the right-hand side are specific for the forecast model. A horizontal line-up of issues suggests that their impact may be hard to isolate based on observations. In the case of observation and model bias, it may be nearly impossible to attribute an observed error to either the data assimilation system or the model.

Table 4.

Various flavors of covariance inflation, discussed in section 4a, provide bulk methods of changing the ensemble spread to a desired level with no knowledge of the corresponding specific error sources. Other methods aim at specifically increasing spread where a problem is known to exist. For example, stochastic kinetic energy backscatter (section 4b) reintroduces error variance that had been lost due to diffusive processes. Errors associated with model physical parameterizations are discussed in section 4c. We end (section 4d) with a discussion of the realism of current methods to arrive at realistic error levels.

a. Covariance inflation

The oldest method to increase ensemble spread, discussed in section 4a(1), is inspired by Kalman filter theory and consists of adding perturbation fields obtained using prescribed model error covariances. An alternative method, that is convenient in simple modeling environments, is to multiply perturbations with a constant inflation factor [section 4a(2)]. Finally, relaxation methods have been developed to address underdispersion in deterministic filter systems [section 4a(3)].

1) Additive inflation

In Kalman filter theory, it is standard (e.g., Cohn and Parrish 1991; Dee 1995) to assume that the forecast-model error is white in time, with mean zero and covariance matrix . To implement this in the context of an EnKF, one can sample random fields from this matrix as in Houtekamer and Mitchell [2005, their Eqs. (10) and (11)]:
e22
e23
In operational environments, it is fairly common to assume that is simply proportional to the static background error covariance assumed by 3D (or 4D) assimilation systems (Mitchell and Houtekamer 2000):
e24
At operational centers having variational assimilation algorithms, the static background error covariance matrix can likely be expressed as a sequence of operators (Derber and Bouttier 1999) and these same operators can be used to generate the random fields of Eq. (23). The tunable parameter α is, however, ad hoc and does not contribute to an understanding of the origin of the missing error sources. When more realistic parameterizations are developed that describe specific components of the system error, it should be possible to reduce α to reflect the reduced amplitude of the unexplained part of the system error (Mitchell and Houtekamer 2000, their section 5).

As an alternative to using Eqs. (23) and (24), another option is to directly sample from an inventory of differences between forecasts of different length valid at the same time (X. Wang et al. 2013). One thus avoids passing through the covariance modeling steps of the variational algorithm. To still obtain an ensemble of perturbation fields, with a rank of at least , one will, however, need a bigger inventory of difference fields. X. Wang et al. (2013, their section 2) use a full year inventory.

The additive error formulation is competitive with other methods to stabilize the EnKF (Houtekamer et al. 2009; Whitaker and Hamill 2012). However, the errors obtained with Eq. (23) are independent of the flow of the day and show only moderate growth in subsequent longer-range integrations (Hamill and Whitaker 2011). Yang et al. (2015) use ensemble singular vectors to obtain flow-dependent additive perturbations with rapid growth rates. This can, in particular, accelerate the spinup of an EnKF after a cold start (see section 6b for a discussion of spinup).

2) Multiplicative inflation

In simple modeling environments, it is popular to use multiplicative inflation (Anderson and Anderson 1999). Here the background error covariance is multiplied by a tunable factor γ:
e25

Multiplicative inflation tends to work well when γ remains fairly close to 1. When multiplicative inflation is used in realistic atmospheric models, larger values are required and the repeated application of such values can lead to unbounded covariance growth in data-sparse areas (Anderson 2009; Miyoshi et al. 2010). To deal with this problem, and avoid expensive tuning of the inflation parameter, Anderson (2009) and Miyoshi (2011) developed adaptive inflation algorithms.

3) Relaxation to prior ensemble information

When cross validation is not used, as in deterministic filters [section 2b(2)], the ensemble spread becomes particularly underdispersive where observations are assimilated. In relaxation methods, the mean analysis is accepted from the deterministic filter. However, since the ensemble spread obtained using Eq. (10) is known to be too small, only part of the proposed spread reduction is accepted for the ensemble of perturbations.

A formulation that only changes the analysis where data have been assimilated is the relaxation to prior perturbation (RTPP) method of Zhang et al. [2004, their Eq. (5)]. Here the computed analysis perturbations are relaxed back to the prior perturbations using the following:
e26
In the limit , the analysis perturbations are identical to the background perturbations and, after a number of assimilation cycles, the perturbations will gradually converge to the leading Lyapunov (i.e., bred) vectors (Toth and Kalnay 1997; Annan 2004; Whitaker and Hamill 2012). Such an ensemble can be expected to have a relatively low effective dimension. This would not seem desirable because, in data assimilation, a higher dimension would permit a more effective use of a high volume of observations. The resulting ensemble can, however, also be expected to have fairly good balance properties and high growth rates, which are both desirable.
In the relaxation to prior spread (RTPS) algorithm [Whitaker and Hamill (2012), their Eq. (2)], the ensemble spread —as opposed to the ensemble of perturbations—is relaxed back to the prior value using the following:
e27
Ying and Zhang (2015) propose an adaptive algorithm to determine the α parameter of the RTPS method, and tested it in the context of the Lorenz 40-variable model.

Tuning a real data assimilation system for optimal performance, one may find best results with values of , or even higher, in Eqs. (26) or (27). Such severe relaxation or inflation indicates that the pure analysis ensemble is not representative of the ensemble mean analysis error and warrants further investigation. Notably, convergence experiments with larger ensemble sizes can be used to determine if the severe underdispersion is due to the absence of cross validation or due to some other cause.

Whitaker and Hamill (2012) recommend using relaxation methods [Eqs. (26) or (27)] to correct for data assimilation errors, which will be more severe when dense observations are available, and additive inflation [Eq. (22)] to account for model error.

4) Issues with inflation

In operational environments, simple inflation methods as discussed above are critical for maintaining sufficient ensemble spread and having good overall performance (Houtekamer et al. 2009; X. Wang et al. 2013). Unfortunately, they will dilute the impact of the flow-dependent statistics developed in the EnKF. It is important to devote more effort to the identification of error sources in the assimilation cycle; this should permit a more appropriate error sampling and a reduced role of inflation methods.

b. Diffusion and truncation

It has been found that models become more active (i.e., start having faster intrinsic growth rates), when the model resolution is increased and smaller scales can be included. This has, for instance, been observed at ECMWF with the so-called “Lorenz diagrams�? [Fig. 1 in Lorenz (1982); Fig. 16 in Simmons (1996)]. These show a more and more rapid growth of differences with the model dynamics in medium-range forecasts as, in a sequence of operational implementations over a number of years, the spatial resolution of the system is improved. At the same time, the quality of the forecasts has been consistently improved. In an ensemble context, higher instrinsic growth rates of the model and reduced error permit the simulation of a larger fraction of the error with the internal dynamics of the forecast model. Recent EnKF (Bonavita et al. 2015, their section 2) and EDA (Bonavita et al. 2016, their section 4) experiments at ECMWF show a desirable increase in ensemble spread as resolution is increased. These various results reflect that diffusion in the forecast model suppresses activity in the smallest scales. As a result of this diffusion, it can be difficult to maintain a suitable level of spread in an EnKF system and subsequent EPS forecasts (Mitchell et al. 2002, their section 5).

A similar diffusive effect can be the result of balancing methods (see section 5 below and Table 4), which, while necessary to obtain good performance, can filter the signal associated with atmospheric tides as well as with rapidly evolving weather systems.

To be realistic, it would be desirable to have an upscale propagation of uncertainty from beyond the truncation limit to the resolved scales. The stochastic kinetic energy backscatter algorithm (SKEB; Shutts 2005) aims to randomly reinsert kinetic energy that has been overdissipated near the model truncation limit. The overdissipation can be a consequence of explicit and implicit diffusion in the model or of insufficient energy upscaling by parameterizations for gravity wave breaking and deep convection (Charron et al. 2010). The use of the SKEB algorithm has generally been seen to have a positive impact in medium-range ensemble prediction systems (Berner et al. 2009; Charron et al. 2010).

In data assimilation systems, however, it has been difficult to obtain a positive impact from the SKEB algorithm (Houtekamer et al. 2009; Bonavita 2011). Even when model truncation is known to be an important source of error, it is hard to obtain better results with SKEB than with simple additive inflation (Whitaker and Hamill 2012).

In practice, when implementing the SKEB algorithm, many choices need to be made. These include the choice and quantification of processes responsible for the drag, the fraction of dissipated energy that needs to be reintroduced, the spectrum in space and time for the backscatter, and the modes (rotational or gravity) on which the perturbations are to be added. In a mesoscale EnKF, Ha et al. (2015) did obtain good results by perturbing the largest scales, going down to an estimated effective resolution of , with a spatially and temporally constant formulation. Since it is neither flow dependent nor scale selective, this SKEB implementation is in fact similar to the additive formulation of Eq. (22), with the noise being added at every model time step (Ha et al. 2015, their section 3a). Recently, Shutts and Callado Pallares (2011) suggested using the related vorticity confinement algorithm (Steinhoff and Underhill 1994) as an alternative to SKEB.

c. Error in physical parameterizations

In this subsection, we will review error that can be associated with the physical parameterizations of the forecast model. This subject has been extensively investigated in the context of medium-range ensemble prediction systems. Many operational centers contribute with their medium-range ensemble forecasts to the TIGGE project (Swinbank et al. 2015, their Table 1). As shown by online verifications of TIGGE data on the “TIGGE Museum�? (e.g., http://gpvjma.ccs.hpcc.jp/TIGGE/tigge_scatter_diagram.html), state-of-the-art systems now maintain a reasonable agreement between spread and error. Many popular methods have been discussed in the proceedings of the 2011 ECMWF workshop on representing model uncertainty and error in numerical weather and climate prediction models (ECMWF 2011). It would be a priori preferable—if only for considerations of consistency, model spinup, and balance—to use the same treatment of model error in both data assimilation and subsequent medium-range forecast applications. However, as noted by Bonavita (2011), the data assimilation environment is more challenging than the medium-range ensemble prediction environment. This is in part because background errors, which have not yet converged to the dominant modes of instability of the system (Wei and Toth 2003), span a higher dimensional space than longer-lead forecast errors. Another reason is that data assimilation applications require multivariate covariances, whereas the users of EPS forecasts are mostly interested in the mean and standard deviation of scalar forecast variables. It is thus not a priori clear that algorithms, which have been developed and proven successful in medium-range applications, will also be beneficial in data assimilation applications.

1) Physical tendency perturbations

To reflect the uncertainty in parameterized physical processes, Buizza et al. (1999) developed a stochastic algorithm for the perturbation of tendencies of the model physics (SPTP). Here, the output of the model physics for member j is multiplied by a random number to provide a perturbed output of the model physics:
e28
In the original paper, the random number was drawn from a uniform distribution between 0.5 and 1.5, with different values being used every 10° latitude–longitude in space and every 6 h in time. The algorithm was shown to increase spread and improve probabilistic scores. It has since been refined using a number of different time scales and a spectral pattern generator of the type described by Li et al. (2008), and is also in use at CMC (Charron et al. 2010).

Experiments with SPTP in an EnKF context (Houtekamer et al. 2009) were not conclusive. Similarly, Bonavita (2011) concluded that “it may be the case that the spatially correlated error structures introduced by the model error schemes do not represent background error covariances well. Further investigations are required on this issue.�?

One issue with Eq. (28) is that for it forces an inconsistency between the deterministic model physics and the applied tendencies . In extreme cases, this can lead to unstable behavior of the model physics (e.g., Working Group 1, p. viii, ECMWF 2011) and, for instance, motivate a reduction of the amplitude of the random perturbations near the surface. There is also little theoretical support for the simple proportionality between the amplitude of tendency and of the error simulated by Eq. (28) (Shutts and Callado Pallares 2011, their Fig. 1). More refined formulations of SPTP, which could use information about the uncertainty in specific physical processes, would also need to consider correlations between errors in these processes and be considerably more complex.

2) Multimodel, multiphysics, and multiparameter approaches

A pragmatic approach to model error sampling is to use multiple models, multiple physical parameterizations, or multiple parameter values (Krishnamurti et al. 1999). In the EPS context, a substantial gain has been seen from combining medium-range ensemble forecasts from different NWP centers (Mylne et al. 2002; Candille 2009). In an operational data assimilation context, such an approach would require a rapid exchange between centers of a large number of analysis and trial fields. In addition, one would have to resolve potentially complex issues associated with using different definitions for grids and variables. Of course, to obtain a high-rank sampling of model error, it could be necessary to use more than two different models. The approach has, to the knowledge of the authors, not been tried in an ensemble-based assimilation context.

The use of multiphysics approaches, which are simpler to implement, generally has a positive impact also in ensemble-based data assimilation approaches (Fujita et al. 2007; Meng and Zhang 2007, 2008a,b; Houtekamer et al. 2009). The use of multiple physical parameterizations does, in particular, permit the sampling of different possible closure assumptions in deep convection and in boundary layer processes. At operational centers, the multiphysics approach is used in a fairly ad hoc manner to sample uncertainty (Houtekamer 2011) as opposed to being an integral component of the development of state-of-the-art physical parameterizations (Houtekamer and Lefaivre 1997). As part of a multiphysics ensemble, it is possible to also perturb various parameters that play a role in the model physical parameterizations. In the Met Office Global and Regional Ensemble Prediction System (MOGREPS) short-range EPS, for example, eight parameters are perturbed using a stochastic first-order autoregressive process (Bowler et al. 2008, their section 4.1). The multiparameter approach has been pursued systematically in the context of climate projections (Murphy et al. 2011).

Using an extended-state-vector approach, model parameters can be estimated simultaneously with the state variables (e.g., Annan et al. 2005; Aksoy et al. 2006; Ruiz and Pulido 2015). One difficulty is that model parameters usually have a global impact whereas an EnKF will use localization to obtain local estimates for the parameters. A unique parameter value will then usually be obtained using a global averaging procedure. Note that we will see the same issue in the context of bias correction for radiance observations (section 7c). Since there is no natural unstable model to make the estimated parameter uncertainty grow in time along dynamical modes, as in Eq. (3), the estimated parameter values can quickly converge to values that are no longer affected much by new data. To prevent this classical type of filter divergence (see appendix A for more discussion), a conditional covariance inflation can be used to maintain the estimated uncertainty at a desired level (Aksoy et al. 2006). When estimating multiple parameters, another challenge is that errors associated with individual parameters tend to counterbalance each other in many complicated ways (Aksoy et al. 2006). In fact, model failure may occur due to numerically unstable combinations of parameters. Despite the relative power of the estimation procedure, to ensure convergence to meaningful parameter values, it is thus advisable to perform experiments in close collaboration with experts on model physics (Ruiz and Pulido 2015). To this end, it is advisable to perform case studies of well-observed phenomena, as is common for model development (Hu et al. 2010).

3) Future directions

A variety of methods are currently used at operational centers to account for model error. Sometimes different methods can be considered to sample complementary aspects of model error, but other times they are mutually exclusive. For instance, SPTP and the multiphysics or the multiparameter approach will sample the same error in somewhat different manners. When these methods are used simultaneously in the same ensemble system, there could be issues with double counting of errors.

There is currently no consensus across the NWP community on how to best deal with model error. A popular approach is to develop physical parameterizations in a deterministic context and subsequently add external stochastic procedures, such as SPTP, to the model to account for uncertainty. Until now, there has been only limited success with this approach in data assimilation contexts. Palmer (2012) argues that it is preferable to develop physical parameterizations directly in a probabilistic context. While there are some such examples for the parameterization of deep convection (Grell and Dévényi 2002; Plant and Craig 2008), there are at this stage no comprehensive sets of stochastic physical parameterizations that provide an alternative to SPTP. One obstacle for further development is that stochastic parameterizations are likely not helpful in maximizing deterministic forecast skill (Palmer 2012). The move toward a comprehensive set of stochastic physical parameterizations would be a substantial departure from current popular model development practice, and involve a major recalibration of model physics.

An alternative approach is to benefit from the rich “gene pool�? provided by a multimodel and multiphysics ensemble. The required large set of physical parameterizations can emerge naturally from a large enough group of scientists. The use of a multiphysics ensemble verifies well in data assimilation environments [section 4c(2)]. In multimodel EPS systems, at least when the component models are of similar quality, the approach is also successful (Candille 2009). In the context of climate predictions (Pennell and Reichler 2011), it has, however, been shown that different models have similar limitations leading to overconfident climate projections. Perhaps the main problem is that the gene pool evolves in an ad hoc manner with little evidence of an overall intelligent design [see Fig. 1 in Houtekamer (2011) and the related discussion].

d. Realism of error sources

With some imagination, one could create a long list of additional error sources, not yet discussed above, in a data assimilation cycle. Additional errors in the system may arise from, for instance, incorrectly specified observation error statistics (Frehlich 2006; Gorin and Tsyrulnikov 2011; Stewart et al. 2013).

However, as one may guess from even a schematic overview such as in Table 4, it is in practice difficult to arrive at a clear one-to-one connection between a source of error, an algorithm to sample that error source and a set of corresponding observations that can serve to validate the algorithm. One could argue, for instance, that the SKEB algorithm corrects for overdissipation by the model (section 4b). In a validation experiment with real data, it may unfortunately be difficult to separate the dissipative effects due to model truncation from similar effects due to balancing methods, and it may not be possible to arrive at optimal parameter values for the SKEB algorithm that truly correspond to model truncation issues alone.

Different ways to account for system error were tested by Houtekamer et al. (2009) using a global EnKF system. While they found some additional value from the use of multiple parameterizations, the biggest impact on ensemble spread and best results were obtained from the addition of isotropic perturbations [section 4a(1)]. It is not clear what specific error sources correspond to this bulk parameterization. Thus, despite more than 10 years of operational experience with the EnKF algorithm, we do not currently have reasonable and convincing algorithms to describe system error. Thus, although current ad hoc methods work fairly well, it would seem there is significant room for improvement.

5. Balance and length of the assimilation window

In 3D global data assimilation systems, the use of a 6-h data assimilation window has a long history (Rutherford 1976). It has always been beneficial to use balancing methods to control gravity wave noise in such systems (e.g., Machenhauer 1977; Temperton and Roch 1991). With the arrival of 4D assimilation systems, the optimal window length (Fisher and Auvinen 2012) needs to be reevaluated and it is not as obvious as before that balancing methods are indispensable.

In section 5a the need for balancing methods is discussed. Time-filtering methods are the subject of section 5b. In an EnKF environment, it would be desirable to reduce the length of the assimilation window (section 5c), and to this end, it may be necessary to reduce imbalance at the source (section 5d).

a. The need for balancing methods

The forecast model simulates the behavior of the true atmosphere. Because of systematic deficiencies of the forecast model, such as having a truncation limit, the model cannot exactly mimic the dynamics of the atmosphere. In addition, data assimilation methods use approximations such as linearity, isotropy, and locality of dynamics in the computation of an analysis increment. These various simplifications translate into a certain lack of coherence between the model, the atmosphere, the observations, and the assimilation system. This incoherence manifests itself as a rapid adjustment to the model climate, via the generation of gravity waves or sometimes sound waves, when the analysis is integrated forward in time with the dynamical model. In the process, part of the analysis increment will be undone. In extreme cases, large-amplitude gravity waves can render the analysis of little value [section 6.2 in Daley (1991)]. Initialization or balancing methods have been developed to modify the initial state in such a way that it does not give rise to spurious gravity waves in a subsequent model integration.

For the Kalman filter, assuming that the model-error covariances do not project on fast modes, it can be shown that initialization is performed automatically at each analysis time (Cohn and Parrish 1991, their section 4b) (i.e., balance is maintained even without a specific initialization procedure). It could be hoped that global operational 4D data assimilation systems, such as 4D-Var and the EnKF, would inherit this property. However, in practice, balancing methods continue to be necessary (Thépaut and Courtier 1991; Polavarapu et al. 2000; Gauthier and Thépaut 2001; Gauthier et al. 2007).

b. Time-filtering methods

In the 4D-Var method, by virtue of the sequence of integrations that is performed during the iterative assimilation process, it is possible to use a cost term, that penalizes oscillations in a model integration, to arrive at an analysis that minimizes simultaneously the balance constraint and the distance to the observations and the background estimate. In an EnKF, the ensemble of prior model integrations has not yet been affected by the data assimilation and applying a balancing or filtering operation on them should have little impact. In fact, imbalance in the EnKF is thought to mostly result from the localization that is used in the computation of the analysis increments. It is only after the localized increments have been computed, that the model can be used to filter imbalance with either a digital filter or an incremental analysis update (IAU) procedure (Bloom et al. 1996; Kleist and Ide 2015b).

In the digital filter, a time filter is applied to a short model integration to remove rapid oscillations (Lynch and Huang 1992; Fillion et al. 1995). Unfortunately time filters do not have a sharp cutoff and their repeated application to a full atmospheric state, in subsequent steps of an assimilation cycle, can cause a substantial change to important aspects of the atmospheric circulation (e.g., the atmospheric tides and stratospheric flow). To reduce such impacts of a balancing method, it is better to apply it on the analysis increment only (Ballish et al. 1992). Examples are the IAU and incremental digital filter algorithms. As shown by Polavarapu et al. (2004), these two incremental methods can be made to have a similar frequency response.

A recent improvement of the IAU method is to have the analysis increments, that are added during the integration, vary with time (Lorenc et al. 2015). Similar to the IAU and nudging approaches, in the mollified EnKF (Bergemann and Reich 2010), appropriate increments are computed at each time step, and spread out in time using a mollified Dirac delta function. A corresponding stable formulation for the LETKF has been developed by Amezcua et al. (2014). A similar approach that is based on 4D nudging and makes use of the Kalman gain matrix, is proposed by Lei et al. (2012).

c. Toward shorter assimilation windows

In the EnKF, the ensemble propagates covariances between subsequent assimilation windows. This is different from 4D-Var systems, which do require a long window to propagate covariances over a long period in order to “forget�? the initially specified (time invariant) covariance (Pires et al. 1996). In fact, in an EnKF system, there is an advantage to having a short window (Fertig et al. 2007). With a long window, local dynamical features would have time to cross and leave the area where localized increments are significant (Bishop and Hodyss 2009; Buehner et al. 2010a). Thus, when small-scale features need to be assimilated with severe spatial localization, it is necessary to have a corresponding short assimilation window (see section 9c and Fig. 3 for a further discussion and illustration of this issue).

The desire to shorten the assimilation window length is, unfortunately, at odds with the desire to use model integrations of length 6 h or more to improve balance. In a global assimilation cycle, with a window length of 6 h, an IAU or digital filter can be used to filter imbalance with periods of 2–4 h. Inevitably, such a filter will also remove any desired short-period signal with such periods (Fillion et al. 1995, their Fig. 1). Going to a shorter, say 1 h, assimilation window, it will not be possible to effectively filter oscillations with periods of 2–4 h (Huang and Lynch 1993, their section 4a). Thus, for a rapidly updating analysis, in which one wants to selectively filter unbalanced motions with a 2–4-h period from balanced motions with the same periods, it may be necessary to use a different balancing method. Recently, Hamrud et al. (2015) proposed a method to slightly adjust the analysis increment of the wind field to arrive at a more balanced evolution of the surface pressure field. Methods such as this one, which do not require several model time steps, could be required in going to ever shorter window lengths. Alternatively, one could improve the data assimilation procedure so that it introduces only insignificant amounts of imbalance.

Going to shorter windows, the forecast model will start more frequently from initial conditions provided by the analysis. This can aggravate the impact of “precipitation spinup�? that, in particular in global assimilation systems, often results from issues with the initialization of, notably, hydrological variables. One may, for instance, use specific humidity as the only hydrological variable in the control vector of the analysis, whereas the model may also require initial conditions for cloud water. As a result, it may take an integration of a day or more before globally averaged precipitation amounts have settled down to climatological stable amounts (section 13.6 in Daley 1991; Krishnamurti et al. 1988). To diagnose issues caused by frequently performing an analysis, one may compare a continuous model run with a model run that has been interrupted to assimilate no observations. Substantial discontinuity may undo the advantage of a shorter assimilation window.

d. Reduction of sources of imbalance

Sources of imbalance specific to EnKF systems are as follows: model error simulation, covariance localization, model start-up procedures, and recentering methods. We will briefly touch upon each of these issues.

Model error simulation, as already discussed extensively in section 4, is performed using a variety of methods. A difficult issue is that it is not clear what fraction of the true analysis error is unbalanced and if the analysis should entirely remove this error or instead only correct the balanced part of the error and remain on the attractor of the model. All else being equal, one should probably favor methods of model error simulation that do not lead to rapid adjustment after their application. Thus, one could favor RTPP over RTPS [section 4a(3)], and for additive error [section 4a(1)] one could prefer using perturbations that have evolved with the model dynamics (Hamill and Whitaker 2011).

Covariance localization is a potentially important source of imbalance in an EnKF (Lorenc 2003; Houtekamer and Mitchell 2005). The localization can likely be improved by relaxing remote covariances estimates toward a climatological covariance estimate as opposed to relaxing to zero at a cutoff distance [see Eq. (12) in Flowerdew (2015)]. Following Flowerdew (2015), in a scalar example with localization �?, an estimated unlocalized gain and a climatological gain , one would replace
e29
with
e30
This may, in particular, permit using climatological information about weak, but nonzero, distant correlations in addition to significant dynamically evolving ensemble-based correlations at close distances. Since this method permits using the observational information over a larger area, it would likely come at a certain additional computational cost. Alternatively, the localization procedure might be improved by performing it in streamfunction-velocity potential space, rather than in wind component space (Kepert 2009). Perhaps an EnKF system could also benefit from advanced covariance modeling as is used in variational methods (Fisher 2004).

In operational data assimilation methods, it is fairly common to have a control vector that has fewer variables than the model state vector . There is an associated loss of information each time the model run is interrupted for the assimilation of data; an IAU procedure can be designed to minimize this loss (Buehner et al. 2015, their section 2g).

In recentering methods (Zhang et al. 2009b, their Fig. 1; X. Wang et al. 2013), the ensemble mean analysis is replaced by another analysis. One could for instance use a higher-resolution E4DVar (see section 9b) analysis for this purpose. Differences in topography or in physical parameterizations could be a source of imbalance for the low-resolution ensemble runs, because the integrations starting from the recentered analyses will have to converge back to the attractor of the low-resolution forecast model. A local procedure of the same type is to relocate the initial condition of a tropical cyclone to, or close to, an observed position (Chang et al. 2014a). The impact of recentering on balance properties such as small-scale variability, jumpiness, and precipitation spinup has been investigated by Lang et al. (2015).

Sources of imbalance for regional systems will be discussed in section 6a below.

6. Regional data assimilation

The frontier of data assimilation is at the high spatial and temporal resolution of limited-area systems, where we have rapidly developing precipitating systems with complex dynamics. These are observed, at high spatial and temporal frequency, by radar systems. For the case of Doppler wind observations, there is a fairly straightforward link between the model state vector and the observed quantities (Snyder and Zhang 2003; Dowell et al. 2004; Aksoy et al. 2009; Zhang et al. 2009a; Chang et al. 2014b). The use of radar reflectivity information may, however, require the use of realistic microphysical parameterizations because biases in model fields associated with reflectivity could be projected onto other model variables through the ensemble covariances (Dowell et al. 2011).

The field of limited-area ensemble-based data assimilation has recently been the focus of extensive research [refer to Meng and Zhang (2011) for a review on this subject]. Moving from the well-established global applications to limited-area applications comes with a number of new problems, some of which will be encountered by global models when resolution is further increased. Having multiple spatial domains, it is necessary to maintain a certain level of consistency across domains (section 6a). A high-resolution domain may be specially created, for a specific area and relatively short time period, to properly predict the evolution of a specific high-impact weather system like a tropical cyclone or a tornadic supercell thunderstorm. For such applications, the proper initialization of the starting ensemble becomes important (section 6b). Radar observations come in large volumes and some preprocessing steps are necessary (section 6c). For the analysis of convective-scale systems (section 6d), we now have various proof-of-concept applications using both Doppler velocity and reflectivity measurements (Xue et al. 2006; Dowell et al. 2011) as well as near-operational applications (Schraff et al. 2016). For the analysis of hurricanes (section 6e), we have systematic evaluations covering many years (Zhang and Weng 2015). In section 6f, we end this section with a discussion of some open issues.

a. Boundary conditions and consistency across multiple domains

The most unique issue of a limited-area model (LAM) EnKF is the need to represent realistic uncertainties at the lateral boundaries of the limited-area domain. Although preferable, a concurrent global ensemble data assimilation system with the same ensemble size and similar configurations is generally not available to provide the lateral boundary conditions (LBCs) for each of the individual ensemble members in the regional ensemble. Torn et al. (2006) were the first to systematically examine the LBC issues, from which they concluded: 1) the regional phenomena or processes of interest should be sufficiently far away from lateral boundaries (which is generally advised for all regional-scale modeling); and 2) boundary perturbations can be generated by random sampling from a specified multivariate Gaussian distribution or a presumed covariance model such as that of the 3D-Var/4D-Var background covariance (Barker 2005; Meng and Zhang 2008a). Proper tuning and scaling of LBC perturbation tendencies can be difficult and computationally prohibitive. To alleviate this problem, Meng and Zhang (2008b) proposed to use a coarser, but larger-domain, ensemble reinitialized at every assimilation cycle with randomly sampled balanced perturbations. This frequently reinitiated larger-domain ensemble can provide somewhat flow-dependent LBC perturbations for the nested regional EnKF domain.

LAMs often have a lower model top than global models (McTaggart-Cowan et al. 2011). This is an issue for the assimilation of satellite radiance observations that depend partially on information from above the model top boundary of the LAM. For LAM applications with large domains that include oceanic areas, where radiance observations are important, one could consequently decide to keep a high model top as in global systems. On the other hand, for convective-scale assimilation of radar observations over land, where many complementary conventional observations are likely available, having less radiance observations due to a lower model top may not have a substantial negative impact.

Maintaining analysis consistency among multiple model domains with different grid resolutions is a challenge for regional systems that often have nested grids. Any inconsistency is likely to lead to some imbalance and adjustment near the domain boundaries. Here we list four, perhaps largely inevitable, EnKF challenges for models with multiple nests:

  1. Resolution differences: The innovation computed with the high-resolution background of the inner domain will be different from the corresponding innovation for the outer domain. Having more resolved scales, the error of representativeness will be smaller, and this increased accuracy can be reflected in a reduction of the specified error variance R of the observations (Lorenc 1986). Consequently, even in the simple case of just one observation in the inner domain, it is impossible to have exactly the same optimal analysis increment on all domains.

  2. Observation sets: In view of the different resolved scales, different observation sets can be used for the different domains. For the outer domain(s), it may, for instance, not be useful to assimilate radar observations, whereas for the inner domain, these observations may well prove to be critical for the assimilation of convective systems (Snook et al. 2015). It is less clear if and how the analysis inside the inner domain should have access to the observations from the outer domain.

  3. Temporal resolution: For the inner domains, model time steps will need to be shorter than those used for the outer domains. For the assimilation of the high temporal resolution observations in the inner domain, one could even decide to use a different, shorter, window length for the assimilation system than for the assimilation of less frequent data in the outer domain.

  4. Differences in model physics: It is not clear how to account for the discontinuity caused by having different model physical parameterizations for different domains. At some resolution, one will likely see additional model variables associated with hydrometeor species in microphysics schemes. It is to be noted that, even if one would use the same parameterizations in all domains, one would still likely need to vary parameter values to account for resolution differences.

The above inconsistencies are likely to have a negative impact on the quality of the analysis. This will make it difficult for a LAM EnKF to perform better than a piloting global EnKF, when evaluated with probabilistic scores like the CRPS (Hersbach 2000) over the same area using the same verifying observations. How to best demonstrate the added value of higher resolution is an issue that is common to many high-resolution modeling efforts (Mass et al. 2002). In systematic comparisons, the biggest improvements will likely be seen for quantities, such as the probability of having an extreme precipitation event, for which small-scale dynamical features are relatively important. Advanced spatial verification methods, that can possibly address the issue, are reviewed by Gilleland et al. (2010).

b. Initialization of the starting ensemble

For global EnKF systems, it is common to sample random but balanced perturbations from static background error covariances provided by the corresponding variational DA systems to represent the initial condition uncertainties for the EnKF (Houtekamer et al. 2005, their section 3). Given enough spinup time, these initial perturbations usually do not impact the EnKF performance overall, since they are only used once at the very first assimilation cycle. Similar approaches are also used often in large-domain, long-term regional cycling EnKF systems (Barker 2005; Meng and Zhang 2008b; Torn and Hakim 2008; Zhang et al. 2009a). Some regional EnKF systems have also used the global ensemble to generate the initial ensemble [e.g., the HWRF Ensemble Data Assimilation System (HEDAS; Aksoy et al. 2013) and CMC (Chang et al. 2014b, their Fig. 5)].

A challenge for applying EnKFs with LAM models is how to generate an ensemble representing the correct prior errors for features often missing from climatological error statistics, like mesoscale weather systems. This can be critical since most severe convective weather phenomena have rather short life spans. For example, Caya et al. (2005) showed that the EnKF may have some disadvantages in comparison to 4D-Var during the earlier cycles of the analysis of a supercell storm. In the absence of established formulations for generating balanced small-scale convective perturbations, some studies have used purely gridpoint random perturbations to provide initial ensembles having uncertainty at small scales (Snyder and Zhang 2003; Dowell et al. 2004; Aksoy et al. 2009). Unfortunately, such ad hoc perturbations may not properly reflect statistics of the background error and may require a fairly long time to evolve into relevant dynamical features. We simply refer to the gradual process of the EnKF starting to have relevant flow-dependent statistics and track specific dynamical phenomena as “spinup.�? This is different from the classical precipitation spinup of section 5c, which we consider to be more of a balance issue.

Kalnay and Yang (2010) proposed the “running-in-place (RIP)�? technique to accelerate the spinup of the EnKF. The procedure uses the most recent observations to improve previous initial ensemble states, leading to further improved ensemble states at the current time after observations are reassimilated. It can also serve to incrementally extract information from observations in the case of long assimilation window and nonlinear error growth (Yang et al. 2012a). Yang et al. (2012b) perform OSSEs with the RIP procedure to accelerate the spinup in the case of typhoon assimilation and prediction. S. Wang et al. (2013) implemented a similar iterative algorithm in the context of an EnSRF. They find that the algorithm can reach a steady level of state estimation error more quickly than the corresponding noniterated version.

A scheme called “cloud analysis,�? which uses the observed radar reflectivity to directly adjust (nudge) the model temperature and/or moisture profiles, has also been found to be very effective in spinning up the observed storms (Albers et al. 1996; Hu et al. 2006). A nudging component, for radar-derived precipitation rates, has recently also been added to a preoperational high-resolution LETKF data assimilation system with generally positive results (Schraff et al. 2016).

The root of the spinup issue is that the finite-size prior ensemble does not and cannot represent the large non-Gaussian uncertainties associated with the rare events in the small scales. Note that even in a hypothetical continuously cycling large high-resolution global ensemble, there would likely be spinup issues associated with limited observation coverage and systematic errors. The problem can be aggravated by the use of too long data assimilation windows and static (i.e., not adaptive) algorithms.

c. Preprocessing steps for radar observations

Weather radars have now been installed on most continents and composite up-to-date images of, for example, the European, North American, Chinese, and Japanese radars clearly show the areas with precipitation in near–real time. Many stations employ advanced dual-polarization radar. In the United States, it is common to use the Weather Surveillance Radar-1988 Dopplers (WSR-88Ds). Beyond the conventional Doppler velocity and reflectivity measurements, these systems also provide dual-polarization measurements with information on differential reflectivity, specific differential phase, and the copolar correlation coefficient. The latter variables provide information about the distributions of particles in the sampling volume. It is, for now, a challenge to use this wealth of information in automated operational data assimilation systems.

One approach, to deal with high spatiotemporal observations such as from Doppler radars or satellites, is to perform data thinning and quality control through superobservations (e.g., Lindskog et al. 2004; Zhang et al. 2009a; Weng and Zhang 2012). This process combines multiple noisy observations into one high-accuracy “super�? observation (SO). A data-thinning and quality control procedure, minimizing the impact of ground clutter and dealiasing errors, was developed in Zhang et al. (2009a) to generate SOs for ground-based Doppler radars (e.g., WSR-88Ds) for initializing tropical cyclones. A similar SO procedure for data thinning and quality control was used in Weng and Zhang (2012) to assimilate airborne Doppler radar observations and in Sippel et al. (2014) for assimilating High-altitude Imaging Wind and Rain Airborne Profiler (HIWRAP) radar observations.

d. Use of radar observations for convective-scale analyses

Snyder and Zhang (2003) and Zhang et al. (2004) presented the first proof-of-concept applications of the EnKF to assimilate synthetic Doppler velocity data into a cloud model. They demonstrated that EnKF analyses can faithfully approximate the truth in terms of both dynamic and thermodynamic variables of a splitting supercell storm. The first application of an EnKF to assimilate real radar observations appeared in Dowell et al. (2004) for a tornadic supercell thunderstorm. Since then, there has been great progress in using the EnKF to assimilate radar observations both for improving numerical weather prediction and/or for understanding finescale structures and dynamical processes of severe weather. Tong and Xue (2005) assimilated Doppler velocity and radar reflectivity observations with a nonhydrostatic NWP model, which included microphysics. Xue et al. (2006) performed an OSSE to evaluate the impact of combining the data from multiple radar platforms. Aksoy et al. (2010) showed benefits of using zero-reflectivity (clear air) observations in suppressing spurious convective cells during the EnKF assimilation of Doppler velocity in the study of multicellular thunderstorms. Dowell et al. (2011) also found a clear advantage in assimilating reflectivity observations in addition to Doppler velocity for a tornadic thunderstorm. Snook et al. (2011) demonstrated that assimilating data from a dense X-band radar network in addition to WSR-88D data can improve the representation of storm-scale circulations, particularly in the lowest few kilometers of the atmosphere.

A convective-scale LETKF with 2.8-km resolution is currently, summer 2016, being evaluated for operational use at the German Weather Service (DWD). In the initial implementation, latent heat nudging is used for the assimilation of radar-derived precipitation rates (Schraff et al. 2016).

e. Use of radar observations for tropical cyclone analyses

Assimilation of Doppler radar observations has also been shown to improve the accuracy of both track and intensity forecasts for tropical cyclones (Zhang et al. 2009a, 2011; Weng and Zhang 2012). In these studies, the assimilation of radial velocity from either the ground-based WSR-88D network or from airborne Tail Doppler Radars (TDR) was shown to draw the EnKF analysis closer to the true inner-core wind field. Dong and Xue (2013) further show the benefit of assimilating reflectivity, in addition to radial velocity, for a landfalling hurricane. Recently, Zhang and Weng (2015) demonstrated very promising performance with an experimental regional-scale real-time EnKF-based convection-permitting analysis and prediction system, through experiments with the more than 100 applicable NOAA P-3 airborne Doppler missions during the 2008–12 Atlantic hurricane seasons. They showed that the forecasts initialized with the EnKF analysis using airborne Doppler observations led to mean absolute intensity forecast errors at the 24- to 120-h lead forecast times that were 20%–40% lower than for the National Hurricane Center’s official forecasts issued at similar times. In the same cycling EnKF system, Weng and Zhang (2016) showed that considerable improvement could already be obtained, even without the Doppler radar data, from the assimilation of aircraft flight-level and dropsonde observations.

Since 2013, assimilation of airborne Doppler radar observations has been used to improve the vortex initialization for the NOAA regional operational hurricane forecast model (HWRF).

f. Other issues with respect to LAM data assimilation

In addition to the issues already mentioned, progress is needed in the following areas.

  1. Need for flexible localization methods: In a given domain, one may have both radar and radiosonde observations and one may find it optimal to use very different localization lengths for these two data types (Snook et al. 2015). The optimal algorithm for the covariance localization is likely a complex combination involving the changing weather patterns, the observational network, and properties of the forecasting system. In view of the large number of adjustable parameters, it is recommended to use flexible and general localization methods [section 3e(3)].

  2. Complexity of microphysics: For a model to accurately simulate the thermal and microphysical processes observed with dual-polarization radar, it may be necessary to use a second- or higher-moment microphysics scheme (Jung et al. 2010a). Even with such a scheme, it may not be possible to constrain all variables involved (Jung et al. 2010b) without having high-quality direct measurements (e.g., in situ data from aircraft).

    • There is also considerable uncertainty in the calculation of reflectivity, since different microphysics schemes carry different numbers of hydrometeor categories and have different assumptions for droplet size distributions (Morrison et al. 2015). Unfortunately, common multiphysics approaches to sample model uncertainty [section 4c(2)] can only be used in an EnKF context if all (micro)physical parameterizations use the same set of model variables, because the computation of the Kalman gain assumes that all members use the same definition of the control vector.

  3. Error of representativeness: Because of inherently small scales and fast error growth, representativeness error will likely be severe at the convective scale. Large representativeness error can lead to large and sometimes unphysical unbalanced analysis increments. In the assimilation of cloudy radiance observations, Geer and Bauer (2011) proposed an algorithm to inflate observation error variances. Although empirical at present, it has been found beneficial in the ECMWF operational data assimilation algorithm. More research needs to be done, when we move toward assimilating radar reflectivity and dual-polarization measurements.

  4. Complexity of observation error for dual-polarization data: An additional challenge presented by the assimilation of polarimetric radar data is dealing with artifacts, such as differential attenuation, nonuniform beam filling, depolarization streaks, and polarimetric three-body scattering, which may negatively affect the data quality (Ryzhkov 2007; Kumjian 2013). For operational use of dual-polarization measurements in a data assimilation system, it will also be necessary to have reliable automated data processing as well as a comprehensive description of the covariance structure of the observational error.

Despite all the complexity mentioned above, Putnam et al. (2014) presented the first successful real-data experiment using a double-moment microphysics scheme for the EnKF assimilation of microphysical states and polarimetric variables. Evidently a major investment in human and computational resources is still required before dual-polarization observations can be used in automated operational systems.

7. The assimilation of satellite observations

Observations from satellites form an important component of the global observational network. Some of these observations, notably atmospheric motion vectors (AMVs) and radio occultation observations from the global positioning system (GPS-RO), can be considered to contain information valid at points in space and treated in essentially the same way as conventional observations from, for instance, aircraft and radiosondes. In this section, we are concerned with radiance observations that depend on temperature and humidity in a layer of the atmosphere [see Fig. 1.1 in Rodgers (2000)]. Perhaps the best known is the Advanced Microwave Sounding Unit (AMSU) instrument, which has about 10 channels that are mostly sensitive to temperature (AMSU-A) and 5 channels that are mostly sensitive to humidity (AMSU-B). A larger number of radiance observations, O(1000) per vertical profile, is available from the Atmospheric Infrared Sounder (AIRS) and the Infrared Atmospheric Sounding Interferometer (IASI).

In the context of variational assimilation systems, the positive impact of radiance observations has been well documented (Derber and Wu 1998; English et al. 2000; see Fig. 1 in Cardinali 2009). Whereas satellite radiance observations can be successfully assimilated with an EnKF system it would seem their impact is not as substantial as in comparable variational systems (Miyoshi et al. 2010; Bonavita et al. 2015). At this moment, it is not clear if there is one dominating cause for this apparent difference in impact. In this section, we will review differences in the assimilation of radiance observations in common variational and EnKF algorithms. We will cover differences in covariance localization (section 7a), the possibly detrimental impact of dense data (section 7b), different bias-correction algorithms (section 7c), the possible impact of covariance cycling (section 7d), and of false assumptions with respect to observation error covariances (section 7e). In summary, we conclude that the complex situation still needs to be sorted out using controlled experiments (section 7f).

a. Covariance localization

As discussed in sections 5c and 9c, there is an issue with temporal localization in the EnKF, because the area with significant correlations will be advected with the flow between the observation time and the analysis time. This could have a negative impact on the handling of observations near the beginning and end of the assimilation window. While this issue may play a role, it is not specific to the assimilation of radiance observations. In this subsection, we focus on algorithmic differences for the vertical localization.

In the operational global EnKFs at NCEP and CMC, Eq. (19) is used for the localization. Since here the forward operator is used to transform from model space to observation space, prior to the application of the localization, this is known as “radiance-space localization.�? This localization requires defining the distance between the different radiance observations and between the radiance observations and the model variables. Unfortunately, since radiances can be sensitive to conditions in a fairly deep layer of the atmosphere, their vertical location is not well defined. Therefore, it may be preferable to localize directly using “model-space localization�? instead (Campbell et al. 2010):
e31
As demonstrated by Campbell et al. (2010) in a column environment, the localization in Eq. (31) allows the model state to be recovered perfectly in a well-observed case, with as many radiance channels as model levels, in the limit of zero observation error; the same is not true using the localization in Eq. (19). For almost all evaluated parameter values, Campbell et al. (2010) obtained better results with the model-space localization. In an operational context, radiance observations do not have near-zero error. In view of their large number, the limit of zero observational error may, however, still be important. Model-space localization is straightforward to implement in variational algorithms by means of an additional operator in the root of the matrix . Because of the dimension of , it is unfortunately not clear how to efficiently implement Eq. (31) in a sequential EnKF.

The model-space localization would, since it is applied directly to , seem to permit the use of a narrow vertical localization function. However, as shown by Rodgers (2000), the radiance observations have very little information regarding narrow vertical structures (i.e., narrow with respect to the width of the response function). Recently, Lei and Whitaker (2015) have shown that, by neglecting negative background error covariances, model-space localization can cause counterintuitive analysis increments. Thus, which localization method performs best is found to depend of the vertical length scale of both the forward operator and the background error.

We note again [section 3e(3)] that the optimal localization for satellite observations (Anderson 2007) can be quite different from the smooth Gaussian-like functions generally used in either radiance-space or model-space localization.

To make effective use of radiance observations, it may turn out to be necessary to use large ensemble sizes and long vertical localization length scales (Lei and Whitaker 2015; Hamrud et al. 2015).

b. Data density

Because of its flow-dependent covariances, the EnKF works relatively well for sparse observational networks. For example, Compo et al. (2006) created a 100-yr reanalysis of the troposphere using surface pressure observations alone. On the other hand, it would appear that dense networks exacerbate sampling error problems (Thomas et al. 2009, their Fig. 6). Nerger (2015) showed that inconsistencies due to covariance localization can accumulate in particular in the EnSRF, which assimilates observations one at a time. In the stochastic EnKF, which handles observations in a sequence of batches, similar problems can be expected from using an insufficient batch size. Notably for stratospheric analyses, the combination of covariance localization, relatively long horizontal length scales in nature, a sequential algorithm, and a dense network of radiance observations might well cause noise in—or even divergence of—the analysis. To make optimal use of radiance observations in an EnKF context, it may be necessary to use more severe data thinning (Hamrud et al. 2015) or superobbing procedures than for comparable variational systems.

c. Bias-correction procedures

Satellite radiance observations are affected by biases, which are often larger than the amplitude of the random component of the error [section 1 in Dee (2005)]. The use of an effective bias-correction algorithm is critical for the assimilation of radiance observations. In variational assimilation systems, it is fairly common (Derber and Wu 1998; Dee 2005) to extend the model state vector with a few global parameters that are used in the estimation of radiance observation bias. With a variational bias-correction algorithm, a joint estimate of the model state and the bias parameters can be obtained.

As discussed in section 2d, an extended-state-vector algorithm can also be used in the context of an EnKF. Aravéquia et al. [2011, their Eq. (8)] applied an LETKF to estimate bias using the following:
e32
Here, for each radiance channel, there are bias-correction coefficients β. Aravéquia et al. (2011) used two predictors p, namely, scan angle and skin temperature. In an EnKF, it is common to use covariance localization for all state variables and consequently many local estimates are obtained for each global parameter. A unique global estimate can subsequently be obtained using a weighted average [Aravéquia et al. (2011), their Eq. (18)]. These authors concluded that “the assimilation of radiance observations with our proposed strategy is a source of analysis improvement that leads to significant forecast improvement in the Southern Hemisphere midlatitudes (p. 1947).�? The observed improvement was especially large in the upper troposphere and the stratosphere.

A simpler approach is to use a variational bias correction in an “offline�? mode to only estimate the bias parameters (Liu et al. 2012; Miyoshi et al. 2010).

Currently variational and direct estimation analysis algorithms coexist at various operational centers. A synergistic solution, avoiding software duplication, is to simply reuse existing estimates of the bias parameters from a variational algorithm in the EnKF procedure. For instance, as of early 2015, CMC used an offline bias-correction procedure, which itself uses input from the operational deterministic analysis system, to provide bias-corrected observations to both the deterministic and ensemble assimilation components (Houtekamer et al. 2014a, their section 3h). Alternatively, similar procedures can be developed independently for both contexts.

At least from a theoretical point of view, it is appealing that the EnKF environment offers the possibility of dealing with evolving cross correlations between bias parameters and other variables of the model state. In a variational environment, such error information will generally not be available and cross correlations are assumed to be zero [Eq. (14) in Dee 2005].

d. Impact of covariance cycling

A distinguishing feature of the EnKF is that covariance information is passed from one assimilation cycle to the next. The result is a coherent system in which an assimilation cycle benefits from covariance information developed over all previous cycles. Such coherence is certainly desirable when all sources of error are well understood and simulated in the EnKF. The covariances obtained with the EnKF will then properly reflect the balance between error growth with model dynamics and error reduction by data assimilation. As shown by Houtekamer et al. (2005, their Fig. 2), an EnKF can generate very narrow vertical correlations with negative vertical lobes for the temperature field in the stratosphere. These narrow correlations could be a consequence of the vertical stability in the stratosphere and the abundance of radiance observations that, as a consequence of their broad response functions, are relatively informative for error with relatively deep vertical structures (Rodgers 2000; McNally 2004). Unfortunately, as an EnKF manages to gradually reduce simulated error on deep error modes, it may be less able to extract additional information from radiance observations because the more and more dominating shallow error modes are not well resolved by these observations.

In contrast, variational systems commonly use climatological covariances at the beginning of an assimilation cycle. These covariances are often obtained with the NMC method (Parrish and Derber 1992), which uses differences between forecasts issued at different times but valid at the same time.

In reality, stratospheric error patterns can have fairly deep vertical structures. Figure 3 in Dee (2005) shows striking patterns with alternating positive–negative errors, with a well-resolved vertical structure, which he interprets as evidence for biases in the system. These error patterns were obtained from mean analysis increments. Since analysis increments are also used in the generation of error fields in the NMC method, one could speculate that similarly deep vertical structures make these error fields particularly appropriate for the assimilation of radiance observations.

In a cycling EnKF, error fields with a broad vertical structure can be either (i) introduced by the regular introduction of additive model error with covariances obtained with the NMC-method [section 4a(1)] or (ii) maintained by using the RTPP algorithm (Zhang et al. 2004). It would, however, be more direct and likely more effective to actually sample the biases that were at the origin of the alternating error patterns.

e. Assumptions regarding observational error

It is currently a fairly standard practice in the data assimilation community to neglect horizontal error correlations for AMSU-A radiance observations. To compensate for this false assumption, the observation error variances are specified to be substantially larger than estimated (Liu and Rabier 2003). With this strategy, it is possible to use radiance observations at a high horizontal density in variational systems. Because of the cycling of error covariances in an EnKF, there will be a memory of the assumed observation-error covariances. Thus, if one would specify long-range correlations for the observational error, one would expect to see corresponding correlations in the simulated analysis error and, perhaps to a lesser extent, also in subsequently obtained simulated background errors. For the assimilation of ocean altimetry observations with an EnKF, Brankart et al. (2009) found it to be beneficial to use consistent observation-error covariances. Miyoshi et al. (2013) find that correlated observation errors correspond with having more information and, as demonstrated with the 40-variable Lorenz model, permit obtaining better analyses.

There is currently no consensus in the literature regarding the actual importance of observation-error correlations associated with AMSU-A radiance observations. For example, Bormann and Bauer (2010) find fairly negligible correlations using various different estimation algorithms. On the other hand, not making the standard assumption that forecast and observation errors are uncorrelated, Gorin and Tsyrulnikov (2011) find that AMSU-A observation errors have significant correlations out to distances as large as 1000 km. Similarly, using an OSSE, Errico et al. (2013, their Fig. 2) determine length scales of 320 to 610 km for the AMSU-A channels, with the longer length scales being for the higher peaking channels.

f. Recommendations regarding satellite observations

Satellite observations are now routinely used in global EnKF systems at operational centers. There have, however, been only few published results devoted to the assimilation of satellite observations in these systems. Consequently, there are a fair number of open questions, discussed above, regarding the impact or potential impact of satellite observations in EnKF systems. It is recommended to use OSSEs to sort out this situation. In such an environment, it is possible to compare the ensemble covariances with the statistics of available full error fields and investigate many hypotheses toward an improved use of satellite observations.

As we will see in section 8, there is a substantial computational cost associated with the assimilation of high volumes of radiance observations in an EnKF. It is thus important to establish that the satellite observations are used in a cost-effective manner.

It is worth mentioning that, despite the challenges, one recent study by Zhang et al. (2016) showed very promising results in using a regional EnKF system to assimilate both simulated and real-data all-sky infrared brightness temperature observations from geostationary satellites for tropical cyclone analysis and prediction.

8. Computational aspects

Much of the recent popularity of the EnKF algorithm is related to the ongoing improvements to computational platforms (Isaksen 2012, their Fig. 1). Section 8a deals with the general qualitative changes that can be expected as EnKF configurations become more computationally challenging with higher model resolution, more observations, and more ensemble members. The workload associated with various aspects of the EnKF algorithm is the subject of section 8b. Challenges associated with having more computer cores are discussed in section 8c, and practical issues associated with running an operational EnKF are the subject of section 8d. In section 8e, we assume that practical issues can be resolved and speculate on the feasibility of reaching and crossing the “gray zone,�? characterized by a horizontal resolution between 1 and 10 km, using a global EnKF system. Finally, in section 8f, we summarize and discuss the impact of evolving computing systems on the general scientific endeavor.

a. Parameters with an impact on quality

The number of floating point operations, memory usage, and communication costs of an EnKF can often be expressed using the three parameters , , and . These parameters also have a direct impact on analysis quality and any changes to any one of these parameters can motivate a readjustment of other parameters of the EnKF. We will now summarize the expected dependencies.

  • : The number of model coordinates.

    • A higher-resolution model will normally be more accurate and less diffusive. With this higher quality, it should be possible to reduce the amount of simulated model error (Bonavita et al. 2015). The contribution of the SKEB algorithm should diminish naturally as the truncation limit is changed. The increased fraction of dynamically evolving error will normally be a factor toward further improvement of the EnKF (section 3b). With the higher resolution, to support the assimilation of newly resolved features, one may want to increase to observe the new features and to increase the rank of and have corresponding error directions. In view of the reduced characteristic lifetime of the newly resolved scales, it will be tempting to reduce the length of the assimilation window. Naturally, a reduced time step may have to be used for the higher-resolution model. Since at operational centers the development of higher-resolution configurations of the forecast model is usually done in the context of a high-resolution deterministic system, increasing in the lower-resolution ensemble context is likely to be fairly straightforward.

  • : The number of observations.

    • At most operational centers, there is no obvious (low cost) way to increase the number of independent observations one has access to. However, with more observations there would be more information available, and this should lead to higher-quality analyses. When increases there will likely be a need to have corresponding additional error directions in , so it may be necessary to increase as well. Alternatively, it could be decided to use either a more severe covariance localization or a multiscale approach.

    • We also note that more observations can be obtained by changing the parameters of data-selection and thinning procedures. These procedures serve to, at least partly, compensate for the neglect of observation-error correlations. In this complex situation, it is possible that using fewer observations leads to an improved analysis quality (Hamrud et al. 2015). This is indicative of an underlying problem that needs to be addressed at a more fundamental level.

  • : The number of ensemble members.

    • Having more members permits more accurate estimates of correlations and directly reduces the rank problem. With increased , to benefit from weak but distant correlations, it will likely be possible to use a less severe covariance localization (Houtekamer and Mitchell 1998, their Fig. 5). With very large ensembles, covariance localization may not even be necessary (Kunii 2014). Another alternative, benefitting from the increased rank provided by the larger ensemble, is to assimilate more observations.

For optimal performance, we need to increase each of these three parameters in a balanced manner. Normally, when additional computational resources become available, some experiments will be performed to decide on an optimal change of parameter values (Houtekamer et al. 2014a, their section 5d). One may, for instance, favor the improved quality of a higher-resolution model over the improved correlation estimates from a larger ensemble (Kunii 2014).

The parameter values that were in use for global EnKF systems at operational centers in February 2015 are listed in Table 1.

b. Overview of current parallel algorithms

To run an EnKF, one has to perform three main tasks: (i) perform an ensemble of model integrations [Eq. (5)], (ii) evaluate the forward operator to obtain values and form extended state vectors [Eq. (18)], and (iii) compute analysis increments [Eq. (4)]. These three tasks can, at least in principle, be performed in sequence with independent parallelization strategies. The cost of step (i) dominates over other costs at CMC and for the research and development EnKF at ECMWF. For NCEP relative costs depend on the precise configuration, but (i) is still most expensive for the EnKF component of the hybrid 3DEnVar/EnKF algorithm. Including the cost due to reduced time steps at higher resolution, the model cost will be approximately . With the trivial parallelization over and good parallelization of the dynamical model , the parallelization of (i) is not a major concern in EnKF development. With regard to the forward operator in step (ii), we have a trivial parallelization over as well as a possible further parallelization over . Clearly, most research efforts have gone into step (iii), for which there is no obvious embarrassingly parallel algorithm due to matrix operations in Eq. (2).

Tippett et al. (2003, their Table 1) give an overview of the scaling of different EnKF algorithms. As is evident from their use at operational centers, the stochastic EnKF, the deterministic EnSRF, and the LETKF are all suitable for large atmospheric data assimilation applications. For detailed descriptions of the stochastic EnKF, and deterministic EnSRF, and LETKF algorithms, see Houtekamer et al. (2014b), Whitaker et al. (2008), and Szunyogh et al. (2008), respectively.

The workload of both the stochastic EnKF and the deterministic EnSRF scales as . The main cost is in the matrix multiplication associated with Eq. (6), which scales well because computations at different grid points are independent, permitting parallelization over [Houtekamer et al. (2014b), their section 4b(6)]. Note that the sparseness resulting from covariance localization, with a function of compact support, can be used toward substantial computational savings. When having a more powerful computational platform, one is often interested in solving more challenging computational problems as opposed to solving an existing problem in a shorter time. Weak scaling experiments ascertain if the wall clock time remains approximately constant when the number of computer cores is increased in proportion to the problem size. Houtekamer et al. (2014b) observe good weak scaling when increasing and , but not when increasing . Thus, it could prove difficult to efficiently use more cores toward assimilating more observations.

The workload of the LETKF scales as (Tippett et al. 2003; Hunt et al. 2007). In the LETKF, when going to the limit of large ensemble size , the main computational cost is in the eigenvalue decomposition of an matrix [Eq. (16)], which scales as (Miyoshi et al. 2014). When going to the limit of large or , other computational challenges may need to be addressed.

c. Evolution of computer architecture

One of the exciting yet challenging aspects of modern supercomputers is the steady increase in the number of computer cores (Hager and Wellein 2011; Isaksen 2012). In Fig. 2, we show the increase in the average number of cores for the world’s 10 fastest computers over the last decade (Strohmaier et al. 2015). This number has increased from 23 872 in 2005 to 800 616 in 2015. Note, however, that the last 3 years have seen a slight decrease in the average number of cores. For now, at NWP centers, EnKF applications have been limited to cores (Table 1). For such an application to make efficient use of the computer, simply applying Amdahl’s law, the parallel fraction of the code will need to be near 0.999. It would seem reasonable to hope that O(10 000) cores will soon become available for operational EnKF systems. For the corresponding code, the parallel fraction will need to be near 0.9999.

Fig. 2.
Fig. 2.

The average number of cores for the top 10 computers of the top500 list (Strohmaier et al. 2015). For each year, the November list has been used.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0440.1

It is hard to estimate the difficulty of reaching a parallel fraction of 0.9999. Using the CMC EnKF as an example, it could be straightforward if the additional computational power is used to increase , but it could require a substantial redesign of algorithms involving communications if, instead, is increased. A reduction of available memory per core could also necessitate substantial modifications. For LETKF experiments with 5760 cores on the K computer, the I/O system was reported to take a substantial fraction of the time (Kunii 2014). It is, of course, possible that different characteristics of new computers, such as permitting faster communications between nodes or having a more powerful I/O system, will lead to higher parallel fractions with essentially no changes to the algorithm.

The positive impact of a close collaboration between computer scientists and data assimilation specialists is illustrated by Miyoshi et al. (2014), who obtained an eightfold speedup by using a more efficient algorithm for eigenvector decomposition. This speedup could be used to double or increase . Similarly, Houtekamer et al. (2014b) report a substantial speedup from a more efficient use of the computer caches (their section 3b) as well as from improved communication patterns (their appendix C). The relatively large potential impact of optimization efforts makes it difficult to objectively compare the suitability of the stochastic EnKF, deterministic EnSRF, and LETKF for current or future High Performance Computing (HPC) environments.

d. Practical issues

When running an EnKF on a supercomputer, one is likely to encounter a number of practical issues.

  • Job class flooding.

    • On many computer systems, a job queueing system is used to manage the tasks submitted by a wide range of users. Normally, this system will attempt to give comparable users a comparable share of the system. Many ensemble applications, such as performing an ensemble of model integrations with Eq. (5), are trivial to parallelize in that one can simply submit a large ensemble of nearly identical jobs to the same job class. Such a strategy may not be well received by the community of users, and artificial constraints may result. More fundamentally, when many identical tasks are run simultaneously on a computer, such as when O(10 000) tasks independently try to access the same information, one may well hit physical limits of the computer system, such as having a limited number of nodes dedicated to input and output processes.

  • Storage capacity.

    • It is often practical and desirable to keep meteorological fields corresponding to experiments available for future reference. As explained by Houtekamer et al. (2014b), the number of files that can be generated by an EnKF is substantial and the systematic storage of all relevant fields will likely be a challenge for disk and tape archiving systems.

  • Design of experiments.

    • An atmospheric EnKF can be coupled to many other systems. One could, for instance, anticipate benefits from two-way coupling with a land surface EnKF and with an ocean EnKF (Sluka et al. 2016). This coupling comes with an additional technical complexity, because the different systems will likely have been designed in slightly different ways and parameters—like the assimilation window length—will have different optimal values for different components. It may also require a substantial team effort to analyze and interpret the interactions in the coupled system.

e. Approaching the gray zone

Here, we will define the gray zone to be the range of scales where a parameterization of deep convection may or may not be necessary. Following “The Grey Zone Project�? (http://projects.knmi.nl/greyzone), this zone is defined to be between horizontal resolutions of 1 and 10 km. In addition, it will be assumed that it is not clear how to deal with convection in this resolution range; therefore, numerical modeling and EnKF data assimilation in this zone should be avoided. We will speculate on the possibility of having a future global EnKF arrive at the gray zone (i.e., at 10-km resolution), and on the possibility of jumping beyond it to a resolution of 1 km.

In Table 5, we provide some highly speculative parameters for EnKF systems going into the gray zone. The future true configuration parameters would need to be determined from a cost–benefit analysis (Houtekamer et al. 2014a, their section 5d). The analysis cost is taken to scale as and the model cost scales as . Note that the rapid increase in compensates for the reduced model time step, and the analysis cost is projected to remain smaller than the cost of running the model. As discussed earlier, several EnKF algorithms exist, and it may well be decided to move to a more efficient algorithm if that is available. Similarly, the given number of longitudes and latitudes assume that a globally uniform latitude–longitude grid is used, but different grid topologies may well be preferable. The predicted power consumption of the last listed year would require several dedicated nuclear power plants, which obviously would not be reasonable. It follows from this observation that the gray zone cannot be crossed with the CMC EnKF algorithm and currently known computer technology.

Table 5.

Parameters of current and hypothetical future EnKF configurations. Computer systems are assumed to become twice as powerful every two years (as evaluated using the LINPACK test). For illustration purposes, parameters of a global uniform grid are given. The parameters , , , , and are assumed to increase in the same relative manner as between 2011 and 2014 at the CMC. Configurations are given every 8 years from 2014 onward. The exact 10-km resolution, that was used to define the gray zone, would be reached in 2032, which is specially mentioned. The cost of the analysis, indicated by “cost �? scales as the product of , , and . The cost of the model scales as . From 2030 onward, the cost of the analysis is also shown scaled by a factor of 160, which reflects the difference in executing speed (as measured with LINPACK) between the 2015 CMC and Tianhe-2 supercomputers. The latter computer is also used to estimate power consumption from 2030 onward.

Table 5.

With a projected doubling of computer capacity every two years, the next generation of CMC scientists—and likely scientists at various other NWP centers—could hope to reach the gray zone in 2032. At that point in time, the CMC computer would need to be approximately 4 times more powerful than the world’s current fastest supercomputer (the Tianhe-2 in Guangzhou, China). Configurations approaching the gray zone, with a resolution of, say, 15 km might well be feasible on the world’s fastest computers in the coming years. A dual-resolution approach, such as with a 4DEnVar algorithm (see section 9c below), could be used to obtain one or a few members beyond the gray zone.

f. Summary

The tremendous recent development of ensemble-based data assimilation systems has been made possible by the advent of massively parallel high-performance computing systems. Current EnKF algorithms obtain parallel fractions between 0.999 and 0.9999. To make efficient use of order 1 000 000 computer cores, it will be necessary to achieve a parallel fraction of 0.999 999, which is two orders of magnitude better than what is currently available. To approach such a fraction will be a substantial challenge.

Inevitably, these developments are increasing the gap between pilot studies with simple chaotic models (Lorenz 2005) that can be run on a laptop computer and applications with current NWP systems (Houtekamer et al. 2014a; Hamrud et al. 2015) that require HPC environments. To span the gap, going from simple to complex, it would help greatly to use models of intermediate complexity, like the SPEEDY model (Molteni 2003), to develop and test newly proposed algorithms.

Inevitably, it will continue to be necessary to have contributions from a range of people with backgrounds in the atmospheric, statistical, and computer sciences to develop future assimilation systems.

9. Hybrids with variational and EnKF components

Hybrid systems now arise at operational centers, as elements of emerging EnKF systems (e.g., Whitaker et al. 2008; Houtekamer et al. 2014a) are being combined with elements of well-functioning preexisting variational assimilation systems (e.g., Parrish and Derber 1992; Rabier et al. 2000; Gauthier et al. 2007). In section 9a, the combination of static and ensemble-based covariances into a hybrid covariance matrix is discussed. It is shown in section 9b how, in a traditional 4D-variational framework, an extended control variable can describe the analysis increment as a sum of a static and ensemble-based components. The 4DEnVar ensemble-variational method, which is discussed in section 9c, is similar but uses the EnKF ensemble, instead of linearized models, to estimate temporal covariances in the assimilation window. In the hybrid gain algorithm (section 9d), analysis increments from EnKF and variational systems are simply weighted to obtain an optimal analysis. A concluding discussion is provided in section 9e.

a. Hybrid background error covariances

In the oldest form of hybrid, the ensemble covariance is mixed (hybridized) with a static background error covariance , as proposed in Hamill and Snyder [2000, their Eq. (4)]:
e33
Here, the tuneable parameter β is used to obtain an optimal weighting of the static and ensemble covariances. The ensemble covariance has rank at most . In contrast, by virtue of its construction using a diagonal matrix and a sequence of transformations [Derber and Bouttier (1999), their Eq. (12)], the static covariance is normally full rank. The weighted sum of the low-rank ensemble covariances and the full-rank static (climatological) background error covariances is full rank. The use of the hybrid covariance will thus reduce the effects of sampling error and rank deficiency (e.g., Wang et al. 2008; Zhang et al. 2009b; X. Wang et al. 2013; Kleist and Ide 2015a,b).
Static background error covariances are also often used in an EnKF framework, in particular in the form of additive inflation by drawing random perturbations from a static background covariance [section 4a(1)]. In that context, the two covariance terms are not averaged but added and the static term is interpreted as a model error covariance [Mitchell and Houtekamer (2000), their Eq. (4)]:
e34
where is prediction error obtained with a perfect model. In the ensemble implementation with Eq. (22), the addition of model error may sample new directions in phase space, but the rank of the -member ensemble cannot increase beyond (Meng and Zhang 2008a).
At an operational center, one could well decide to use only a single static covariance matrix for both the weighted average of the hybrid [Eq. (33)] and the model error simulation in the EnKF [Eq. (34)]. Substituting Eq. (34) into Eq. (33), we find that
e35
Comparing Eqs. (34) and (35), it is seen that the hybrid matrix gives a reduced weight to flow-dependent covariances as compared to the matrix of the EnKF. Equivalently, it can be stated that the model error sampling is relatively more important in the hybrid system where it also serves toward further regularization of the rank problem.

In the EnKF framework, the addition of perturbations generated from static error covariances may also be done at the stage of the posterior covariance (Hunt et al. 2004; Houtekamer and Mitchell 2005) allowing the forecast model , to evolve these perturbations to the subsequent assimilation window.

b. E4DVar with the α control variable

With the α control variable approach proposed by Lorenc (2003), the cost function J in the hybrid ensemble-variational system (EnVar) can be expressed as
e36
Here, is a localizing covariance and α contains smoothly varying fields of weights for the ensemble of background error fields. The background term in the cost function [Eq. (36)] is then computed as a function of the traditional preconditioning control variable (Barker et al. 2012) and the ensemble-based (α) control variable (Lorenc 2003). The cost of the distance to the observations is measured using as usual in variational minimization.
The α control variable and direct covariance combination approaches, using Eq. (33), have been proven to be theoretically equivalent by Wang et al. (2007). Following these authors, we have
e37
e38
e39
e40
where the total analysis increment is the sum of a static component and an ensemble-based component .

As stated by Lorenc (2003), it is straightforward to include a temporal component as in 4D-Var algorithms. Here ensembles are used to define an increment valid at the beginning of the assimilation window and the tangent linear and adjoint forecast models are used to compute the distance to observations and corresponding gradients. In such an E4DVar algorithm, “within the window the time evolution of covariances is modelled by 4D-Var, and the EnKF models their evolution from one window to the next�? (p. 3199 in Lorenc 2003). The E4DVar algorithm is known to outperform uncoupled EnKF and 4D-Var algorithms (Zhang et al. 2009b; Buehner et al. 2010a,b). A hybrid of an EnKF and a 4D-Var is currently used in the global system of the Met Office (Clayton et al. 2013).

An E4DVar was implemented with a mesoscale model (WRF) by Zhang and Zhang (2012) and Poterjoy and Zhang (2014).

c. Not using linearized models with 4DEnVar

In the E4DVar algorithm, discussed above, it is assumed that an accurate linearized forecast model and its adjoint exist and are available for use in data assimilation applications. An alternative, exploited in 4DEnVar algorithms, is to obtain information on the temporal evolution from the ensemble of background trajectories of an EnKF (Liu et al. 2008).

The control vector that is used in the 4DEnVar algorithm has a static and an ensemble component as in the standard α control vector. The static component is held constant over the assimilation window and for the ensemble component one will actually use the ensemble of nonlinear model integrations. The lack of evolution of the static component is a disadvantage as compared to corresponding EnKF, 4D-Var, and E4DVar implementations. Also in view of issues with the interpretation of the static term (see section 9a), it could be decided to leave the handling of model or system error to the EnKF component of the 4DEnVar [i.e., set as in Wang and Lei (2014)].

Another issue is that, in the context of either an EnKF or a 4DEnVar, it is not clear how to localize in a coherent manner in both space and time. This can be problematic when there is rapid advection of information during the assimilation time window (Bishop and Hodyss 2009; Buehner et al. 2010a). The issue is illustrated in Fig. 3. Here it is shown how a local volume in space can evolve in a Lagrangian manner during an assimilation window from areas A to B and finally C. In E4DVar this evolution is handled in a coherent manner by the linearized forecast model that is used to transport the localized initial covariances across the time window. In EnKF systems, of the type discussed in this paper, the localization is always for area A. Thus, at the central time, the analysis increment would be erroneously confined to area A and probably be rather small and noisy because the significant ensemble-based correlations are only expected in area B. The flow of information with time has been examined extensively in the context of observation targeting studies (Bishop and Toth 1999). To reduce the detrimental impact of the neglected Lagrangian evolution in the 4DEnVar, it could be decided to use relatively generous localization lengths, or alternatively, one could reduce the length of the assimilation window.

Fig. 3.
Fig. 3.

Schematic diagram to dramatize the difficulty of localization within the assimilation window. An observation is made in area A and valid at the beginning of time window. During the time window a Lagrangian particle can move to area B, at the central analysis time, and to area C at the final time. At the central time, one may want to have the impact of the observation localized in the circle around area B and at the final time in the circle around area C. The assumption is that the localizing area is very small compared to the distance the Lagrangian particle can cover during the assimilation window.

Citation: Monthly Weather Review 144, 12; 10.1175/MWR-D-15-0440.1

As a consequence of these issues, the analysis quality obtained using a 4DEnVar framework will likely be inferior to the quality obtained with a E4DVar system (Clayton et al. 2013; Lorenc et al. 2015; Poterjoy and Zhang 2015, 2016).

The first operational implementation of a 4D ensemble-variational (4DEnVar) algorithm occurred at the CMC where, since November 2014, it is used for the deterministic global (Buehner et al. 2015) and regional (Caron et al. 2015) analyses. Currently, NCEP is also in the process of upgrading to 4DEnVar from the current 3D composite system (3DEnVar) (D. Kleist 2015, personal communication).

d. The hybrid gain algorithm

So far, it has been assumed that the EnKF system and the variational solver have each been developed fairly independently to a comparable level of quality such that it becomes advantageous for an operational center to create a hybrid system.

Working in the context of the Lorenz 40-component model, Penny (2014) proposed to hybridize the gain matrices of a variational and an EnKF assimilation system by weighting the analyses from the two component systems (which here were a 3D-Var and an LETKF). This is equivalent to recentering the EnKF mean analysis and the deterministic analysis using the following:
e41
Here γ is some tuneable constant that depends on the relative quality of the two systems. The algorithm is illustrated in Fig. 4 of Bonavita et al. (2015).

The approach has been explored further at NCEP in the context of a global ocean data assimilation system (Penny et al. 2015). Similar to what is seen with the more common hybrids, which combine the ensemble and background error covariances (section 9a), the hybrid gain approach provides benefit when the ensemble size is small, because the analysis increment is not restricted to the space spanned by ensemble members. Because of the use of a climatological covariance in the variational component, it also improves the analysis in regions where observations are sparse.

Bonavita et al. (2015) applied Penny’s hybrid gain algorithm to hybridize the gain matrices from independent EnKF and 4D-Var systems. They found that the hybrid gain system outperformed its two component systems (i.e., EnKF and static B 4D-Var). It was also competitive with a reduced-resolution version of the hybrid 4D-Var–EDA system used operationally at ECMWF. That operational system uses an ensemble of independent adjoint-based 4D-Var analyses, where the ensemble is used to estimate flow-dependent error variances (Bonavita et al. 2012). A practical advantage of the hybrid gain approach is that it permits independent parallel development of EnKF and variational systems to have a beneficial impact on the resulting hybrid system (Bonavita et al. 2015).

e. Open issues and recommendations

Generally speaking, among hybrid algorithms, it would seem that the E4DVar method is superior to the 4DEnVar methods. This may be mostly due to the latter’s use of a static term in the hybrid background error covariances and to the difficulty of localizing covariances over a long window. Possibly, removing the static term and using less localization with larger ensembles and a shorter assimilation window would invert the situation with now superior results for the 4DEnVar due to its use of a nonlinear model to transport covariance information over the assimilation window. The use of a 4DEnVar algorithm may also promote faster model development because with this approach accurate linearized models would no longer be required by the data assimilation systems.

Beyond discipline, it is not clear if there is a general development strategy that will minimize configuration differences between the components of the hybrid. An extreme solution would be to abandon the hybrid and either (i) obtain all required analyses from the direct solver of the EnKF or (ii) obtain all required analyses from an ensemble of variational solutions. Current comparisons of computational cost would likely favor (i), but this situation could change if efficient parallel algorithms for ensemble minimization become available. On the other hand, it would seem from in particular the experience with the assimilation of radiance observations that higher-quality results can be obtained with (ii). There is, however, no generally accepted explanation for this apparently higher quality.

It is hard to predict how the quality of the component variational and EnKF systems will evolve in the future. For the foreseeable future, the hybrid gain algorithm can serve as a simple and practical method to combine desirable aspects of both families of algorithms.

10. Summary and discussion

In recent years, there has been a flurry of activity with respect to global (e.g., Table 1) and regional EnKF systems. Houtekamer et al. (2014a) see a virtuous cycle, in which increasingly powerful computer systems, an improved sampling of errors with larger ensembles, more active higher-resolution dynamics, and more realistic models with less systematic problems all conspire toward an ever higher-quality approximation of the underlying Kalman filter. Nevertheless, as we will summarize below, the burgeoning field of ensemble-based data assimilation is full of unresolved issues.

a. Stochastic or deterministic filters

Perhaps the main split in the EnKF community is with respect to the use of a stochastic or a deterministic algorithm.

With the stochastic algorithm, we have the perhaps naive hope that eventually we will manage to identify all the major sources of error (Table 4) and find appropriate ways to sample these in the Monte Carlo framework. In this view, localization deals only with the rank problem associated with a small ensemble size and additive covariance inflation is a stand-in for (to be developed) algorithms that will account for, and sample, specific weaknesses in the system.

The deterministic algorithm is more pragmatic. The aim is to obtain an accurate best estimate using a minimum number of ensemble members and, to this end, it is better to avoid random perturbations to observations. If, for some reason, the ensemble statistics are not reliable they can be corrected by an appropriate adaptive algorithm. This can be a localization or relaxation algorithm of which the parameters are tuned to minimize analysis and prediction error.

The current split between stochastic algorithms with cross validation and deterministic algorithms with relaxation methods could possibly be resolved if cross validation could be implemented in deterministic algorithms.

b. The nature of system error

Properly accounting for estimated uncertainties of the observations and letting errors evolve and grow with the dynamics of the model, one arrives at estimated error levels for the analysis or background that are much too optimistic. Ideally, the missing error sources would be identified and included in the Monte Carlo error simulation of the EnKF system. To have a comprehensive quantitative list of the weaknesses in an NWP system would be of immense value for the realism of the EnKF system, and it would greatly facilitate work toward improvement of the NWP system itself. The task, however, is daunting. Since the number of possible issues in the system is huge, it may prove easier to find a needle in a haystack with dimensions (Tarantola 2006).

Lacking a comprehensive list, a variety of bulk methods has been developed to maintain a realistic amount of spread in an EnKF system (section 4). Simple methods for covariance inflation (section 4a), like relaxation to prior, additive covariance inflation or a combination of these two methods, provide generally satisfying reliability and it has so far proven difficult to obtain better results with more sophisticated methods. Some improvement has, however, been observed from using multiple physical parameterizations [section 4c(2)] to sample the “model error�? component of the total “system error.�? There is currently no consensus in the community on how to move toward schemes that comprehensively sample model error [section 4c(3)]. Options include the development of either a complete set of stochastic parameterizations, covering all parameterized processes, or a comprehensive set of possible deterministic parameterizations.

c. Going beyond the synoptic scales

The global EnKF systems originated in environments appropriate for synoptic-scale data assimilation. At these scales, the temporal evolution is slow, fairly well described by the numerical model, and it is sufficient to do the analysis every 6 h. Covariance localization can be done in various ways, but must use relatively broad functions— km in the horizontal—to preserve geostrophic balance in the analysis increments. Traditional balancing methods, like the digital filter, can be used to filter any undesired gravity waves.

Going to shorter temporal scales and shorter assimilation window lengths, it would seem advantageous to move to the incremental analysis update (IAU) procedure, in which analysis increments are gradually added as a forcing to a continuous model integration. When the window length reduces, the IAU will, however, start filtering shorter temporal scales and may become less effective as a balancing method. Perhaps, we need to return to balancing methods that act on tendencies observed at the initial model time step (Hamrud et al. 2015).

Adding shorter horizontal scales creates a difficult superposition of large-scale, almost geostrophic, dynamics and small-scale intermittent convective activity. Traditional localization methods, with prescribed smooth functions, cannot deal with this situation unless the ensemble is made large enough that no localization is required for small scales. Miyoshi and Kondo (2013) propose to use a multiscale localization approach, where a regular small-scale analysis obtained with severe localization is augmented with remote larger-scale components obtained from analyses using smoothed perturbations. Alternatively, one may have to use a flow-dependent and observation-type-dependent localization procedure. For data assimilation in hurricane environments, successive covariance localization (SCL; Zhang et al. 2009a) permits zooming in on interesting features in the well-observed central area. In thunderstorm environments, it is common to use observation-type-dependent localization (Meng and Zhang 2008a; Snook et al. 2015). In particular in the context of sequential algorithms, it could appear that one is free to localize each observation with specific criteria. A caveat is that interaction between the sequential processing and the localization can lead to instability (Nerger 2015). How to best localize the information from multiple observations is an unresolved problem [section 3e(3)].

There is a significant user need for accurate forecasts of severe convective storms. Such systems are often observed at high temporal and spatial resolution with dual-polarization radar. Likely, for a routine high-quality EnKF convective-scale analysis one would need -m horizontal resolution, an accurate higher-moment microphysics scheme, and an accurate description of observational errors. In addition, one will have to deal with sporadic irregular dynamics, non-Gaussian error dynamics, and many additional, mostly unobserved, model variables. It is too early to tell if the problem can successfully be addressed in the context of an EnKF.

d. Satellite observations

Global EnKF systems now routinely assimilate a fair amount of radiance observations with some success. Although there are few corresponding studies, it does nevertheless seem that their impact is not as large and critical to good performance as in corresponding variational systems. Possible explanations are the following: suboptimal covariance localization, a difficulty of the EnKF to deal with dense observation sets, imperfect bias removal, and neglected observation-error correlations.

It would be important for the EnKF community to learn how to make optimal use of all available observation types. A systematic analysis of many of the issues relating to the assimilation of radiance observations could be done in the context of an observing system simulation experiment (OSSE).

e. Hybrid systems

In hybrid systems, one tries to combine components of existing EnKF and variational analysis systems to obtain a better coupled system. Similar to the experience with multimodel ensembles like the North American Ensemble Forecast System (NAEFS; Candille 2009), a substantial improvement is generally associated with the combination of two fairly independent systems of similar quality. This has been shown by Bonavita et al. (2015) in the implementation of the gain hybrid of the 4D-Var and LETKF assimilation systems.

A difficulty with a hybrid system is that it requires a sustained investment and development for the two component systems as well as for the combined system. It is not clear that the operational centers can and will support a hybrid system for an extended period of time.

f. Future of the EnKF

It would seem that EnKF systems can continue to benefit from improvements in computational platforms for some time to come. The use of notably higher resolution and more members should fairly easily translate into higher analysis quality. It would seem possible, in the future, to have global EnKF configurations that approach the gray zone (a horizontal resolution of approximately 10 km). In this endeavor, it will likely be of value that a moderately large variety of ensemble-based data assimilation systems, using different parallelization and (due to their fundamental impact on computational efficiency) localization strategies, exists at operational centers.

Improved boundary conditions, including a reasonable estimate of uncertainty, could be provided by coupling of an atmospheric EnKF with similar systems for the ocean, the ice, the land surface, chemical constituents, etc. In principle, it is possible to have a unique data assimilation application that uses all observations to estimate the state vector of all component systems using a single gain matrix (Tardif et al. 2015; Sluka et al. 2016). Challenges for such coupled assimilation systems come from different spatial and temporal scales in the component systems, from possibly weak and, therefore, hard to estimate correlations across boundaries and from the likely complex dynamics of the coupled system.

It may also be possible to alleviate some of the constraints that seem to be imposed by the Kalman filter framework. For instance, Lien et al. (2013, 2016a,b) recently proposed several ways of addressing the non-Gaussian distributions of precipitation, including the use of a Gaussian transformation of variables, and the assimilation of both zero precipitation and nonzero precipitation observations.

In particular because of the huge electricity requirement, it does not appear possible to have an EnKF configuration that functions beyond the gray zone, at a horizontal resolution of 1 km or higher, with current computer technology. For global convection-resolving data assimilation, it may be necessary to use dual-resolution configurations.

Acknowledgments

This paper started as a summary of the material presented and discussed at the Sixth EnKF Workshop that was held 18–22 May 2014 near Buffalo, New York, and we want to thank all workshop participants. We thank Herschel Mitchell and Jonathan Poterjoy for internal reviews with many valuable suggestions. We thank Luc Fillion for, notably, discussions regarding balance and regional data assimilation. We thank the reviewers Geir Evensen and Eugenia Kalnay, as well as an anonymous reviewer, for numerous valuable suggestions that helped us to improve the manuscript. FZ is partially supported by NSF Grant 1305798, Office of Naval Research Grant N000140910526, and NASA Grant NNX12AJ79G.

APPENDIX A

Types of Filter Divergence

In filter divergence, the best estimate provided by the EnKF becomes disconnected from reality. In classical filter divergence, well known from Kalman filter systems, the ensemble spread becomes much too small. An EnKF can also suffer from catastrophic filter divergence in which the ensemble spread becomes very large.

a. Classical filter divergence

It is well known that the Kalman filter can display unstable behavior (Lewis et al. 2006, their section 28.4; Snyder 2015, his section 3.5.4). The following issues can play a role in filter divergence: model bias and errors, errors in specified statistics, round-off errors, and nonlinearity in the system. A symptom of the divergence is a high-condition number of the matrix . A standard approach for dealing with this problem is to reformulate the algorithm using square root matrices (Lewis et al. 2006, their section 28.4) or to add a high-rank system noise (Fitzgerald 1971). Similarly, in modern atmospheric applications of the EnKF, the rank problem that is associated with small ensemble sizes will naturally lead to a high or even infinite condition number (section 3d). Covariance localization and the addition of system error are often used to protect against the associated filter divergence (Whitaker and Hamill 2002, see revised Fig. 3 in the corrigendum to their article). Note, however, that covariance localization, in combination with a sequential algorithm, can also be a cause of filter divergence (section 7b above; Nerger 2015). In environments with stable dynamics—or no dynamics at all as in the case of parameter estimation (Aksoy et al. 2006)—there may also be issues, and it is a good practice to always verify that the estimated variances of the background error remain significant.

b. Catastrophic filter divergence

In the EnKF context, there are what would seem to be new types of instability (Houtekamer and Mitchell 2005, their section 3b; Kelly et al. 2015). The issue here is that the ensemble spread becomes very large. This will lead to large-amplitude analysis increments and, when unchecked, can lead to initial conditions that can no longer be handled by the numerical model.

In the case of observations that are only weakly correlated with the model state, the estimation noise can be larger than the signal and the analysis can be less accurate than the background for some variables. When a cross-validation procedure is used, the EnKF is “aware�? of this situation and the analysis spread will exceed the spread in the background (Houtekamer and Mitchell 2005, their section 3b). This increased spread subsequently enables larger analysis increments when subsequent sets of similar observations are assimilated in the sequential analysis algorithm. When the number of observations is large, this divergence can act very quickly. This instability can be reduced or controlled notably by using larger ensembles and more severe covariance localization.

In the example by Kelly et al. (2015), the observation is perpendicular to the ensemble subspace and the catastrophic blow-up results as an interplay between the numerical model and the forward operator. In this case, additive covariance inflation can stabilize the filter.

Finally, we note once more that filter divergence can also result from the use of multiplicative inflation in data-sparse areas (Anderson 2009) and potentially from other methods to inflate ensemble spread such as the addition of physical tendency perturbations.

Since, as we have seen, both large and small ensemble spread can cause filter divergence, EnKF systems intended for operational applications have to be designed and tested with great care.

APPENDIX B

Systems Available for Download

Most ensemble Kalman filter algorithms are conceptually fairly simply. However, to obtain good performance on a supercomputer with a true forecasting model and a variety of observation types represents a major effort. Starting essentially from scratch, but embedded in an operational environment, Drs. Houtekamer and Mitchell and coworkers needed approximately 10 years to arrive at their first operational implementation (Houtekamer et al. 2005). Similarly, the EnSRF, which had been developed originally at NCAR, has been implemented for operational use at NCEP (Whitaker et al. 2008) and for research use at ECMWF. Operational centers, with no experience in ensemble-based data assimilation, can likely benefit from the experience and code developed at other centers. Here we want to refer to three specific efforts to make ensemble-based data assimilation available to a larger group of users.

The Data Assimilation Research Testbed (DART) is a community facility (Anderson et al. 2009) for ensemble-based data assimilation that is developed and supported by the National Center for Atmospheric Research (NCAR). It includes many models ranging from very simple to fairly complex and has a large community of users. The code is accessible using subversion from a central code repository (http://www.image.ucar.edu/DAReS/DART/). First results in complex environments can reportedly be obtained in a few months.

The source code of a WRF-based EnKF system (Zhang et al. 2009a, 2011; Weng and Zhang 2012), which has been continuously developed at The Pennsylvania State University, and is used for experimental real-time convection-permitting hurricane analysis and prediction, can be freely downloaded online at http://adapt.psu.edu/index.php?loc=outreach.

The source code of the LETKF (Hunt et al. 2007), which has originally been developed at the University of Maryland, has been made available for download by T. Miyoshi at https://code.google.com/p/miyoshi. The LETKF has a large, rather loose, community of users.

The above initiatives will facilitate the work toward obtaining initial results. Beyond that, a substantial amount of experimentation will likely be required to arrive at a robust system with a satisfying quality.

REFERENCES

  • Aksoy, A., F. Zhang, and J. W. Nielsen-Gammon, 2006: Ensemble-based simultaneous state and parameter estimation in a two-dimensional sea-breeze model. Mon. Wea. Rev., 134, 29512970, doi:10.1175/MWR3224.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., D. C. Dowell, and C. Snyder, 2009: A multicase comparative assessment of the ensemble Kalman filter for assimilation of radar observations. Part I: Storm-scale analyses. Mon. Wea. Rev., 137, 18051824, doi:10.1175/2008MWR2691.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., D. C. Dowell, and C. Snyder, 2010: A multicase comparative assessment of the ensemble Kalman filter for assimilation of radar observations. Part II: Short-range ensemble forecasts. Mon. Wea. Rev., 138, 12731292, doi:10.1175/2009MWR3086.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., S. D. Aberson, T. Vukicevic, K. J. Sellwood, S. Lorsolo, and X. Zhang, 2013: Assimilation of high-resolution tropical cyclone observations with an ensemble Kalman filter using NOAA/AOML/HRD’s HEDAS: Evaluation of the 2008–11 vortex-scale analyses. Mon. Wea. Rev., 141, 18421865, doi:10.1175/MWR-D-12-00194.1.

    • Search Google Scholar
    • Export Citation
  • Albers, S. C., J. A. McGinley, D. L. Birkenheuer, and J. R. Smart, 1996: The Local Analysis and Prediction System (LAPS): Analysis of clouds, precipitation, and temperature. Wea. Forecasting, 11, 273287, doi:10.1175/1520-0434(1996)011<0273:TLAAPS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Amezcua, J., K. Ide, E. Kalnay, and S. Reich, 2014: Ensemble transform Kalman–Bucy filters. Quart. J. Roy. Meteor. Soc., 140, 9951004, doi:10.1002/qj.2186.

    • Search Google Scholar
    • Export Citation
  • 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
  • Anderson, J. L., 2003: A local least squares framework for ensemble filtering. Mon. Wea. Rev., 131, 634642, doi:10.1175/1520-0493(2003)131<0634:ALLSFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2007: Exploring the need for localization in ensemble data assimilation using a hierarchical ensemble filter. Physica D, 230, 99111, doi:10.1016/j.physd.2006.02.011.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2009: Spatially and temporally varying adaptive covariance inflation for ensemble filters. Tellus, 61A, 7283, doi:10.1111/j.1600-0870.2008.00361.x.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2012: Localization and sampling error correction in ensemble Kalman filter data assimilation. Mon. Wea. Rev., 140, 23592371, doi:10.1175/MWR-D-11-00013.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2016: Reducing correlation sampling error in ensemble Kalman filter data assimilation. Mon. Wea. Rev., 144, 913925, doi:10.1175/MWR-D-15-0052.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., and S. L. Anderson, 1999: A Monte Carlo implementation of the nonlinear filtering problem to produce ensemble assimilations and forecasts. Mon. Wea. Rev., 127, 27412758, doi:10.1175/1520-0493(1999)127<2741:AMCIOT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., T. Hoar, K. Raeder, H. Liu, N. Collins, R. Torn, and A. Avellano, 2009: The Data Assimilation Research Testbed: A community facility. Bull. Amer. Meteor. Soc., 90, 12831296, doi:10.1175/2009BAMS2618.1.

    • Search Google Scholar
    • Export Citation
  • Annan, J. D., 2004: On the orthogonality of bred vectors. Mon. Wea. Rev., 132, 843849, doi:10.1175/1520-0493(2004)132<0843:OTOOBV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Annan, J. D., J. C. Hargreaves, N. R. Edwards, and R. Marsh, 2005: Parameter estimation in an intermediate complexity earth system model using an ensemble Kalman filter. Ocean Modell., 8, 135154, doi:10.1016/j.ocemod.2003.12.004.

    • Search Google Scholar
    • Export Citation
  • Aravéquia, J. A., I. Szunyogh, E. J. Fertig, E. Kalnay, D. Kuhl, and E. J. Kostelich, 2011: Evaluation of a strategy for the assimilation of satellite radiance observations with the local ensemble transform Kalman filter. Mon. Wea. Rev., 139, 19321951, doi:10.1175/2010MWR3515.1.

    • Search Google Scholar
    • Export Citation
  • Ballish, B., X. Cao, E. Kalnay, and M. Kanamitsu, 1992: Incremental nonlinear normal-mode initialization. Mon. Wea. Rev., 120, 17231734, doi:10.1175/1520-0493(1992)120<1723:INNMI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Barker, D. M., 2005: Southern high-latitude ensemble data assimilation in the Antartic mesoscale prediction system. Mon. Wea. Rev., 133, 34313449, doi:10.1175/MWR3042.1.

    • Search Google Scholar
    • Export Citation
  • Barker, D. M., and Coauthors, 2012: The Weather Research and Forecasting Model’s Community Variational/Ensemble Data Assimilation System: WRFDA. Bull. Amer. Meteor. Soc., 93, 831843, doi:10.1175/BAMS-D-11-00167.1.

    • Search Google Scholar
    • Export Citation
  • Bergemann, K., and S. Reich, 2010: A mollified ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 136, 16361643, doi:10.1002/qj.672.

    • Search Google Scholar
    • Export Citation
  • Berner, J., G. J. Shutts, M. Leutbecher, and T. N. Palmer, 2009: A spectral stochastic kinetic energy backscatter scheme and its impact on flow-dependent predictability in the ECMWF ensemble prediction system. J. Atmos. Sci., 66, 603626, doi:10.1175/2008JAS2677.1.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and Z. Toth, 1999: Ensemble transformation and adaptive observations. J. Atmos. Sci., 56, 17481765, doi:10.1175/1520-0469(1999)056<1748:ETAAO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and D. Hodyss, 2009: Ensemble covariances adaptively localized with ECO-RAP. Part 1: Tests on simple error models. Tellus, 61A, 8496, doi:10.1111/j.1600-0870.2008.00371.x.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., B. J. Etherton, and S. J. Majumdar, 2001: Adaptive sampling with the ensemble transform Kalman filter. Part I: Theoretical aspects. Mon. Wea. Rev., 129, 420436, doi:10.1175/1520-0493(2001)129<0420:ASWTET>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bloom, S. C., L. L. Takacs, A. M. da Silva, and D. Ledvina, 1996: Data assimilation using incremental analysis updates. Mon. Wea. Rev., 124, 12561271, doi:10.1175/1520-0493(1996)124<1256:DAUIAU>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., 2011: Impact and diagnosis of model error in the ECMWF ensemble of data assimilations. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 303318.

  • Bonavita, M., L. Torrisi, and F. Marcucci, 2010: Ensemble data assimilation with the CNMCA regional forecasting system. Quart. J. Roy. Meteor. Soc., 136, 132145, doi:10.1002/qj.553.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., L. Isaksen, and E. Hólm, 2012: On the use of EDA background error variances in the ECMWF 4D-Var. Quart. J. Roy. Meteor. Soc., 138, 15401559, doi:10.1002/qj.1899.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., M. Hamrud, and L. Isaksen, 2015: EnKF and hybrid gain ensemble data assimilation. Part II: EnKF and hybrid gain results. Mon. Wea. Rev., 143, 48654882, doi:10.1175/MWR-D-15-0071.1.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., E. Hólm, L. Isaksen, and M. Fisher, 2016: The evolution of the ECMWF hybrid data assimilation system. Quart. J. Roy. Meteor. Soc., 142, 287303, doi:10.1002/qj.2652.

    • Search Google Scholar
    • Export Citation
  • Bormann, N., and P. Bauer, 2010: Estimates of spatial and interchannel observation-error characteristics for current sounder radiances for numerical weather prediction. I: Methods and application to ATOVS data. Quart. J. Roy. Meteor. Soc., 136, 10361050, doi:10.1002/qj.616.

    • Search Google Scholar
    • Export Citation
  • Bowler, N. E., A. Arribas, K. R. Mylne, K. B. Robertson, and S. E. Beare, 2008: The MOGREPS short-range ensemble prediction system. Quart. J. Roy. Meteor. Soc., 134, 703722, doi:10.1002/qj.234.

    • Search Google Scholar
    • Export Citation
  • Brankart, J.-M., C. Ubelmann, C.-E. Testut, E. Cosme, P. Brasseur, and J. Verron, 2009: Efficient parameterization of the observation error covariance matrix for square root or ensemble Kalman filters: Application to ocean altimetry. Mon. Wea. Rev., 137, 19081927, doi:10.1175/2008MWR2693.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., 2012: Evaluation of a spatial/spectral covariance localization approach for atmospheric data assimilation. Mon. Wea. Rev., 140, 617636, doi:10.1175/MWR-D-10-05052.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., P. L. Houtekamer, C. Charette, H. L. Mitchell, and B. He, 2010a: Intercomparison of variational data assimilation and the ensemble Kalman filter for global deterministic NWP. Part I: Description and single-observation experiments. Mon. Wea. Rev., 138, 15501566, doi:10.1175/2009MWR3157.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., P. L. Houtekamer, C. Charette, H. L. Mitchell, and B. He, 2010b: Intercomparison of variational data assimilation and the ensemble Kalman filter for global deterministic NWP. Part II: One-month experiments with real observations. Mon. Wea. Rev., 138, 15671586, doi:10.1175/2009MWR3158.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., and Coauthors, 2015: Implementation of deterministic weather forecasting systems based on ensemble–variational data assimilation at Environment Canada. Part I: The global system. Mon. Wea. Rev., 143, 25322559, doi:10.1175/MWR-D-14-00354.1.

    • Search Google Scholar
    • Export Citation
  • Buizza, R., M. Miller, and T. N. Palmer, 1999: Stochastic representation of model uncertainties in the ECMWF ensemble prediction system. Quart. J. Roy. Meteor. Soc., 125, 28872908, doi:10.1002/qj.49712556006.

    • Search Google Scholar
    • Export Citation
  • Burgers, G., P. J. van Leeuwen, and G. Evensen, 1998: Analysis scheme in the ensemble Kalman filter. Mon. Wea. Rev., 126, 17191724, doi:10.1175/1520-0493(1998)126<1719:ASITEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Campbell, W. F., C. H. Bishop, and D. Hodyss, 2010: Vertical covariance localization for satellite radiances in ensemble Kalman filters. Mon. Wea. Rev., 138, 282290, doi:10.1175/2009MWR3017.1.

    • Search Google Scholar
    • Export Citation
  • Candille, G., 2009: The multiensemble approach: The NAEFS example. Mon. Wea. Rev., 137, 16551665, doi:10.1175/2008MWR2682.1.

  • Candille, G., and O. Talagrand, 2005: Evaluation of probabilistic prediction systems for a scalar variable. Quart. J. Roy. Meteor. Soc., 131, 21312150, doi:10.1256/qj.04.71.

    • Search Google Scholar
    • Export Citation
  • Candille, G., and O. Talagrand, 2008: Impact of observational error on the validation of ensemble prediction systems. Quart. J. Roy. Meteor. Soc., 134, 959971, doi:10.1002/qj.268.

    • Search Google Scholar
    • Export Citation
  • Cardinali, C., 2009: Monitoring the observation impact on the short-range forecast. Quart. J. Roy. Meteor. Soc., 135, 239250, doi:10.1002/qj.366.

    • Search Google Scholar
    • Export Citation
  • Caron, J.-F., T. Milewski, M. Buehner, L. Fillion, M. Reszka, S. Macpherson, and J. St-James, 2015: Implementation of deterministic weather forecasting systems based on ensemble-variational data assimilation at Environment Canada. Part II: The regional system. Mon. Wea. Rev., 143, 25602580, doi:10.1175/MWR-D-14-00353.1.

    • Search Google Scholar
    • Export Citation
  • Caya, A., J. Sun, and C. Snyder, 2005: A comparison between the 4DVAR and the ensemble Kalman filter techniques for radar data assimilation. Mon. Wea. Rev., 133, 30813094, doi:10.1175/MWR3021.1.

    • Search Google Scholar
    • Export Citation
  • Chang, C.-C., S.-C. Yang, and C. Keppenne, 2014a: Applications of the mean recentering scheme to improve typhoon track prediction: A case study of typhoon Nanmadol (2011). J. Meteor. Soc. Japan, 92, 559584, doi:10.2151/jmsj.2014-604.

    • Search Google Scholar
    • Export Citation
  • Chang, W., K.-S. Chung, L. Fillion, and S.-J. Baek, 2014b: Radar data assimilation in the Canadian high-resolution ensemble Kalman filter system: Performance and verification with real summer cases. Mon. Wea. Rev., 142, 21182138, doi:10.1175/MWR-D-13-00291.1.

    • Search Google Scholar
    • Export Citation
  • Charron, M., G. Pellerin, L. Spacek, P. L. Houtekamer, N. Gagnon, H. L. Mitchell, and L. Michelin, 2010: Toward random sampling of model error in the Canadian Ensemble Prediction System. Mon. Wea. Rev., 138, 18771901, doi:10.1175/2009MWR3187.1.

    • Search Google Scholar
    • Export Citation
  • Clayton, A. M., A. C. Lorenc, and D. M. Barker, 2013: Operational implementation of a hybrid ensemble/4D-Var global data assimilation system at the Met Office. Quart. J. Roy. Meteor. Soc., 139, 14451461, doi:10.1002/qj.2054.

    • Search Google Scholar
    • Export Citation
  • Cohn, S. E., and D. F. Parrish, 1991: The behavior of forecast error covariances for a Kalman filter in two dimensions. Mon. Wea. Rev., 119, 17571785, doi:10.1175/1520-0493(1991)119<1757:TBOFEC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Cohn, S. E., A. da Silva, J. Guo, M. Sienkiewicz, and D. Lamich, 1998: Assessing the effects of data selection with the DAO physical-space statistical analysis system. Mon. Wea. Rev., 126, 29132926, doi:10.1175/1520-0493(1998)126<2913:ATEODS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Compo, G. P., J. S. Whitaker, and P. D. Sardeshmukh, 2006: Feasibility of a 100-year reanalysis using only surface pressure data. Bull. Amer. Meteor. Soc., 87, 175190, doi:10.1175/BAMS-87-2-175.

    • Search Google Scholar
    • Export Citation
  • Daley, R., 1991: Atmospheric Data Analysis. Cambridge University Press, 457 pp.

  • Daley, R., 1995: Estimating the wind field from chemical constituent observations: Experiments with a one-dimensional extended Kalman filter. Mon. Wea. Rev., 123, 181198, doi:10.1175/1520-0493(1995)123<0181:ETWFFC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Daley, R., and T. Mayer, 1986: Estimates of global analysis error from the global weather experiment observational network. Mon. Wea. Rev., 114, 16421653, doi:10.1175/1520-0493(1986)114<1642:EOGAEF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • 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
  • Dee, D. P., 2005: Bias and data assimilation. Quart. J. Roy. Meteor. Soc., 131, 33233343, doi:10.1256/qj.05.137.

  • Derber, J. C., and W.-S. Wu, 1998: The use of TOVS cloud-cleared radiances in the NCEP SSI analysis system. Mon. Wea. Rev., 126, 22872299, doi:10.1175/1520-0493(1998)126<2287:TUOTCC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Derber, J. C., and F. Bouttier, 1999: A reformulation of the background error covariance in the ECMWF global data assimilation system. Tellus, 51A, 195221, doi:10.1034/j.1600-0870.1999.t01-2-00003.x.

    • Search Google Scholar
    • Export Citation
  • Desroziers, G., L. Berre, B. Chapnik, and P. Poli, 2005: Diagnosis of observation, background and analysis-error statistics in observation space. Quart. J. Roy. Meteor. Soc., 131, 33853396, doi:10.1256/qj.05.108.

    • Search Google Scholar
    • Export Citation
  • Dillon, M. E., and Coauthors, 2016: Application of the WRF-LETKF data assimilation system over southern South America: Sensitivity to model physics. Wea. Forecasting, 31, 217236, doi:10.1175/WAF-D-14-00157.1.

    • Search Google Scholar
    • Export Citation
  • Dong, J., and M. Xue, 2013: Assimilation of radial velocity and reflectivity data from coastal WSR-88D radars using an ensemble Kalman filter for the analysis and forecast of landfalling hurricane Ike (2008). Quart. J. Roy. Meteor. Soc., 139, 467487, doi:10.1002/qj.1970.

    • Search Google Scholar
    • Export Citation
  • Dowell, D. C., F. Zhang, L. J. Wicker, C. Snyder, and N. A. Crook, 2004: Wind and temperature retrievals in the 17 May 1981 Arcadia, Oklahoma, supercell: Ensemble Kalman filter experiments. Mon. Wea. Rev., 132, 19822005, doi:10.1175/1520-0493(2004)132<1982:WATRIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dowell, D. C., L. J. Wicker, and C. Snyder, 2011: Ensemble Kalman filter assimilation of radar observations of the 8 May 2003 Oklahoma City supercell: Influences of reflectivity observations on storm-scale analyses. Mon. Wea. Rev., 139, 272294, doi:10.1175/2010MWR3438.1.

    • Search Google Scholar
    • Export Citation
  • ECMWF, 2011: Proceedings of the ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, 20–24 June 2011. ECMWF, 370 pp.

  • English, S. J., R. J. Renshaw, P. C. Dibben, A. J. Smith, P. J. Rayer, C. Poulsen, F. W. Saunders, and J. R. Eyre, 2000: A comparison of the impact of TOVS and ATOVS satellite sounding data on the accuracy of numerical weather forecasts. Quart. J. Roy. Meteor. Soc., 126, 29112931, doi:10.1002/qj.49712656915.

    • Search Google Scholar
    • Export Citation
  • Errico, R. M., R. Yang, N. C. Privé, K.-S. Tai, R. Todling, M. E. Sienkiewicz, and J. Guo, 2013: Development and validation of observing-system simulation experiments at NASA’s Global Modeling and Assimilation Office. Quart. J. Roy. Meteor. Soc., 139, 11621178, doi:10.1002/qj.2027.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 1994: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99, 10 14310 162, doi:10.1029/94JC00572.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2003: The ensemble Kalman filter: Theoretical formulation and practical implementation. Ocean Dyn., 53, 343367, doi:10.1007/s10236-003-0036-9.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2004: Sampling strategies and square root analysis schemes for the EnKF. Ocean Dyn., 54, 539560, doi:10.1007/s10236-004-0099-2.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2009: The ensemble Kalman filter for combined state and parameter estimation. IEEE Control Syst., 29, 83104, doi:10.1109/MCS.2009.932223.

    • Search Google Scholar
    • Export Citation
  • Fertig, E. J., J. Harlim, and B. R. Hunt, 2007: A comparative study of 4D-VAR and a 4D ensemble Kalman filter: Perfect model simulations with Lorenz-96. Tellus, 59A, 96100, doi:10.1111/j.1600-0870.2006.00205.x.

    • Search Google Scholar
    • Export Citation
  • Fillion, L., H. L. Mitchell, H. Ritchie, and A. Staniforth, 1995: The impact of a digital filter finalization technique in a global data assimilation system. Tellus, 47A, 304323, doi:10.1034/j.1600-0870.1995.t01-2-00002.x.

    • Search Google Scholar
    • Export Citation
  • Fisher, M., 2004: Background error covariance modelling. Proc. ECMWF Seminar on Recent Developments in Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 4563.

  • Fisher, M., and H. Auvinen, 2012: Long window 4D-Var. Proc. ECMWF Seminar on Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 189202.

  • Fishman, G. S., 1996: Monte Carlo: Concepts, Algorithms and Applications. Springer, 698 pp.

  • Fitzgerald, R. J., 1971: Divergence of the Kalman filter. IEEE Trans. Automat. Control, 16, 736747, doi:10.1109/TAC.1971.1099836.

  • Flowerdew, J., 2015: Towards a theory of optimal localisation. Tellus, 67A, 25257, doi:10.3402/tellusa.v67.25257.

  • Frehlich, R., 2006: Adaptive data assimilation including the effect of spatial variations in observation error. Quart. J. Roy. Meteor. Soc., 132, 12251257, doi:10.1256/qj.05.146.

    • Search Google Scholar
    • Export Citation
  • Fujita, T., D. J. Stensrud, and D. C. Dowell, 2007: Surface data assimilation using an ensemble Kalman filter approach with initial condition and model physics uncertainties. Mon. Wea. Rev., 135, 18461868, doi:10.1175/MWR3391.1.

    • Search Google Scholar
    • Export Citation
  • Gaspari, G., and S. E. Cohn, 1999: Construction of correlation functions in two and three dimensions. Quart. J. Roy. Meteor. Soc., 125, 723757, doi:10.1002/qj.49712555417.

    • Search Google Scholar
    • Export Citation
  • Gauthier, P., and J.-N. Thépaut, 2001: Impact of the digital filter as a weak constraint in the preoperational 4DVAR assimilation system of Météo-France. Mon. Wea. Rev., 129, 20892102, doi:10.1175/1520-0493(2001)129<2089:IOTDFA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Gauthier, P., M. Tanguay, S. Laroche, S. Pellerin, and J. Morneau, 2007: Extension of 3DVAR to 4DVAR: Implementation of 4DVAR at the Meteorological Service of Canada. Mon. Wea. Rev., 135, 23392354, doi:10.1175/MWR3394.1.

    • Search Google Scholar
    • Export Citation
  • Geer, A. J., and P. Bauer, 2011: Observation errors in all-sky data assimilation. Quart. J. Roy. Meteor. Soc., 137, 20242037, doi:10.1002/qj.830.

    • Search Google Scholar
    • Export Citation
  • Ghil, M., and P. Malanotte-Rizzoli, 1991: Data assimilation in meteorology and oceanography. Advances in Geophysics, Vol. 33, Academic Press, 141–266.

  • Ghil, M., S. Cohn, J. Tavantzis, K. Bube, and E. Isaacson, 1981: Applications of estimation theory to numerical weather prediction. Dynamic Meteorology—Data Assimilation Methods, L. Bengtsson, M. Ghil, and E. Källén, Eds., Springer-Verlag, 139–224.

  • Gilleland, E., D. A. Ahijevych, B. G. Brown, and E. E. Ebert, 2010: Verifying forecasts spatially. Bull. Amer. Meteor. Soc., 91, 13651373, doi:10.1175/2010BAMS2819.1.

    • Search Google Scholar
    • Export Citation
  • Golub, G. H., and C. F. Van Loan, 1996: Matrix Computations. 3rd ed. Johns Hopkins University Press, 694 pp.

  • Gorin, V. E., and M. D. Tsyrulnikov, 2011: Estimation of multivariate observation-error statistics for AMSU-A data. Mon. Wea. Rev., 139, 37653780, doi:10.1175/2011MWR3554.1.

    • Search Google Scholar
    • Export Citation
  • Grell, G. A., and D. Dévényi, 2002: A generalized approach to parameterizing convection combining ensemble and data assimilation techniques. Geophys. Res. Lett., 29, doi:10.1029/2002GL015311.

    • Search Google Scholar
    • Export Citation
  • Greybush, S. J., E. Kalnay, T. Miyoshi, K. Ide, and B. R. Hunt, 2011: Balance and ensemble Kalman filter localization techniques. Mon. Wea. Rev., 139, 511522, doi:10.1175/2010MWR3328.1.

    • Search Google Scholar
    • Export Citation
  • Ha, S., J. Berner, and C. Snyder, 2015: A comparison of model error representations in mesoscale ensemble data assimilation. Mon. Wea. Rev., 143, 38933911, doi:10.1175/MWR-D-14-00395.1.

    • Search Google Scholar
    • Export Citation
  • Hager, G., and G. Wellein, 2011: Introduction to High Performance Computing for Scientists and Engineers. Chapman & Hall/CRC, 330 pp.

  • Hamill, T. M., 2001: Interpretation of rank histograms for verifying ensemble forecasts. Mon. Wea. Rev., 129, 550560, doi:10.1175/1520-0493(2001)129<0550:IORHFV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., 2006: Ensemble-based atmospheric data assimilation. Predictability of Weather and Climate, T. Palmer and R. Hagedorn, Eds., Cambridge University Press, 124–156.

  • Hamill, T. M., and C. Snyder, 2000: A hybrid ensemble Kalman filter-3D variational analysis scheme. Mon. Wea. Rev., 128, 29052919, doi:10.1175/1520-0493(2000)128<2905:AHEKFV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., and J. S. Whitaker, 2011: What constrains spread growth in forecasts initialized from ensemble Kalman filters? Mon. Wea. Rev., 139, 117131, doi:10.1175/2010MWR3246.1.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., J. S. Whitaker, and C. Snyder, 2001: Distance-dependent filtering of background error covariance estimates in an ensemble Kalman filter. Mon. Wea. Rev., 129, 27762790, doi:10.1175/1520-0493(2001)129<2776:DDFOBE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamrud, M., M. Bonavita, and L. Isaksen, 2015: EnKF and hybrid gain ensemble data assimilation. Part I: EnKF implementation. Mon. Wea. Rev., 143, 48474864, doi:10.1175/MWR-D-14-00333.1.

    • Search Google Scholar
    • Export Citation
  • Hersbach, H., 2000: Decomposition of the continuous ranked probability score for ensemble prediction systems. Wea. Forecasting, 15, 559570, doi:10.1175/1520-0434(2000)015<0559:DOTCRP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., 1993: Global and local skill forecasts. Mon. Wea. Rev., 121, 18341846, doi:10.1175/1520-0493(1993)121<1834:GALSF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., 2011: The use of multiple parameterizations in ensembles. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 163173.

  • Houtekamer, P. L., and L. Lefaivre, 1997: Using ensemble forecasts for model validation. Mon. Wea. Rev., 125, 24162426, doi:10.1175/1520-0493(1997)125<2416:UEFFMV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 1998: Data assimilation using an ensemble Kalman filter technique. Mon. Wea. Rev., 126, 796811, doi:10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 2001: A sequential ensemble Kalman filter for atmospheric data assimilation. Mon. Wea. Rev., 129, 123137, doi:10.1175/1520-0493(2001)129<0123:ASEKFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 2005: Ensemble Kalman filtering. Quart. J. Roy. Meteor. Soc., 131, 32693289, doi:10.1256/qj.05.135.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., L. Lefaivre, J. Derome, H. Ritchie, and H. L. Mitchell, 1996: A system simulation approach to ensemble prediction. Mon. Wea. Rev., 124, 12251242, doi:10.1175/1520-0493(1996)124<1225:ASSATE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., H. L. Mitchell, G. Pellerin, M. Buehner, M. Charron, L. Spacek, and B. Hansen, 2005: Atmospheric data assimilation with an ensemble Kalman filter: Results with real observations. Mon. Wea. Rev., 133, 604620, doi:10.1175/MWR-2864.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., H. L. Mitchell, and X. Deng, 2009: Model error representation in an operational ensemble Kalman filter. Mon. Wea. Rev., 137, 21262143, doi:10.1175/2008MWR2737.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., X. Deng, H. L. Mitchell, S.-J. Baek, and N. Gagnon, 2014a: Higher resolution in an operational ensemble Kalman filter. Mon. Wea. Rev., 142, 11431162, doi:10.1175/MWR-D-13-00138.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., B. He, and H. L. Mitchell, 2014b: Parallel implementation of an ensemble Kalman filter. Mon. Wea. Rev., 142, 11631182, doi:10.1175/MWR-D-13-00011.1.

    • Search Google Scholar
    • Export Citation
  • Hu, M., M. Xue, and K. Brewster, 2006: 3DVAR and cloud analysis with WSR-88D level-II data for the prediction of the Fort Worth, Texas, tornadic thunderstorms. Part I: Cloud analysis and its impact. Mon. Wea. Rev., 134, 675698, doi:10.1175/MWR3092.1.

    • Search Google Scholar
    • Export Citation
  • Hu, X.-M., F. Zhang, and J. W. Nielsen-Gammon, 2010: Ensemble-based simultaneous state and parameter estimation for treatment of mesoscale model error: A real-data study. Geophys. Res. Lett., 37, L08802, doi:10.1029/2010GL043017.

    • Search Google Scholar
    • Export Citation
  • Huang, X.-Y., and P. Lynch, 1993: Diabetic digital-filtering initialization: Application to the HIRLAM model. Mon. Wea. Rev., 121, 589603, doi:10.1175/1520-0493(1993)121<0589:DDFIAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hunt, B. R., and Coauthors, 2004: Four-dimensional ensemble Kalman filtering. Tellus, 56A, 273277, doi:10.1111/j.1600-0870.2004.00066.x.

    • Search Google Scholar
    • Export Citation
  • Hunt, B. R., E. J. Kostelich, and I. Szunyogh, 2007: Efficient data assimilation for spatiotemporal chaos: A local ensemble transform Kalman filter. Physica D, 230, 112126, doi:10.1016/j.physd.2006.11.008.

    • Search Google Scholar
    • Export Citation
  • Isaksen, L., 2012: Data assimilation on future computer architectures. Proc. ECMWF Seminar on Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 301322.

  • Jones, R. H., 1965: Optimal estimation of initial conditions for numerical prediction. J. Atmos. Sci., 22, 658663, doi:10.1175/1520-0469(1965)022<0658:OEOICF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jung, Y., M. Xue, and G. Zhang, 2010a: Simulations of polarimetric radar signatures of a supercell storm using a two-moment bulk microphysics scheme. J. Appl. Meteor. Climatol., 49, 146163, doi:10.1175/2009JAMC2178.1.

    • Search Google Scholar
    • Export Citation
  • Jung, Y., M. Xue, and G. Zhang, 2010b: Simultaneous estimation of microphysical parameters and the atmospheric state using simulated polarimetric radar data and an ensemble Kalman filter in the presence of an observation operator error. Mon. Wea. Rev., 138, 539562, doi:10.1175/2009MWR2748.1.

    • Search Google Scholar
    • Export Citation
  • Kalman, R. E., 1960: A new approach to linear filtering and prediction problems. J. Basic Eng., 82, 3545, doi:10.1115/1.3662552.

  • 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., and A. Dalcher, 1987: Forecasting forecast skill. Mon. Wea. Rev., 115, 349356, doi:10.1175/1520-0493(1987)115<0349:FFS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., and S.-C. Yang, 2010: Accelerating the spin-up of ensemble Kalman filtering. Quart. J. Roy. Meteor. Soc., 136, 16441651, doi:10.1002/qj.652.

    • Search Google Scholar
    • Export Citation
  • Kang, J.-S., E. Kalnay, J. Liu, I. Fung, T. Miyoshi, and K. Ide, 2011: “Variable localization�? in an ensemble Kalman filter: Application to the carbon cycle data assimilation. J. Geophys. Res., 116, D09110, doi:10.1029/2010JD014673.

    • Search Google Scholar
    • Export Citation
  • Kang, J.-S., E. Kalnay, T. Miyoshi, J. Liu, and I. Fung, 2012: Estimation of surface carbon fluxes with an advanced data assimilation methodology. J. Geophys. Res., 117, D24101, doi:10.1029/2012JD018259.

    • Search Google Scholar
    • Export Citation
  • Kelly, D., A. J. Majda, and X. T. Tong, 2015: Concrete ensemble Kalman filters with rigorous catastrophic filter divergence. Proc. Natl. Acad. Sci. USA, 112, 10 58910 594, doi:10.1073/pnas.1511063112.

    • Search Google Scholar
    • Export Citation
  • Kepert, J. D., 2009: Covariance localisation and balance in an ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 135, 11571176, doi:10.1002/qj.443.

    • Search Google Scholar
    • Export Citation
  • Kleist, D. T., and K. Ide, 2015a: An OSSE-based evaluation of hybrid variational-ensemble data assimilation for the NCEP GFS. Part I: System description and 3D-hybrid results. Mon. Wea. Rev., 143, 433451, doi:10.1175/MWR-D-13-00351.1.

    • Search Google Scholar
    • Export Citation
  • Kleist, D. T., and K. Ide, 2015b: An OSSE-based evaluation of hybrid variational-ensemble data assimilation for the NCEP GFS. Part II: 4DEnVar and hybrid variants. Mon. Wea. Rev., 143, 452470, doi:10.1175/MWR-D-13-00350.1.

    • Search Google Scholar
    • Export Citation
  • Krishnamurti, T. N., H. S. Bedi, W. Heckley, and K. Ingles, 1988: Reduction of the spinup time for evaporation and precipitation in a spectral model. Mon. Wea. Rev., 116, 907920, doi:10.1175/1520-0493(1988)116<0907:ROTSTF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Krishnamurti, T. N., C. M. Kishtawal, T. E. LaRow, D. R. Bachiochi, Z. Zhang, C. E. Williford, S. Gadgil, and S. Surendran, 1999: Improved weather and seasonal climate forecasts from multimodel superensemble. Science, 285, 15481550, doi:10.1126/science.285.5433.1548.

    • Search Google Scholar
    • Export Citation
  • Kumjian, M. R., 2013: Principles and applications of dual-polarization weather radar. Part III: Artifacts. J. Oper. Meteor., 1, 265274, doi:10.15191/nwajom.2013.0121.

    • Search Google Scholar
    • Export Citation
  • Kunii, M., 2014: The 1000-member ensemble Kalman filtering with the JMA nonhydrostatic mesoscale model on the K computer. J. Meteor. Soc. Japan, 92, 623633, doi:10.2151/jmsj.2014-607.

    • Search Google Scholar
    • Export Citation
  • Lang, S. T. K., M. Bonavita, and M. Leutbecher, 2015: On the impact of re-centering initial conditions for ensemble forecasts. Quart. J. Roy. Meteor. Soc., 141, 25712581, doi:10.1002/qj.2543.

    • Search Google Scholar
    • Export Citation
  • Lange, H., and G. C. Craig, 2014: The impact of data assimilation length scales on analysis and prediction of convective storms. Mon. Wea. Rev., 142, 37813808, doi:10.1175/MWR-D-13-00304.1.

    • Search Google Scholar
    • Export Citation
  • Lavaysse, C., M. Carrera, S. Bélair, N. Gagnon, R. Frenette, M. Charron, and M. K. Yau, 2013: Impact of surface parameter uncertainties within the Canadian regional ensemble prediction system. Mon. Wea. Rev., 141, 15061526, doi:10.1175/MWR-D-11-00354.1.

    • Search Google Scholar
    • Export Citation
  • Lawson, W. G., and J. A. Hansen, 2004: Implications of stochastic and deterministic filters as ensemble-based data assimilation methods in varying regimes of error growth. Mon. Wea. Rev., 132, 19661981, doi:10.1175/1520-0493(2004)132<1966:IOSADF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. L. Anderson, 2014: Empirical localization of observations for serial ensemble Kalman filter data assimilation in an atmospheric general circulation model. Mon. Wea. Rev., 142, 18351851, doi:10.1175/MWR-D-13-00288.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. S. Whitaker, 2015: Model space localization is not always better than observation space localization for assimilation of satellite radiances. Mon. Wea. Rev., 143, 39483955, doi:10.1175/MWR-D-14-00413.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., D. R. Stauffer, and A. Deng, 2012: A hybrid nudging-ensemble Kalman filter approach to data assimilation in WRF/DART. Quart. J. Roy. Meteor. Soc., 138, 20662078, doi:10.1002/qj.1939.

    • Search Google Scholar
    • Export Citation
  • Leith, C. E., 1974: Theoretical skill of Monte Carlo forecasts. Mon. Wea. Rev., 102, 409418, doi:10.1175/1520-0493(1974)102<0409:TSOMCF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lewis, J. M., S. Lakshmivarahan, and S. K. Dhall, 2006: Dynamic Data Assimilation: A Least Squares Approach. Cambridge University Press, 654 pp.

  • Li, X., M. Charron, L. Spacek, and G. Candille, 2008: A regional ensemble prediction system based on moist targeted singular vectors and stochastic parameter perturbations. Mon. Wea. Rev., 136, 443462, doi:10.1175/2007MWR2109.1.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., E. Kalnay, and T. Miyoshi, 2013: Effective assimilation of global precipitation: Simulation experiments. Tellus, 65A, 19915, doi:10.3402/tellusa.v65i0.19915.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., E. Kalnay, T. Miyoshi, and G. J. Huffman, 2016a: Statistical properties of global precipitation in the NCEP GFS model and TMPA observations for data assimilation. Mon. Wea. Rev., 144, 663679, doi:10.1175/MWR-D-15-0150.1.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., T. Miyoshi, and E. Kalnay, 2016b: Assimilation of TRMM multisatellite precipitation analysis with a low-resolution NCEP global forecast system. Mon. Wea. Rev., 144, 643661, doi:10.1175/MWR-D-15-0149.1.

    • Search Google Scholar
    • Export Citation
  • Lindskog, M., K. Salonen, H. Järvinen, and D. B. Michelson, 2004: Doppler radar wind data assimilation with HIRLAM 3DVAR. Mon. Wea. Rev., 132, 10811092, doi:10.1175/1520-0493(2004)132<1081:DRWDAW>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Liu, C., Q. Xiao, and B. Wang, 2008: An ensemble-based four-dimensional variational data assimilation scheme. Part I: Technical formulation and preliminary test. Mon. Wea. Rev., 136, 33633373, doi:10.1175/2008MWR2312.1.

    • Search Google Scholar
    • Export Citation
  • Liu, Z.-Q., and F. Rabier, 2003: The potential of high-density observations for numerical weather prediction: A study with simulated observations. Quart. J. Roy. Meteor. Soc., 129, 30133035, doi:10.1256/qj.02.170.

    • Search Google Scholar
    • Export Citation
  • Liu, Z.-Q., C. S. Schwartz, C. Snyder, and S.-Y. Ha, 2012: Impact of assimilating AMSU-A radiances on forecasts of 2008 Atlantic tropical cyclones initialized with a limited-area ensemble Kalman filter. Mon. Wea. Rev., 140, 40174034, doi:10.1175/MWR-D-12-00083.1.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., 1986: Analysis methods for numerical weather prediction. Quart. J. Roy. Meteor. Soc., 112, 11771194, doi:10.1002/qj.49711247414.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., 2003: The potential of the ensemble Kalman filter for NWP—A comparison with 4D-Var. Quart. J. Roy. Meteor. Soc., 129, 31833203, doi:10.1256/qj.02.132.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., N. E. Bowler, A. M. Clayton, S. R. Pring, and D. Fairbairn, 2015: Comparison of hybrid-4DEnVar and hybrid-4DVar data assimilation methods for global NWP. Mon. Wea. Rev., 143, 212229, doi:10.1175/MWR-D-14-00195.1.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 1965: A study of the predictability of a 28-variable atmospheric model. Tellus, 17A, 321333, doi:10.1111/j.2153-3490.1965.tb01424.x.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 1982: Atmospheric predictability experiments with a large numerical model. Tellus, 34A, 505513, doi:10.1111/j.2153-3490.1982.tb01839.x.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 2005: Designing chaotic models. J. Atmos. Sci., 62, 15741587, doi:10.1175/JAS3430.1.

  • Lynch, P., and X.-Y. Huang, 1992: Initialization of the HIRLAM model using a digital filter. Mon. Wea. Rev., 120, 10191034, doi:10.1175/1520-0493(1992)120<1019:IOTHMU>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Machenhauer, B., 1977: On the dynamics of gravity oscillations in a shallow water model, with application to normal mode initialization. Contrib. Atmos. Phys., 50, 253271.

    • Search Google Scholar
    • Export Citation
  • Mandel, J., 2006: Efficient implementation of the ensemble Kalman filter. Center for Computational Mathematics Rep. 231, University of Colorado at Denver and Health Sciences Center, Denver, CO, 9 pp.

  • Mass, C. F., D. Ovens, K. Westrick, and B. A. Colle, 2002: Does increasing horizontal resolution produce more skillful forecasts? Bull. Amer. Meteor. Soc., 83, 407430, doi:10.1175/1520-0477(2002)083<0407:DIHRPM>2.3.CO;2.

    • Search Google Scholar
    • Export Citation
  • McNally, T., 2004: The assimilation of stratospheric satellite data at ECMWF. Proc. ECMWF/SPARC Workshop on Modelling and Assimilation for the Stratosphere and Tropopause, Shinfield Park, Reading, United Kingdom, ECMWF, 103106.

  • McTaggart-Cowan, R., C. Girard, A. Plante, and M. Desgagné, 2011: The utility of upper-boundary nesting in NWP. Mon. Wea. Rev., 139, 21172144, doi:10.1175/2010MWR3633.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2007: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part II: Imperfect model experiments. Mon. Wea. Rev., 135, 14031423, doi:10.1175/MWR3352.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2008a: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part III: Comparison with 3DVAR in a real-data case study. Mon. Wea. Rev., 136, 522540, doi:10.1175/2007MWR2106.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2008b: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part IV: Comparison with 3DVAR in a month-long experiment. Mon. Wea. Rev., 136, 36713682, doi:10.1175/2008MWR2270.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2011: Limited-area ensemble-based data assimilation. Mon. Wea. Rev., 139, 20252045, doi:10.1175/2011MWR3418.1.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., and P. L. Houtekamer, 2000: An adaptive ensemble Kalman filter. Mon. Wea. Rev., 128, 416433, doi:10.1175/1520-0493(2000)128<0416:AAEKF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., and P. L. Houtekamer, 2009: Ensemble Kalman filter configurations and their performance with the logistic map. Mon. Wea. Rev., 137, 43254343, doi:10.1175/2009MWR2823.1.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., P. L. Houtekamer, and G. Pellerin, 2002: Ensemble size, balance, and model-error representation in an ensemble Kalman filter. Mon. Wea. Rev., 130, 27912808, doi:10.1175/1520-0493(2002)130<2791:ESBAME>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., 2011: The Gaussian approach to adaptive covariance inflation and its implementation with the local ensemble transform Kalman filter. Mon. Wea. Rev., 139, 15191535, doi:10.1175/2010MWR3570.1.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., and K. Kondo, 2013: A multi-scale localization approach to an ensemble Kalman filter. SOLA, 9, 170173, doi:10.2151/sola.2013-038.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., Y. Sato, and T. Kadowaki, 2010: Ensemble Kalman filter and 4D-Var intercomparison with the Japanese operational global analysis and prediction system. Mon. Wea. Rev., 138, 28462866, doi:10.1175/2010MWR3209.1.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., E. Kalnay, and H. Li, 2013: Estimating and including observation-error correlations in data assimilation. Inverse Probl. Sci. Eng., 21, 387398, doi:10.1080/17415977.2012.712527.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., K. Kondo, and T. Imamura, 2014: The 10,240-member ensemble Kalman filtering with an intermediate AGCM. Geophys. Res. Lett., 41, 52645271, doi:10.1002/2014GL060863.

    • Search Google Scholar
    • Export Citation
  • Molteni, F., 2003: Atmospheric simulations using a GCM with simplified physical parametrizations. I: Model climatology and variability in multi-decadal experiments. Climate Dyn., 20, 175191.

    • Search Google Scholar
    • Export Citation
  • Morrison, H., A. Morales, and C. Villanueva-Birriel, 2015: Concurrent sensitivities of an idealized deep convective storm to parameterization of microphysics, horizontal grid resolution, and environmental static stability. Mon. Wea. Rev., 143, 20822104, doi:10.1175/MWR-D-14-00271.1.

    • Search Google Scholar
    • Export Citation
  • Murphy, J., and Coauthors, 2011: Perturbed parameter ensembles as a tool for sampling model uncertainties and making climate projections. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 183208.

  • Mylne, K. R., R. E. Evans, and R. T. Clark, 2002: Multi-model multi-analysis ensembles in quasi-operational medium-range forecasting. Quart. J. Roy. Meteor. Soc., 128, 361384, doi:10.1256/00359000260498923.

    • Search Google Scholar
    • Export Citation
  • Nerger, L., 2015: On serial observation processing in localized ensemble Kalman filters. Mon. Wea. Rev., 143, 15541567, doi:10.1175/MWR-D-14-00182.1.

    • Search Google Scholar
    • Export Citation
  • Newton, C. W., 1954: Analysis and data problems in relation to numerical prediction. Bull. Amer. Meteor. Soc., 35, 287294.

  • Oczkowski, M., I. Szunyogh, and D. J. Patil, 2005: Mechanisms for the development of locally low-dimensional atmospheric dynamics. J. Atmos. Sci., 62, 11351156, doi:10.1175/JAS3403.1.

    • Search Google Scholar
    • Export Citation
  • Ott, E., and Coauthors, 2004: A local ensemble Kalman filter for atmospheric data assimilation. Tellus, 56A, 415428, doi:10.1111/j.1600-0870.2004.00076.x.

    • Search Google Scholar
    • Export Citation
  • Palmer, T. N., 2012: Towards the probabilistic earth-system simulator: A vision for the future of climate and weather prediction. Quart. J. Roy. Meteor. Soc., 138, 841861, doi:10.1002/qj.1923.

    • Search Google Scholar
    • Export Citation
  • Pan, Y., K. Zhu, M. Xue, X. Wang, M. Hu, S. G. Benjamin, S. S. Weygandt, and J. S. Whitaker, 2014: A GSI-based coupled EnSRF-En3DVar hybrid data assimilation system for the operational rapid refresh model: Tests at a reduced resolution. Mon. Wea. Rev., 142, 37563780, doi:10.1175/MWR-D-13-00242.1.

    • Search Google Scholar
    • Export Citation
  • Parrish, D. F., and J. C. Derber, 1992: The National Meteorological Center’s spectral statistical-interpolation analysis system. Mon. Wea. Rev., 120, 17471763, doi:10.1175/1520-0493(1992)120<1747:TNMCSS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Pennell, C., and T. Reichler, 2011: On the effective number of climate models. J. Climate, 24, 23582367, doi:10.1175/2010JCLI3814.1.

  • Penny, S. G., 2014: The hybrid local ensemble transform Kalman filter. Mon. Wea. Rev., 142, 21392149, doi:10.1175/MWR-D-13-00131.1.

  • Penny, S. G., D. W. Behringer, J. A. Carton, and E. Kalnay, 2015: A hybrid global ocean data assimilation system at NCEP. Mon. Wea. Rev., 143, 46604677, doi:10.1175/MWR-D-14-00376.1.

    • Search Google Scholar
    • Export Citation
  • Petersen, D. P., 1968: On the concept and implementation of sequential analysis for linear random fields. Tellus, 20A, 673686, doi:10.1111/j.2153-3490.1968.tb00410.x.

    • Search Google Scholar
    • Export Citation
  • Pires, C., R. Vautard, and O. Talagrand, 1996: On extending the limits of variational assimilation in nonlinear chaotic systems. Tellus, 48A, 96121, doi:10.1034/j.1600-0870.1996.00006.x.

    • Search Google Scholar
    • Export Citation
  • Plant, R. S., and G. C. Craig, 2008: A stochastic parameterization for deep convection based on equilibrium statistics. J. Atmos. Sci., 65, 87105, doi:10.1175/2007JAS2263.1.

    • Search Google Scholar
    • Export Citation
  • Polavarapu, S., M. Tanguay, and L. Fillion, 2000: Four-dimensional variational data assimilation with digital filter initialization. Mon. Wea. Rev., 128, 24912510, doi:10.1175/1520-0493(2000)128<2491:FDVDAW>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Polavarapu, S., S. Ren, A. M. Clayton, D. Sankey, and Y. Rochon, 2004: On the relationship between incremental analysis updating and incremental digital filtering. Mon. Wea. Rev., 132, 24952502, doi:10.1175/1520-0493(2004)132<2495:OTRBIA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2014: Intercomparison and coupling of ensemble and four-dimensional variational data assimilation methods for the analysis and forecasting of Hurricane Karl (2010). Mon. Wea. Rev., 142, 33473364, doi:10.1175/MWR-D-13-00394.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2015: Systematic comparison of four-dimensional data assimilation methods with and without the tangent linear model using hybrid background error covariance: E4DVar versus 4DEnVar. Mon. Wea. Rev., 143, 16011621, doi:10.1175/MWR-D-14-00224.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2016: Comparison of hybrid four-dimensional data assimilation methods with and without the tangent linear and adjoint models for predicting the life cycle of Hurricane Karl (2010). Mon. Wea. Rev., 144, 14491468, doi:10.1175/MWR-D-15-0116.1.

    • Search Google Scholar
    • Export Citation
  • Potter, J. E., 1964: W matrix augmentation. MIT Instrumentation Laboratory, Memo. SGA 5-64, Massachusetts Institute of Technology, Cambridge, MA.

  • Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, 1992: Numerical Recipes in FORTRAN: The Art of Scientific Computing. 2nd ed. Cambridge University Press, 933 pp.

  • Putnam, B. J., M. Xue, Y. Jung, N. Snook, and G. Zhang, 2014: The analysis and prediction of microphysical states and polarimetric radar variables in a mesoscale convective system using double-moment microphysics, multinetwork radar data, and the ensemble Kalman filter. Mon. Wea. Rev., 142, 141162, doi:10.1175/MWR-D-13-00042.1.

    • Search Google Scholar
    • Export Citation
  • Rabier, F., H. Järvinen, E. Klinker, J.-F. Mahfouf, and A. Simmons, 2000: The ECMWF operational implementation of four-dimensional variational assimilation. I: Experimental results with simplified physics. Quart. J. Roy. Meteor. Soc., 126, 11431170, doi:10.1002/qj.49712656415.

    • Search Google Scholar
    • Export Citation
  • Rodgers, C. D., 2000: Inverse Methods for Atmospheric Sounding. World Scientific, 240 pp.

  • Ruiz, J., and M. Pulido, 2015: Parameter estimation using ensemble-based data assimilation in the presence of model error. Mon. Wea. Rev., 143, 15681582, doi:10.1175/MWR-D-14-00017.1.

    • Search Google Scholar
    • Export Citation
  • Rutherford, I. D., 1976: An operational 3-dimensional multivariate statistical objective analysis scheme. Proc. JOC Study Group Conf. on Four-Dimensional Data Assimilation, Rep. 11, Paris, France, WMO/JCSU, 98121.

  • Ryzhkov, A. V., 2007: The impact of beam broadening on the quality of radar polarimetric data. J. Atmos. Oceanic Technol., 24, 729744, doi:10.1175/JTECH2003.1.

    • Search Google Scholar
    • Export Citation
  • Sakov, P., and L. Bertino, 2011: Relation between two common localisation methods for the EnKF. Comput. Geosci., 15, 225237, doi:10.1007/s10596-010-9202-6.

    • Search Google Scholar
    • Export Citation
  • Schraff, C., H. Reich, A. Rhodin, A. Schomburg, K. Stephan, A. Periáñez, and R. Potthast, 2016: Kilometre-scale ensemble data assimilation for the COSMO model (KENDA). Quart. J. Roy. Meteor. Soc., 142, 14531472, doi:10.1002/qj.2748.

    • Search Google Scholar
    • Export Citation
  • Schwartz, C. S., G. S. Romine, R. A. Sobash, K. R. Fossell, and M. L. Weisman, 2015: NCAR’s experimental real-time convection-allowing ensemble prediction system. Wea. Forecasting, 30, 16451654, doi:10.1175/WAF-D-15-0103.1.

    • Search Google Scholar
    • Export Citation
  • Shutts, G., 2005: A kinetic energy backscatter algorithm for use in ensemble prediction systems. Quart. J. Roy. Meteor. Soc., 131, 30793102, doi:10.1256/qj.04.106.

    • Search Google Scholar
    • Export Citation
  • Shutts, G., and A. Callado Pallares, 2011: Tracking down the origin of NWP model uncertainty: coarse-graining studies. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 221231.

  • Simmons, A. J., 1996: The skill of 500 hPa height forecasts. Proc. ECMWF Seminar on Predictability, Shinfield Park, Reading, United Kingdom, ECMWF, 1968.

  • Sippel, J. A., F. Zhang, Y. Weng, L. Tian, G. M. Heymsfield, and S. A. Braun, 2014: Ensemble Kalman filter assimilation of HIWRAP observations of Hurricane Karl (2010) from the unmanned Global Hawk aircraft. Mon. Wea. Rev., 142, 45594580, doi:10.1175/MWR-D-14-00042.1.

    • Search Google Scholar
    • Export Citation
  • Sluka, T. C., S. G. Penny, E. Kalnay, and T. Miyoshi, 2016: Assimilating atmospheric observations into the ocean using strongly coupled ensemble data assimilation. Geophys. Res. Lett., 43, 752759, doi:10.1002/2015GL067238.

    • Search Google Scholar
    • Export Citation
  • Snook, N., M. Xue, and Y. Jung, 2011: Analysis of a tornadic mesoscale convective vortex based on ensemble Kalman filter assimilation of CASA X-band and WSR-88D radar data. Mon. Wea. Rev., 139, 34463468, doi:10.1175/MWR-D-10-05053.1.

    • Search Google Scholar
    • Export Citation
  • Snook, N., M. Xue, and Y. Jung, 2015: Multiscale EnKF assimilation of radar and conventional observations and ensemble forecasting for a tornadic mesoscale convective system. Mon. Wea. Rev., 143, 10351057, doi:10.1175/MWR-D-13-00262.1.

    • Search Google Scholar
    • Export Citation
  • Snyder, C., 2015: Introduction to the Kalman filter. Les Houches 2012: Advanced Data Assimilation for Geosciences, E. Blayo et al., Eds., Oxford University Press, 75–120.

  • Snyder, C., and F. Zhang, 2003: Assimilation of simulated Doppler radar observations with an ensemble Kalman filter. Mon. Wea. Rev., 131, 16631677, doi:10.1175//2555.1.

    • Search Google Scholar
    • Export Citation
  • Steinhoff, J., and D. Underhill, 1994: Modification of the Euler equations for “vorticity confinement�?: Application to the computation of interacting vortex rings. Phys. Fluids A Fluid Dyn., 6, 27382744, doi:10.1063/1.868164.

    • Search Google Scholar
    • Export Citation
  • Stewart, L. M., S. L. Dance, and N. K. Nichols, 2013: Data assimilation with correlated observation errors: Experiments with a 1-D shallow water model. Tellus, 65A, 19546, doi:10.3402/tellusa.v65i0.19546.

    • Search Google Scholar
    • Export Citation
  • Strohmaier, E., J. Dongarra, H. Simon, and M. Meuer, 2015: The Top500 list: Twenty years of insight into HPC performance. Accessed December 2015. [Available online at www.top500.org/lists/2015/11.]

  • Swinbank, R., and Coauthors, 2015: The TIGGE project and its achievements. Bull. Amer. Meteor. Soc., 97, 4967, doi:10.1175/BAMS-D-13-00191.1.

    • Search Google Scholar
    • Export Citation
  • Szunyogh, I., E. J. Kostelich, G. Gyarmati, E. Kalnay, B. R. Hunt, E. Ott, E. Satterfield, and J. A. Yorke, 2008: A local ensemble transform Kalman filter data assimilation system for the NCEP global model. Tellus, 60A, 113130, doi:10.1111/j.1600-0870.2007.00274.x.

    • Search Google Scholar
    • Export Citation
  • Tarantola, A., 1987: Inverse Problem Theory. Elsevier, 613 pp.

  • Tarantola, A., 2006: Popper, Bayes and the inverse problem. Nature Phys., 2, 492494, doi:10.1038/nphys375.

  • Tardif, R., G. J. Hakim, and C. Snyder, 2015: Coupled atmosphere–ocean data assimilation experiments with a low-order model and CMIP5 model data. Climate Dyn., 45, 14151427, doi:10.1007/s00382-014-2390-3.

    • Search Google Scholar
    • Export Citation
  • Temperton, C., and M. Roch, 1991: Implicit normal mode initialization for an operational regional model. Mon. Wea. Rev., 119, 667677, doi:10.1175/1520-0493(1991)119<0667:INMIFA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Thépaut, J.-N., and P. Courtier, 1991: Four-dimensional variational data assimilation using the adjoint of a multilevel primitive-equation model. Quart. J. Roy. Meteor. Soc., 117, 12251254, doi:10.1002/qj.49711750206.

    • Search Google Scholar
    • Export Citation
  • Thomas, S. J., J. P. Hacker, and J. L. Anderson, 2009: A robust formulation of the ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 135, 507521, doi:10.1002/qj.372.

    • Search Google Scholar
    • Export Citation
  • Thompson, T. E., L. J. Wicker, X. Wang, and C. Potvin, 2015: A comparison between the local ensemble transform Kalman filter and the ensemble square root filter for the assimilation of radar data in convective-scale models. Quart. J. Roy. Meteor. Soc., 141, 11631176, doi:10.1002/qj.2423.

    • Search Google Scholar
    • Export Citation
  • Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker, 2003: Ensemble square root filters. Mon. Wea. Rev., 131, 14851490, doi:10.1175/1520-0493(2003)131<1485:ESRF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Tong, M., and M. Xue, 2005: Ensemble Kalman filter assimilation of Doppler radar data with a compressible nonhydrostatic model: OSS experiments. Mon. Wea. Rev., 133, 17891807, doi:10.1175/MWR2898.1.

    • Search Google Scholar
    • Export Citation
  • Torn, R. D., and G. J. Hakim, 2008: Performance characteristics of a pseudo-operational ensemble Kalman filter. Mon. Wea. Rev., 136, 39473963, doi:10.1175/2008MWR2443.1.

    • Search Google Scholar
    • Export Citation
  • Torn, R. D., G. J. Hakim, and C. Snyder, 2006: Boundary conditions for limited-area ensemble Kalman filters. Mon. Wea. Rev., 134, 24902502, doi:10.1175/MWR3187.1.

    • Search Google Scholar
    • Export Citation
  • Toth, Z., and E. Kalnay, 1997: Ensemble forecasting at NCEP and the breeding method. Mon. Wea. Rev., 125, 32973319, doi:10.1175/1520-0493(1997)125<3297:EFANAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Verlaan, M., and A. W. Heemink, 2001: Nonlinearity in data assimilation applications: A practical method for analysis. Mon. Wea. Rev., 129, 15781589, doi:10.1175/1520-0493(2001)129<1578:NIDAAA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Wang, S., M. Xue, A. Schenkman, and J. Min, 2013: An iterative ensemble square root filter and tests with simulated radar data for storm-scale data assimilation. Quart. J. Roy. Meteor. Soc., 139, 18881903, doi:10.1002/qj.2077.

    • Search Google Scholar
    • Export Citation
  • Wang, X., and T. Lei, 2014: GSI-based four-dimensional ensemble-variational (4DEnsVar) data assimilation: Formulation and single- resolution experiments with real data for NCEP global forecast system. Mon. Wea. Rev., 142, 33033325, doi:10.1175/MWR-D-13-00303.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., C. Snyder, and T. M. Hamill, 2007: On the theoretical equivalence of differently proposed ensemble-3DVAR hybrid analysis schemes. Mon. Wea. Rev., 135, 222227, doi:10.1175/MWR3282.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., D. M. Barker, C. Snyder, and T. M. Hamill, 2008: A hybrid ETKF–3DVAR data assimilation scheme for the WRF model. Part I: Observing system simulation experiment. Mon. Wea. Rev., 136, 51165131, doi:10.1175/2008MWR2444.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., D. Parrish, D. Kleist, and J. Whitaker, 2013: GSI 3DVar-based ensemble-variational hybrid data assimilation for NCEP global forecast system: Single-resolution experiments. Mon. Wea. Rev., 141, 40984117, doi:10.1175/MWR-D-12-00141.1.

    • Search Google Scholar
    • Export Citation
  • Wei, M., and Z. Toth, 2003: A new measure of ensemble performance: Perturbation versus error correlation analysis (PECA). Mon. Wea. Rev., 131, 15491565, doi:10.1175//1520-0493(2003)131<1549:ANMOEP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Weng, Y., and F. Zhang, 2012: Assimilating airborne Doppler radar observations with an ensemble Kalman filter for convection-permitting hurricane initialization and prediction: Katrina (2005). Mon. Wea. Rev., 140, 841859, doi:10.1175/2011MWR3602.1.

    • Search Google Scholar
    • Export Citation
  • Weng, Y., and F. Zhang, 2016: Advances in convection-permitting tropical cyclone analysis and prediction through EnKF assimilation of reconnaissance aircraft observations. J. Meteor. Soc. Japan, 94, 345358, doi:10.2151/jmsj.2016-018.

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and T. M. Hamill, 2002: Ensemble data assimilation without perturbed observations. Mon. Wea. Rev., 130, 19131924, doi:10.1175/1520-0493(2002)130<1913:EDAWPO>2.0.CO;2; Corrigendum, 134, 1722, doi:10.1175/MWR3156.1

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and T. M. Hamill, 2012: Evaluating methods to account for system errors in ensemble data assimilation. Mon. Wea. Rev., 140, 30783089, doi:10.1175/MWR-D-11-00276.1.

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., T. M. Hamill, X. Wei, Y. Song, and Z. Toth, 2008: Ensemble data assimilation with the NCEP global forecast system. Mon. Wea. Rev., 136, 463482, doi:10.1175/2007MWR2018.1.

    • Search Google Scholar
    • Export Citation
  • Xue, M., M. Tong, and K. K. Droegemeier, 2006: An OSSE framework based on the ensemble square root Kalman filter for evaluating the impact of data from radar networks on thunderstorm analysis and forecasting. J. Atmos. Oceanic Technol., 23, 4666, doi:10.1175/JTECH1835.1.

    • Search Google Scholar
    • Export Citation
  • Xue, M., J. Schleif, F. Kong, K. W. Thomas, Y. Wang, and K. Zhu, 2013: Track and intensity forecasting of hurricanes: Impact of convection-permitting resolution and global ensemble Kalman filter analysis on 2010 Atlantic season forecasts. Wea. Forecasting, 28, 13661384, doi:10.1175/WAF-D-12-00063.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, B. Hunt, and N. E. Bowler, 2009: Weight interpolation for efficient data assimilation with the Local Ensemble Transform Kalman Filter. Quart. J. Roy. Meteor. Soc., 135, 251262, doi:10.1002/qj.353.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and B. Hunt, 2012a: Handling nonlinearity in an ensemble Kalman filter: Experiments with the three-variable Lorenz model. Mon. Wea. Rev., 140, 26282646, doi:10.1175/MWR-D-11-00313.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and T. Miyoshi, 2012b: Accelerating the EnKF spinup for typhoon assimilation and prediction. Wea. Forecasting, 27, 878897, doi:10.1175/WAF-D-11-00153.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and T. Enomoto, 2015: Ensemble singular vectors and their use as additive inflation in EnKF. Tellus, 67A, 26536, doi:10.3402/tellusa.v67.26536.

    • Search Google Scholar
    • Export Citation
  • Ying, Y., and F. Zhang, 2015: An adaptive covariance relaxation method for ensemble data assimilation. Quart. J. Roy. Meteor. Soc., 141, 28982906, doi:10.1002/qj.2576.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., and Y. Weng, 2015: Predicting hurricane intensity and associated hazards: A five-year real-time forecast experiment with assimilation of airborne Doppler radar observations. Bull. Amer. Meteor. Soc., 96, 2533, doi:10.1175/BAMS-D-13-00231.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., C. Snyder, and J. Sun, 2004: Impacts of initial estimate and observation availability on convective-scale data assimilation with an ensemble Kalman filter. Mon. Wea. Rev., 132, 12381253, doi:10.1175/1520-0493(2004)132<1238:IOIEAO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., Y. Weng, J. A. Sippel, Z. Meng, and C. H. Bishop, 2009a: Cloud-resolving hurricane initialization and prediction through assimilation of Doppler radar observations with an ensemble Kalman filter. Mon. Wea. Rev., 137, 21052125, doi:10.1175/2009MWR2645.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., M. Zhang, and J. A. Hansen, 2009b: Coupling ensemble Kalman filter with four-dimensional variational data assimilation. Adv. Atmos. Sci., 26, 18, doi:10.1007/s00376-009-0001-8.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., Y. Weng, J. F. Gamache, and F. D. Marks, 2011: Performance of convection-permitting hurricane initialization and prediction during 2008–2010 with ensemble data assimilation of inner-core airborne Doppler radar observations. Geophys. Res. Lett., 38, L15810, doi:10.1029/2011GL048469.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., M. Minamide, and E. E. Clothiaux, 2016: Potential impacts of assimilating all-sky infrared satellite radiances from GOES-R on convection-permitting analysis and prediction of tropical cyclones. Geophys. Res. Lett., 43, 29542963, doi:10.1002/2016GL068468.

    • Search Google Scholar
    • Export Citation
  • Zhang, M., and F. Zhang, 2012: E4DVar: Coupling an ensemble Kalman filter with four-dimensional variational data assimilation in a limited-area weather prediction model. Mon. Wea. Rev., 140, 587600, doi:10.1175/MWR-D-11-00023.1.

    • Search Google Scholar
    • Export Citation
  • Zhen, Y., and F. Zhang, 2014: A probabilistic approach to adaptive covariance localization for serial ensemble square root filters. Mon. Wea. Rev., 142, 44994518, doi:10.1175/MWR-D-13-00390.1.

    • Search Google Scholar
    • Export Citation
  • Zhou, Y., D. McLaughlin, D. Entekhabi, and G.-H. Crystal Ng, 2008: An ensemble multiscale filter for large nonlinear data assimilation problems. Mon. Wea. Rev., 136, 678698, doi:10.1175/2007MWR2064.1.

    • Search Google Scholar
    • Export Citation
Save
  • Aksoy, A., F. Zhang, and J. W. Nielsen-Gammon, 2006: Ensemble-based simultaneous state and parameter estimation in a two-dimensional sea-breeze model. Mon. Wea. Rev., 134, 29512970, doi:10.1175/MWR3224.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., D. C. Dowell, and C. Snyder, 2009: A multicase comparative assessment of the ensemble Kalman filter for assimilation of radar observations. Part I: Storm-scale analyses. Mon. Wea. Rev., 137, 18051824, doi:10.1175/2008MWR2691.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., D. C. Dowell, and C. Snyder, 2010: A multicase comparative assessment of the ensemble Kalman filter for assimilation of radar observations. Part II: Short-range ensemble forecasts. Mon. Wea. Rev., 138, 12731292, doi:10.1175/2009MWR3086.1.

    • Search Google Scholar
    • Export Citation
  • Aksoy, A., S. D. Aberson, T. Vukicevic, K. J. Sellwood, S. Lorsolo, and X. Zhang, 2013: Assimilation of high-resolution tropical cyclone observations with an ensemble Kalman filter using NOAA/AOML/HRD’s HEDAS: Evaluation of the 2008–11 vortex-scale analyses. Mon. Wea. Rev., 141, 18421865, doi:10.1175/MWR-D-12-00194.1.

    • Search Google Scholar
    • Export Citation
  • Albers, S. C., J. A. McGinley, D. L. Birkenheuer, and J. R. Smart, 1996: The Local Analysis and Prediction System (LAPS): Analysis of clouds, precipitation, and temperature. Wea. Forecasting, 11, 273287, doi:10.1175/1520-0434(1996)011<0273:TLAAPS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Amezcua, J., K. Ide, E. Kalnay, and S. Reich, 2014: Ensemble transform Kalman–Bucy filters. Quart. J. Roy. Meteor. Soc., 140, 9951004, doi:10.1002/qj.2186.

    • Search Google Scholar
    • Export Citation
  • 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
  • Anderson, J. L., 2003: A local least squares framework for ensemble filtering. Mon. Wea. Rev., 131, 634642, doi:10.1175/1520-0493(2003)131<0634:ALLSFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2007: Exploring the need for localization in ensemble data assimilation using a hierarchical ensemble filter. Physica D, 230, 99111, doi:10.1016/j.physd.2006.02.011.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2009: Spatially and temporally varying adaptive covariance inflation for ensemble filters. Tellus, 61A, 7283, doi:10.1111/j.1600-0870.2008.00361.x.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2012: Localization and sampling error correction in ensemble Kalman filter data assimilation. Mon. Wea. Rev., 140, 23592371, doi:10.1175/MWR-D-11-00013.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., 2016: Reducing correlation sampling error in ensemble Kalman filter data assimilation. Mon. Wea. Rev., 144, 913925, doi:10.1175/MWR-D-15-0052.1.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., and S. L. Anderson, 1999: A Monte Carlo implementation of the nonlinear filtering problem to produce ensemble assimilations and forecasts. Mon. Wea. Rev., 127, 27412758, doi:10.1175/1520-0493(1999)127<2741:AMCIOT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Anderson, J. L., T. Hoar, K. Raeder, H. Liu, N. Collins, R. Torn, and A. Avellano, 2009: The Data Assimilation Research Testbed: A community facility. Bull. Amer. Meteor. Soc., 90, 12831296, doi:10.1175/2009BAMS2618.1.

    • Search Google Scholar
    • Export Citation
  • Annan, J. D., 2004: On the orthogonality of bred vectors. Mon. Wea. Rev., 132, 843849, doi:10.1175/1520-0493(2004)132<0843:OTOOBV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Annan, J. D., J. C. Hargreaves, N. R. Edwards, and R. Marsh, 2005: Parameter estimation in an intermediate complexity earth system model using an ensemble Kalman filter. Ocean Modell., 8, 135154, doi:10.1016/j.ocemod.2003.12.004.

    • Search Google Scholar
    • Export Citation
  • Aravéquia, J. A., I. Szunyogh, E. J. Fertig, E. Kalnay, D. Kuhl, and E. J. Kostelich, 2011: Evaluation of a strategy for the assimilation of satellite radiance observations with the local ensemble transform Kalman filter. Mon. Wea. Rev., 139, 19321951, doi:10.1175/2010MWR3515.1.

    • Search Google Scholar
    • Export Citation
  • Ballish, B., X. Cao, E. Kalnay, and M. Kanamitsu, 1992: Incremental nonlinear normal-mode initialization. Mon. Wea. Rev., 120, 17231734, doi:10.1175/1520-0493(1992)120<1723:INNMI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Barker, D. M., 2005: Southern high-latitude ensemble data assimilation in the Antartic mesoscale prediction system. Mon. Wea. Rev., 133, 34313449, doi:10.1175/MWR3042.1.

    • Search Google Scholar
    • Export Citation
  • Barker, D. M., and Coauthors, 2012: The Weather Research and Forecasting Model’s Community Variational/Ensemble Data Assimilation System: WRFDA. Bull. Amer. Meteor. Soc., 93, 831843, doi:10.1175/BAMS-D-11-00167.1.

    • Search Google Scholar
    • Export Citation
  • Bergemann, K., and S. Reich, 2010: A mollified ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 136, 16361643, doi:10.1002/qj.672.

    • Search Google Scholar
    • Export Citation
  • Berner, J., G. J. Shutts, M. Leutbecher, and T. N. Palmer, 2009: A spectral stochastic kinetic energy backscatter scheme and its impact on flow-dependent predictability in the ECMWF ensemble prediction system. J. Atmos. Sci., 66, 603626, doi:10.1175/2008JAS2677.1.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and Z. Toth, 1999: Ensemble transformation and adaptive observations. J. Atmos. Sci., 56, 17481765, doi:10.1175/1520-0469(1999)056<1748:ETAAO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., and D. Hodyss, 2009: Ensemble covariances adaptively localized with ECO-RAP. Part 1: Tests on simple error models. Tellus, 61A, 8496, doi:10.1111/j.1600-0870.2008.00371.x.

    • Search Google Scholar
    • Export Citation
  • Bishop, C. H., B. J. Etherton, and S. J. Majumdar, 2001: Adaptive sampling with the ensemble transform Kalman filter. Part I: Theoretical aspects. Mon. Wea. Rev., 129, 420436, doi:10.1175/1520-0493(2001)129<0420:ASWTET>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bloom, S. C., L. L. Takacs, A. M. da Silva, and D. Ledvina, 1996: Data assimilation using incremental analysis updates. Mon. Wea. Rev., 124, 12561271, doi:10.1175/1520-0493(1996)124<1256:DAUIAU>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., 2011: Impact and diagnosis of model error in the ECMWF ensemble of data assimilations. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 303318.

  • Bonavita, M., L. Torrisi, and F. Marcucci, 2010: Ensemble data assimilation with the CNMCA regional forecasting system. Quart. J. Roy. Meteor. Soc., 136, 132145, doi:10.1002/qj.553.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., L. Isaksen, and E. Hólm, 2012: On the use of EDA background error variances in the ECMWF 4D-Var. Quart. J. Roy. Meteor. Soc., 138, 15401559, doi:10.1002/qj.1899.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., M. Hamrud, and L. Isaksen, 2015: EnKF and hybrid gain ensemble data assimilation. Part II: EnKF and hybrid gain results. Mon. Wea. Rev., 143, 48654882, doi:10.1175/MWR-D-15-0071.1.

    • Search Google Scholar
    • Export Citation
  • Bonavita, M., E. Hólm, L. Isaksen, and M. Fisher, 2016: The evolution of the ECMWF hybrid data assimilation system. Quart. J. Roy. Meteor. Soc., 142, 287303, doi:10.1002/qj.2652.

    • Search Google Scholar
    • Export Citation
  • Bormann, N., and P. Bauer, 2010: Estimates of spatial and interchannel observation-error characteristics for current sounder radiances for numerical weather prediction. I: Methods and application to ATOVS data. Quart. J. Roy. Meteor. Soc., 136, 10361050, doi:10.1002/qj.616.

    • Search Google Scholar
    • Export Citation
  • Bowler, N. E., A. Arribas, K. R. Mylne, K. B. Robertson, and S. E. Beare, 2008: The MOGREPS short-range ensemble prediction system. Quart. J. Roy. Meteor. Soc., 134, 703722, doi:10.1002/qj.234.

    • Search Google Scholar
    • Export Citation
  • Brankart, J.-M., C. Ubelmann, C.-E. Testut, E. Cosme, P. Brasseur, and J. Verron, 2009: Efficient parameterization of the observation error covariance matrix for square root or ensemble Kalman filters: Application to ocean altimetry. Mon. Wea. Rev., 137, 19081927, doi:10.1175/2008MWR2693.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., 2012: Evaluation of a spatial/spectral covariance localization approach for atmospheric data assimilation. Mon. Wea. Rev., 140, 617636, doi:10.1175/MWR-D-10-05052.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., P. L. Houtekamer, C. Charette, H. L. Mitchell, and B. He, 2010a: Intercomparison of variational data assimilation and the ensemble Kalman filter for global deterministic NWP. Part I: Description and single-observation experiments. Mon. Wea. Rev., 138, 15501566, doi:10.1175/2009MWR3157.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., P. L. Houtekamer, C. Charette, H. L. Mitchell, and B. He, 2010b: Intercomparison of variational data assimilation and the ensemble Kalman filter for global deterministic NWP. Part II: One-month experiments with real observations. Mon. Wea. Rev., 138, 15671586, doi:10.1175/2009MWR3158.1.

    • Search Google Scholar
    • Export Citation
  • Buehner, M., and Coauthors, 2015: Implementation of deterministic weather forecasting systems based on ensemble–variational data assimilation at Environment Canada. Part I: The global system. Mon. Wea. Rev., 143, 25322559, doi:10.1175/MWR-D-14-00354.1.

    • Search Google Scholar
    • Export Citation
  • Buizza, R., M. Miller, and T. N. Palmer, 1999: Stochastic representation of model uncertainties in the ECMWF ensemble prediction system. Quart. J. Roy. Meteor. Soc., 125, 28872908, doi:10.1002/qj.49712556006.

    • Search Google Scholar
    • Export Citation
  • Burgers, G., P. J. van Leeuwen, and G. Evensen, 1998: Analysis scheme in the ensemble Kalman filter. Mon. Wea. Rev., 126, 17191724, doi:10.1175/1520-0493(1998)126<1719:ASITEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Campbell, W. F., C. H. Bishop, and D. Hodyss, 2010: Vertical covariance localization for satellite radiances in ensemble Kalman filters. Mon. Wea. Rev., 138, 282290, doi:10.1175/2009MWR3017.1.

    • Search Google Scholar
    • Export Citation
  • Candille, G., 2009: The multiensemble approach: The NAEFS example. Mon. Wea. Rev., 137, 16551665, doi:10.1175/2008MWR2682.1.

  • Candille, G., and O. Talagrand, 2005: Evaluation of probabilistic prediction systems for a scalar variable. Quart. J. Roy. Meteor. Soc., 131, 21312150, doi:10.1256/qj.04.71.

    • Search Google Scholar
    • Export Citation
  • Candille, G., and O. Talagrand, 2008: Impact of observational error on the validation of ensemble prediction systems. Quart. J. Roy. Meteor. Soc., 134, 959971, doi:10.1002/qj.268.

    • Search Google Scholar
    • Export Citation
  • Cardinali, C., 2009: Monitoring the observation impact on the short-range forecast. Quart. J. Roy. Meteor. Soc., 135, 239250, doi:10.1002/qj.366.

    • Search Google Scholar
    • Export Citation
  • Caron, J.-F., T. Milewski, M. Buehner, L. Fillion, M. Reszka, S. Macpherson, and J. St-James, 2015: Implementation of deterministic weather forecasting systems based on ensemble-variational data assimilation at Environment Canada. Part II: The regional system. Mon. Wea. Rev., 143, 25602580, doi:10.1175/MWR-D-14-00353.1.

    • Search Google Scholar
    • Export Citation
  • Caya, A., J. Sun, and C. Snyder, 2005: A comparison between the 4DVAR and the ensemble Kalman filter techniques for radar data assimilation. Mon. Wea. Rev., 133, 30813094, doi:10.1175/MWR3021.1.

    • Search Google Scholar
    • Export Citation
  • Chang, C.-C., S.-C. Yang, and C. Keppenne, 2014a: Applications of the mean recentering scheme to improve typhoon track prediction: A case study of typhoon Nanmadol (2011). J. Meteor. Soc. Japan, 92, 559584, doi:10.2151/jmsj.2014-604.

    • Search Google Scholar
    • Export Citation
  • Chang, W., K.-S. Chung, L. Fillion, and S.-J. Baek, 2014b: Radar data assimilation in the Canadian high-resolution ensemble Kalman filter system: Performance and verification with real summer cases. Mon. Wea. Rev., 142, 21182138, doi:10.1175/MWR-D-13-00291.1.

    • Search Google Scholar
    • Export Citation
  • Charron, M., G. Pellerin, L. Spacek, P. L. Houtekamer, N. Gagnon, H. L. Mitchell, and L. Michelin, 2010: Toward random sampling of model error in the Canadian Ensemble Prediction System. Mon. Wea. Rev., 138, 18771901, doi:10.1175/2009MWR3187.1.

    • Search Google Scholar
    • Export Citation
  • Clayton, A. M., A. C. Lorenc, and D. M. Barker, 2013: Operational implementation of a hybrid ensemble/4D-Var global data assimilation system at the Met Office. Quart. J. Roy. Meteor. Soc., 139, 14451461, doi:10.1002/qj.2054.

    • Search Google Scholar
    • Export Citation
  • Cohn, S. E., and D. F. Parrish, 1991: The behavior of forecast error covariances for a Kalman filter in two dimensions. Mon. Wea. Rev., 119, 17571785, doi:10.1175/1520-0493(1991)119<1757:TBOFEC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Cohn, S. E., A. da Silva, J. Guo, M. Sienkiewicz, and D. Lamich, 1998: Assessing the effects of data selection with the DAO physical-space statistical analysis system. Mon. Wea. Rev., 126, 29132926, doi:10.1175/1520-0493(1998)126<2913:ATEODS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Compo, G. P., J. S. Whitaker, and P. D. Sardeshmukh, 2006: Feasibility of a 100-year reanalysis using only surface pressure data. Bull. Amer. Meteor. Soc., 87, 175190, doi:10.1175/BAMS-87-2-175.

    • Search Google Scholar
    • Export Citation
  • Daley, R., 1991: Atmospheric Data Analysis. Cambridge University Press, 457 pp.

  • Daley, R., 1995: Estimating the wind field from chemical constituent observations: Experiments with a one-dimensional extended Kalman filter. Mon. Wea. Rev., 123, 181198, doi:10.1175/1520-0493(1995)123<0181:ETWFFC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Daley, R., and T. Mayer, 1986: Estimates of global analysis error from the global weather experiment observational network. Mon. Wea. Rev., 114, 16421653, doi:10.1175/1520-0493(1986)114<1642:EOGAEF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • 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
  • Dee, D. P., 2005: Bias and data assimilation. Quart. J. Roy. Meteor. Soc., 131, 33233343, doi:10.1256/qj.05.137.

  • Derber, J. C., and W.-S. Wu, 1998: The use of TOVS cloud-cleared radiances in the NCEP SSI analysis system. Mon. Wea. Rev., 126, 22872299, doi:10.1175/1520-0493(1998)126<2287:TUOTCC>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Derber, J. C., and F. Bouttier, 1999: A reformulation of the background error covariance in the ECMWF global data assimilation system. Tellus, 51A, 195221, doi:10.1034/j.1600-0870.1999.t01-2-00003.x.

    • Search Google Scholar
    • Export Citation
  • Desroziers, G., L. Berre, B. Chapnik, and P. Poli, 2005: Diagnosis of observation, background and analysis-error statistics in observation space. Quart. J. Roy. Meteor. Soc., 131, 33853396, doi:10.1256/qj.05.108.

    • Search Google Scholar
    • Export Citation
  • Dillon, M. E., and Coauthors, 2016: Application of the WRF-LETKF data assimilation system over southern South America: Sensitivity to model physics. Wea. Forecasting, 31, 217236, doi:10.1175/WAF-D-14-00157.1.

    • Search Google Scholar
    • Export Citation
  • Dong, J., and M. Xue, 2013: Assimilation of radial velocity and reflectivity data from coastal WSR-88D radars using an ensemble Kalman filter for the analysis and forecast of landfalling hurricane Ike (2008). Quart. J. Roy. Meteor. Soc., 139, 467487, doi:10.1002/qj.1970.

    • Search Google Scholar
    • Export Citation
  • Dowell, D. C., F. Zhang, L. J. Wicker, C. Snyder, and N. A. Crook, 2004: Wind and temperature retrievals in the 17 May 1981 Arcadia, Oklahoma, supercell: Ensemble Kalman filter experiments. Mon. Wea. Rev., 132, 19822005, doi:10.1175/1520-0493(2004)132<1982:WATRIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dowell, D. C., L. J. Wicker, and C. Snyder, 2011: Ensemble Kalman filter assimilation of radar observations of the 8 May 2003 Oklahoma City supercell: Influences of reflectivity observations on storm-scale analyses. Mon. Wea. Rev., 139, 272294, doi:10.1175/2010MWR3438.1.

    • Search Google Scholar
    • Export Citation
  • ECMWF, 2011: Proceedings of the ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, 20–24 June 2011. ECMWF, 370 pp.

  • English, S. J., R. J. Renshaw, P. C. Dibben, A. J. Smith, P. J. Rayer, C. Poulsen, F. W. Saunders, and J. R. Eyre, 2000: A comparison of the impact of TOVS and ATOVS satellite sounding data on the accuracy of numerical weather forecasts. Quart. J. Roy. Meteor. Soc., 126, 29112931, doi:10.1002/qj.49712656915.

    • Search Google Scholar
    • Export Citation
  • Errico, R. M., R. Yang, N. C. Privé, K.-S. Tai, R. Todling, M. E. Sienkiewicz, and J. Guo, 2013: Development and validation of observing-system simulation experiments at NASA’s Global Modeling and Assimilation Office. Quart. J. Roy. Meteor. Soc., 139, 11621178, doi:10.1002/qj.2027.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 1994: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99, 10 14310 162, doi:10.1029/94JC00572.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2003: The ensemble Kalman filter: Theoretical formulation and practical implementation. Ocean Dyn., 53, 343367, doi:10.1007/s10236-003-0036-9.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2004: Sampling strategies and square root analysis schemes for the EnKF. Ocean Dyn., 54, 539560, doi:10.1007/s10236-004-0099-2.

    • Search Google Scholar
    • Export Citation
  • Evensen, G., 2009: The ensemble Kalman filter for combined state and parameter estimation. IEEE Control Syst., 29, 83104, doi:10.1109/MCS.2009.932223.

    • Search Google Scholar
    • Export Citation
  • Fertig, E. J., J. Harlim, and B. R. Hunt, 2007: A comparative study of 4D-VAR and a 4D ensemble Kalman filter: Perfect model simulations with Lorenz-96. Tellus, 59A, 96100, doi:10.1111/j.1600-0870.2006.00205.x.

    • Search Google Scholar
    • Export Citation
  • Fillion, L., H. L. Mitchell, H. Ritchie, and A. Staniforth, 1995: The impact of a digital filter finalization technique in a global data assimilation system. Tellus, 47A, 304323, doi:10.1034/j.1600-0870.1995.t01-2-00002.x.

    • Search Google Scholar
    • Export Citation
  • Fisher, M., 2004: Background error covariance modelling. Proc. ECMWF Seminar on Recent Developments in Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 4563.

  • Fisher, M., and H. Auvinen, 2012: Long window 4D-Var. Proc. ECMWF Seminar on Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 189202.

  • Fishman, G. S., 1996: Monte Carlo: Concepts, Algorithms and Applications. Springer, 698 pp.

  • Fitzgerald, R. J., 1971: Divergence of the Kalman filter. IEEE Trans. Automat. Control, 16, 736747, doi:10.1109/TAC.1971.1099836.

  • Flowerdew, J., 2015: Towards a theory of optimal localisation. Tellus, 67A, 25257, doi:10.3402/tellusa.v67.25257.

  • Frehlich, R., 2006: Adaptive data assimilation including the effect of spatial variations in observation error. Quart. J. Roy. Meteor. Soc., 132, 12251257, doi:10.1256/qj.05.146.

    • Search Google Scholar
    • Export Citation
  • Fujita, T., D. J. Stensrud, and D. C. Dowell, 2007: Surface data assimilation using an ensemble Kalman filter approach with initial condition and model physics uncertainties. Mon. Wea. Rev., 135, 18461868, doi:10.1175/MWR3391.1.

    • Search Google Scholar
    • Export Citation
  • Gaspari, G., and S. E. Cohn, 1999: Construction of correlation functions in two and three dimensions. Quart. J. Roy. Meteor. Soc., 125, 723757, doi:10.1002/qj.49712555417.

    • Search Google Scholar
    • Export Citation
  • Gauthier, P., and J.-N. Thépaut, 2001: Impact of the digital filter as a weak constraint in the preoperational 4DVAR assimilation system of Météo-France. Mon. Wea. Rev., 129, 20892102, doi:10.1175/1520-0493(2001)129<2089:IOTDFA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Gauthier, P., M. Tanguay, S. Laroche, S. Pellerin, and J. Morneau, 2007: Extension of 3DVAR to 4DVAR: Implementation of 4DVAR at the Meteorological Service of Canada. Mon. Wea. Rev., 135, 23392354, doi:10.1175/MWR3394.1.

    • Search Google Scholar
    • Export Citation
  • Geer, A. J., and P. Bauer, 2011: Observation errors in all-sky data assimilation. Quart. J. Roy. Meteor. Soc., 137, 20242037, doi:10.1002/qj.830.

    • Search Google Scholar
    • Export Citation
  • Ghil, M., and P. Malanotte-Rizzoli, 1991: Data assimilation in meteorology and oceanography. Advances in Geophysics, Vol. 33, Academic Press, 141–266.

  • Ghil, M., S. Cohn, J. Tavantzis, K. Bube, and E. Isaacson, 1981: Applications of estimation theory to numerical weather prediction. Dynamic Meteorology—Data Assimilation Methods, L. Bengtsson, M. Ghil, and E. Källén, Eds., Springer-Verlag, 139–224.

  • Gilleland, E., D. A. Ahijevych, B. G. Brown, and E. E. Ebert, 2010: Verifying forecasts spatially. Bull. Amer. Meteor. Soc., 91, 13651373, doi:10.1175/2010BAMS2819.1.

    • Search Google Scholar
    • Export Citation
  • Golub, G. H., and C. F. Van Loan, 1996: Matrix Computations. 3rd ed. Johns Hopkins University Press, 694 pp.

  • Gorin, V. E., and M. D. Tsyrulnikov, 2011: Estimation of multivariate observation-error statistics for AMSU-A data. Mon. Wea. Rev., 139, 37653780, doi:10.1175/2011MWR3554.1.

    • Search Google Scholar
    • Export Citation
  • Grell, G. A., and D. Dévényi, 2002: A generalized approach to parameterizing convection combining ensemble and data assimilation techniques. Geophys. Res. Lett., 29, doi:10.1029/2002GL015311.

    • Search Google Scholar
    • Export Citation
  • Greybush, S. J., E. Kalnay, T. Miyoshi, K. Ide, and B. R. Hunt, 2011: Balance and ensemble Kalman filter localization techniques. Mon. Wea. Rev., 139, 511522, doi:10.1175/2010MWR3328.1.

    • Search Google Scholar
    • Export Citation
  • Ha, S., J. Berner, and C. Snyder, 2015: A comparison of model error representations in mesoscale ensemble data assimilation. Mon. Wea. Rev., 143, 38933911, doi:10.1175/MWR-D-14-00395.1.

    • Search Google Scholar
    • Export Citation
  • Hager, G., and G. Wellein, 2011: Introduction to High Performance Computing for Scientists and Engineers. Chapman & Hall/CRC, 330 pp.

  • Hamill, T. M., 2001: Interpretation of rank histograms for verifying ensemble forecasts. Mon. Wea. Rev., 129, 550560, doi:10.1175/1520-0493(2001)129<0550:IORHFV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., 2006: Ensemble-based atmospheric data assimilation. Predictability of Weather and Climate, T. Palmer and R. Hagedorn, Eds., Cambridge University Press, 124–156.

  • Hamill, T. M., and C. Snyder, 2000: A hybrid ensemble Kalman filter-3D variational analysis scheme. Mon. Wea. Rev., 128, 29052919, doi:10.1175/1520-0493(2000)128<2905:AHEKFV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., and J. S. Whitaker, 2011: What constrains spread growth in forecasts initialized from ensemble Kalman filters? Mon. Wea. Rev., 139, 117131, doi:10.1175/2010MWR3246.1.

    • Search Google Scholar
    • Export Citation
  • Hamill, T. M., J. S. Whitaker, and C. Snyder, 2001: Distance-dependent filtering of background error covariance estimates in an ensemble Kalman filter. Mon. Wea. Rev., 129, 27762790, doi:10.1175/1520-0493(2001)129<2776:DDFOBE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hamrud, M., M. Bonavita, and L. Isaksen, 2015: EnKF and hybrid gain ensemble data assimilation. Part I: EnKF implementation. Mon. Wea. Rev., 143, 48474864, doi:10.1175/MWR-D-14-00333.1.

    • Search Google Scholar
    • Export Citation
  • Hersbach, H., 2000: Decomposition of the continuous ranked probability score for ensemble prediction systems. Wea. Forecasting, 15, 559570, doi:10.1175/1520-0434(2000)015<0559:DOTCRP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., 1993: Global and local skill forecasts. Mon. Wea. Rev., 121, 18341846, doi:10.1175/1520-0493(1993)121<1834:GALSF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., 2011: The use of multiple parameterizations in ensembles. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 163173.

  • Houtekamer, P. L., and L. Lefaivre, 1997: Using ensemble forecasts for model validation. Mon. Wea. Rev., 125, 24162426, doi:10.1175/1520-0493(1997)125<2416:UEFFMV>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 1998: Data assimilation using an ensemble Kalman filter technique. Mon. Wea. Rev., 126, 796811, doi:10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 2001: A sequential ensemble Kalman filter for atmospheric data assimilation. Mon. Wea. Rev., 129, 123137, doi:10.1175/1520-0493(2001)129<0123:ASEKFF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., and H. L. Mitchell, 2005: Ensemble Kalman filtering. Quart. J. Roy. Meteor. Soc., 131, 32693289, doi:10.1256/qj.05.135.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., L. Lefaivre, J. Derome, H. Ritchie, and H. L. Mitchell, 1996: A system simulation approach to ensemble prediction. Mon. Wea. Rev., 124, 12251242, doi:10.1175/1520-0493(1996)124<1225:ASSATE>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., H. L. Mitchell, G. Pellerin, M. Buehner, M. Charron, L. Spacek, and B. Hansen, 2005: Atmospheric data assimilation with an ensemble Kalman filter: Results with real observations. Mon. Wea. Rev., 133, 604620, doi:10.1175/MWR-2864.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., H. L. Mitchell, and X. Deng, 2009: Model error representation in an operational ensemble Kalman filter. Mon. Wea. Rev., 137, 21262143, doi:10.1175/2008MWR2737.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., X. Deng, H. L. Mitchell, S.-J. Baek, and N. Gagnon, 2014a: Higher resolution in an operational ensemble Kalman filter. Mon. Wea. Rev., 142, 11431162, doi:10.1175/MWR-D-13-00138.1.

    • Search Google Scholar
    • Export Citation
  • Houtekamer, P. L., B. He, and H. L. Mitchell, 2014b: Parallel implementation of an ensemble Kalman filter. Mon. Wea. Rev., 142, 11631182, doi:10.1175/MWR-D-13-00011.1.

    • Search Google Scholar
    • Export Citation
  • Hu, M., M. Xue, and K. Brewster, 2006: 3DVAR and cloud analysis with WSR-88D level-II data for the prediction of the Fort Worth, Texas, tornadic thunderstorms. Part I: Cloud analysis and its impact. Mon. Wea. Rev., 134, 675698, doi:10.1175/MWR3092.1.

    • Search Google Scholar
    • Export Citation
  • Hu, X.-M., F. Zhang, and J. W. Nielsen-Gammon, 2010: Ensemble-based simultaneous state and parameter estimation for treatment of mesoscale model error: A real-data study. Geophys. Res. Lett., 37, L08802, doi:10.1029/2010GL043017.

    • Search Google Scholar
    • Export Citation
  • Huang, X.-Y., and P. Lynch, 1993: Diabetic digital-filtering initialization: Application to the HIRLAM model. Mon. Wea. Rev., 121, 589603, doi:10.1175/1520-0493(1993)121<0589:DDFIAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Hunt, B. R., and Coauthors, 2004: Four-dimensional ensemble Kalman filtering. Tellus, 56A, 273277, doi:10.1111/j.1600-0870.2004.00066.x.

    • Search Google Scholar
    • Export Citation
  • Hunt, B. R., E. J. Kostelich, and I. Szunyogh, 2007: Efficient data assimilation for spatiotemporal chaos: A local ensemble transform Kalman filter. Physica D, 230, 112126, doi:10.1016/j.physd.2006.11.008.

    • Search Google Scholar
    • Export Citation
  • Isaksen, L., 2012: Data assimilation on future computer architectures. Proc. ECMWF Seminar on Data Assimilation for Atmosphere and Ocean, Shinfield Park, Reading, United Kingdom, ECMWF, 301322.

  • Jones, R. H., 1965: Optimal estimation of initial conditions for numerical prediction. J. Atmos. Sci., 22, 658663, doi:10.1175/1520-0469(1965)022<0658:OEOICF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jung, Y., M. Xue, and G. Zhang, 2010a: Simulations of polarimetric radar signatures of a supercell storm using a two-moment bulk microphysics scheme. J. Appl. Meteor. Climatol., 49, 146163, doi:10.1175/2009JAMC2178.1.

    • Search Google Scholar
    • Export Citation
  • Jung, Y., M. Xue, and G. Zhang, 2010b: Simultaneous estimation of microphysical parameters and the atmospheric state using simulated polarimetric radar data and an ensemble Kalman filter in the presence of an observation operator error. Mon. Wea. Rev., 138, 539562, doi:10.1175/2009MWR2748.1.

    • Search Google Scholar
    • Export Citation
  • Kalman, R. E., 1960: A new approach to linear filtering and prediction problems. J. Basic Eng., 82, 3545, doi:10.1115/1.3662552.

  • 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., and A. Dalcher, 1987: Forecasting forecast skill. Mon. Wea. Rev., 115, 349356, doi:10.1175/1520-0493(1987)115<0349:FFS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Kalnay, E., and S.-C. Yang, 2010: Accelerating the spin-up of ensemble Kalman filtering. Quart. J. Roy. Meteor. Soc., 136, 16441651, doi:10.1002/qj.652.

    • Search Google Scholar
    • Export Citation
  • Kang, J.-S., E. Kalnay, J. Liu, I. Fung, T. Miyoshi, and K. Ide, 2011: “Variable localization�? in an ensemble Kalman filter: Application to the carbon cycle data assimilation. J. Geophys. Res., 116, D09110, doi:10.1029/2010JD014673.

    • Search Google Scholar
    • Export Citation
  • Kang, J.-S., E. Kalnay, T. Miyoshi, J. Liu, and I. Fung, 2012: Estimation of surface carbon fluxes with an advanced data assimilation methodology. J. Geophys. Res., 117, D24101, doi:10.1029/2012JD018259.

    • Search Google Scholar
    • Export Citation
  • Kelly, D., A. J. Majda, and X. T. Tong, 2015: Concrete ensemble Kalman filters with rigorous catastrophic filter divergence. Proc. Natl. Acad. Sci. USA, 112, 10 58910 594, doi:10.1073/pnas.1511063112.

    • Search Google Scholar
    • Export Citation
  • Kepert, J. D., 2009: Covariance localisation and balance in an ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 135, 11571176, doi:10.1002/qj.443.

    • Search Google Scholar
    • Export Citation
  • Kleist, D. T., and K. Ide, 2015a: An OSSE-based evaluation of hybrid variational-ensemble data assimilation for the NCEP GFS. Part I: System description and 3D-hybrid results. Mon. Wea. Rev., 143, 433451, doi:10.1175/MWR-D-13-00351.1.

    • Search Google Scholar
    • Export Citation
  • Kleist, D. T., and K. Ide, 2015b: An OSSE-based evaluation of hybrid variational-ensemble data assimilation for the NCEP GFS. Part II: 4DEnVar and hybrid variants. Mon. Wea. Rev., 143, 452470, doi:10.1175/MWR-D-13-00350.1.

    • Search Google Scholar
    • Export Citation
  • Krishnamurti, T. N., H. S. Bedi, W. Heckley, and K. Ingles, 1988: Reduction of the spinup time for evaporation and precipitation in a spectral model. Mon. Wea. Rev., 116, 907920, doi:10.1175/1520-0493(1988)116<0907:ROTSTF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Krishnamurti, T. N., C. M. Kishtawal, T. E. LaRow, D. R. Bachiochi, Z. Zhang, C. E. Williford, S. Gadgil, and S. Surendran, 1999: Improved weather and seasonal climate forecasts from multimodel superensemble. Science, 285, 15481550, doi:10.1126/science.285.5433.1548.

    • Search Google Scholar
    • Export Citation
  • Kumjian, M. R., 2013: Principles and applications of dual-polarization weather radar. Part III: Artifacts. J. Oper. Meteor., 1, 265274, doi:10.15191/nwajom.2013.0121.

    • Search Google Scholar
    • Export Citation
  • Kunii, M., 2014: The 1000-member ensemble Kalman filtering with the JMA nonhydrostatic mesoscale model on the K computer. J. Meteor. Soc. Japan, 92, 623633, doi:10.2151/jmsj.2014-607.

    • Search Google Scholar
    • Export Citation
  • Lang, S. T. K., M. Bonavita, and M. Leutbecher, 2015: On the impact of re-centering initial conditions for ensemble forecasts. Quart. J. Roy. Meteor. Soc., 141, 25712581, doi:10.1002/qj.2543.

    • Search Google Scholar
    • Export Citation
  • Lange, H., and G. C. Craig, 2014: The impact of data assimilation length scales on analysis and prediction of convective storms. Mon. Wea. Rev., 142, 37813808, doi:10.1175/MWR-D-13-00304.1.

    • Search Google Scholar
    • Export Citation
  • Lavaysse, C., M. Carrera, S. Bélair, N. Gagnon, R. Frenette, M. Charron, and M. K. Yau, 2013: Impact of surface parameter uncertainties within the Canadian regional ensemble prediction system. Mon. Wea. Rev., 141, 15061526, doi:10.1175/MWR-D-11-00354.1.

    • Search Google Scholar
    • Export Citation
  • Lawson, W. G., and J. A. Hansen, 2004: Implications of stochastic and deterministic filters as ensemble-based data assimilation methods in varying regimes of error growth. Mon. Wea. Rev., 132, 19661981, doi:10.1175/1520-0493(2004)132<1966:IOSADF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. L. Anderson, 2014: Empirical localization of observations for serial ensemble Kalman filter data assimilation in an atmospheric general circulation model. Mon. Wea. Rev., 142, 18351851, doi:10.1175/MWR-D-13-00288.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., and J. S. Whitaker, 2015: Model space localization is not always better than observation space localization for assimilation of satellite radiances. Mon. Wea. Rev., 143, 39483955, doi:10.1175/MWR-D-14-00413.1.

    • Search Google Scholar
    • Export Citation
  • Lei, L., D. R. Stauffer, and A. Deng, 2012: A hybrid nudging-ensemble Kalman filter approach to data assimilation in WRF/DART. Quart. J. Roy. Meteor. Soc., 138, 20662078, doi:10.1002/qj.1939.

    • Search Google Scholar
    • Export Citation
  • Leith, C. E., 1974: Theoretical skill of Monte Carlo forecasts. Mon. Wea. Rev., 102, 409418, doi:10.1175/1520-0493(1974)102<0409:TSOMCF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Lewis, J. M., S. Lakshmivarahan, and S. K. Dhall, 2006: Dynamic Data Assimilation: A Least Squares Approach. Cambridge University Press, 654 pp.

  • Li, X., M. Charron, L. Spacek, and G. Candille, 2008: A regional ensemble prediction system based on moist targeted singular vectors and stochastic parameter perturbations. Mon. Wea. Rev., 136, 443462, doi:10.1175/2007MWR2109.1.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., E. Kalnay, and T. Miyoshi, 2013: Effective assimilation of global precipitation: Simulation experiments. Tellus, 65A, 19915, doi:10.3402/tellusa.v65i0.19915.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., E. Kalnay, T. Miyoshi, and G. J. Huffman, 2016a: Statistical properties of global precipitation in the NCEP GFS model and TMPA observations for data assimilation. Mon. Wea. Rev., 144, 663679, doi:10.1175/MWR-D-15-0150.1.

    • Search Google Scholar
    • Export Citation
  • Lien, G.-Y., T. Miyoshi, and E. Kalnay, 2016b: Assimilation of TRMM multisatellite precipitation analysis with a low-resolution NCEP global forecast system. Mon. Wea. Rev., 144, 643661, doi:10.1175/MWR-D-15-0149.1.

    • Search Google Scholar
    • Export Citation
  • Lindskog, M., K. Salonen, H. Järvinen, and D. B. Michelson, 2004: Doppler radar wind data assimilation with HIRLAM 3DVAR. Mon. Wea. Rev., 132, 10811092, doi:10.1175/1520-0493(2004)132<1081:DRWDAW>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Liu, C., Q. Xiao, and B. Wang, 2008: An ensemble-based four-dimensional variational data assimilation scheme. Part I: Technical formulation and preliminary test. Mon. Wea. Rev., 136, 33633373, doi:10.1175/2008MWR2312.1.

    • Search Google Scholar
    • Export Citation
  • Liu, Z.-Q., and F. Rabier, 2003: The potential of high-density observations for numerical weather prediction: A study with simulated observations. Quart. J. Roy. Meteor. Soc., 129, 30133035, doi:10.1256/qj.02.170.

    • Search Google Scholar
    • Export Citation
  • Liu, Z.-Q., C. S. Schwartz, C. Snyder, and S.-Y. Ha, 2012: Impact of assimilating AMSU-A radiances on forecasts of 2008 Atlantic tropical cyclones initialized with a limited-area ensemble Kalman filter. Mon. Wea. Rev., 140, 40174034, doi:10.1175/MWR-D-12-00083.1.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., 1986: Analysis methods for numerical weather prediction. Quart. J. Roy. Meteor. Soc., 112, 11771194, doi:10.1002/qj.49711247414.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., 2003: The potential of the ensemble Kalman filter for NWP—A comparison with 4D-Var. Quart. J. Roy. Meteor. Soc., 129, 31833203, doi:10.1256/qj.02.132.

    • Search Google Scholar
    • Export Citation
  • Lorenc, A. C., N. E. Bowler, A. M. Clayton, S. R. Pring, and D. Fairbairn, 2015: Comparison of hybrid-4DEnVar and hybrid-4DVar data assimilation methods for global NWP. Mon. Wea. Rev., 143, 212229, doi:10.1175/MWR-D-14-00195.1.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 1965: A study of the predictability of a 28-variable atmospheric model. Tellus, 17A, 321333, doi:10.1111/j.2153-3490.1965.tb01424.x.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 1982: Atmospheric predictability experiments with a large numerical model. Tellus, 34A, 505513, doi:10.1111/j.2153-3490.1982.tb01839.x.

    • Search Google Scholar
    • Export Citation
  • Lorenz, E. N., 2005: Designing chaotic models. J. Atmos. Sci., 62, 15741587, doi:10.1175/JAS3430.1.

  • Lynch, P., and X.-Y. Huang, 1992: Initialization of the HIRLAM model using a digital filter. Mon. Wea. Rev., 120, 10191034, doi:10.1175/1520-0493(1992)120<1019:IOTHMU>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Machenhauer, B., 1977: On the dynamics of gravity oscillations in a shallow water model, with application to normal mode initialization. Contrib. Atmos. Phys., 50, 253271.

    • Search Google Scholar
    • Export Citation
  • Mandel, J., 2006: Efficient implementation of the ensemble Kalman filter. Center for Computational Mathematics Rep. 231, University of Colorado at Denver and Health Sciences Center, Denver, CO, 9 pp.

  • Mass, C. F., D. Ovens, K. Westrick, and B. A. Colle, 2002: Does increasing horizontal resolution produce more skillful forecasts? Bull. Amer. Meteor. Soc., 83, 407430, doi:10.1175/1520-0477(2002)083<0407:DIHRPM>2.3.CO;2.

    • Search Google Scholar
    • Export Citation
  • McNally, T., 2004: The assimilation of stratospheric satellite data at ECMWF. Proc. ECMWF/SPARC Workshop on Modelling and Assimilation for the Stratosphere and Tropopause, Shinfield Park, Reading, United Kingdom, ECMWF, 103106.

  • McTaggart-Cowan, R., C. Girard, A. Plante, and M. Desgagné, 2011: The utility of upper-boundary nesting in NWP. Mon. Wea. Rev., 139, 21172144, doi:10.1175/2010MWR3633.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2007: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part II: Imperfect model experiments. Mon. Wea. Rev., 135, 14031423, doi:10.1175/MWR3352.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2008a: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part III: Comparison with 3DVAR in a real-data case study. Mon. Wea. Rev., 136, 522540, doi:10.1175/2007MWR2106.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2008b: Tests of an ensemble Kalman filter for mesoscale and regional-scale data assimilation. Part IV: Comparison with 3DVAR in a month-long experiment. Mon. Wea. Rev., 136, 36713682, doi:10.1175/2008MWR2270.1.

    • Search Google Scholar
    • Export Citation
  • Meng, Z., and F. Zhang, 2011: Limited-area ensemble-based data assimilation. Mon. Wea. Rev., 139, 20252045, doi:10.1175/2011MWR3418.1.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., and P. L. Houtekamer, 2000: An adaptive ensemble Kalman filter. Mon. Wea. Rev., 128, 416433, doi:10.1175/1520-0493(2000)128<0416:AAEKF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., and P. L. Houtekamer, 2009: Ensemble Kalman filter configurations and their performance with the logistic map. Mon. Wea. Rev., 137, 43254343, doi:10.1175/2009MWR2823.1.

    • Search Google Scholar
    • Export Citation
  • Mitchell, H. L., P. L. Houtekamer, and G. Pellerin, 2002: Ensemble size, balance, and model-error representation in an ensemble Kalman filter. Mon. Wea. Rev., 130, 27912808, doi:10.1175/1520-0493(2002)130<2791:ESBAME>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., 2011: The Gaussian approach to adaptive covariance inflation and its implementation with the local ensemble transform Kalman filter. Mon. Wea. Rev., 139, 15191535, doi:10.1175/2010MWR3570.1.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., and K. Kondo, 2013: A multi-scale localization approach to an ensemble Kalman filter. SOLA, 9, 170173, doi:10.2151/sola.2013-038.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., Y. Sato, and T. Kadowaki, 2010: Ensemble Kalman filter and 4D-Var intercomparison with the Japanese operational global analysis and prediction system. Mon. Wea. Rev., 138, 28462866, doi:10.1175/2010MWR3209.1.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., E. Kalnay, and H. Li, 2013: Estimating and including observation-error correlations in data assimilation. Inverse Probl. Sci. Eng., 21, 387398, doi:10.1080/17415977.2012.712527.

    • Search Google Scholar
    • Export Citation
  • Miyoshi, T., K. Kondo, and T. Imamura, 2014: The 10,240-member ensemble Kalman filtering with an intermediate AGCM. Geophys. Res. Lett., 41, 52645271, doi:10.1002/2014GL060863.

    • Search Google Scholar
    • Export Citation
  • Molteni, F., 2003: Atmospheric simulations using a GCM with simplified physical parametrizations. I: Model climatology and variability in multi-decadal experiments. Climate Dyn., 20, 175191.

    • Search Google Scholar
    • Export Citation
  • Morrison, H., A. Morales, and C. Villanueva-Birriel, 2015: Concurrent sensitivities of an idealized deep convective storm to parameterization of microphysics, horizontal grid resolution, and environmental static stability. Mon. Wea. Rev., 143, 20822104, doi:10.1175/MWR-D-14-00271.1.

    • Search Google Scholar
    • Export Citation
  • Murphy, J., and Coauthors, 2011: Perturbed parameter ensembles as a tool for sampling model uncertainties and making climate projections. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 183208.

  • Mylne, K. R., R. E. Evans, and R. T. Clark, 2002: Multi-model multi-analysis ensembles in quasi-operational medium-range forecasting. Quart. J. Roy. Meteor. Soc., 128, 361384, doi:10.1256/00359000260498923.

    • Search Google Scholar
    • Export Citation
  • Nerger, L., 2015: On serial observation processing in localized ensemble Kalman filters. Mon. Wea. Rev., 143, 15541567, doi:10.1175/MWR-D-14-00182.1.

    • Search Google Scholar
    • Export Citation
  • Newton, C. W., 1954: Analysis and data problems in relation to numerical prediction. Bull. Amer. Meteor. Soc., 35, 287294.

  • Oczkowski, M., I. Szunyogh, and D. J. Patil, 2005: Mechanisms for the development of locally low-dimensional atmospheric dynamics. J. Atmos. Sci., 62, 11351156, doi:10.1175/JAS3403.1.

    • Search Google Scholar
    • Export Citation
  • Ott, E., and Coauthors, 2004: A local ensemble Kalman filter for atmospheric data assimilation. Tellus, 56A, 415428, doi:10.1111/j.1600-0870.2004.00076.x.

    • Search Google Scholar
    • Export Citation
  • Palmer, T. N., 2012: Towards the probabilistic earth-system simulator: A vision for the future of climate and weather prediction. Quart. J. Roy. Meteor. Soc., 138, 841861, doi:10.1002/qj.1923.

    • Search Google Scholar
    • Export Citation
  • Pan, Y., K. Zhu, M. Xue, X. Wang, M. Hu, S. G. Benjamin, S. S. Weygandt, and J. S. Whitaker, 2014: A GSI-based coupled EnSRF-En3DVar hybrid data assimilation system for the operational rapid refresh model: Tests at a reduced resolution. Mon. Wea. Rev., 142, 37563780, doi:10.1175/MWR-D-13-00242.1.

    • Search Google Scholar
    • Export Citation
  • Parrish, D. F., and J. C. Derber, 1992: The National Meteorological Center’s spectral statistical-interpolation analysis system. Mon. Wea. Rev., 120, 17471763, doi:10.1175/1520-0493(1992)120<1747:TNMCSS>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Pennell, C., and T. Reichler, 2011: On the effective number of climate models. J. Climate, 24, 23582367, doi:10.1175/2010JCLI3814.1.

  • Penny, S. G., 2014: The hybrid local ensemble transform Kalman filter. Mon. Wea. Rev., 142, 21392149, doi:10.1175/MWR-D-13-00131.1.

  • Penny, S. G., D. W. Behringer, J. A. Carton, and E. Kalnay, 2015: A hybrid global ocean data assimilation system at NCEP. Mon. Wea. Rev., 143, 46604677, doi:10.1175/MWR-D-14-00376.1.

    • Search Google Scholar
    • Export Citation
  • Petersen, D. P., 1968: On the concept and implementation of sequential analysis for linear random fields. Tellus, 20A, 673686, doi:10.1111/j.2153-3490.1968.tb00410.x.

    • Search Google Scholar
    • Export Citation
  • Pires, C., R. Vautard, and O. Talagrand, 1996: On extending the limits of variational assimilation in nonlinear chaotic systems. Tellus, 48A, 96121, doi:10.1034/j.1600-0870.1996.00006.x.

    • Search Google Scholar
    • Export Citation
  • Plant, R. S., and G. C. Craig, 2008: A stochastic parameterization for deep convection based on equilibrium statistics. J. Atmos. Sci., 65, 87105, doi:10.1175/2007JAS2263.1.

    • Search Google Scholar
    • Export Citation
  • Polavarapu, S., M. Tanguay, and L. Fillion, 2000: Four-dimensional variational data assimilation with digital filter initialization. Mon. Wea. Rev., 128, 24912510, doi:10.1175/1520-0493(2000)128<2491:FDVDAW>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Polavarapu, S., S. Ren, A. M. Clayton, D. Sankey, and Y. Rochon, 2004: On the relationship between incremental analysis updating and incremental digital filtering. Mon. Wea. Rev., 132, 24952502, doi:10.1175/1520-0493(2004)132<2495:OTRBIA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2014: Intercomparison and coupling of ensemble and four-dimensional variational data assimilation methods for the analysis and forecasting of Hurricane Karl (2010). Mon. Wea. Rev., 142, 33473364, doi:10.1175/MWR-D-13-00394.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2015: Systematic comparison of four-dimensional data assimilation methods with and without the tangent linear model using hybrid background error covariance: E4DVar versus 4DEnVar. Mon. Wea. Rev., 143, 16011621, doi:10.1175/MWR-D-14-00224.1.

    • Search Google Scholar
    • Export Citation
  • Poterjoy, J., and F. Zhang, 2016: Comparison of hybrid four-dimensional data assimilation methods with and without the tangent linear and adjoint models for predicting the life cycle of Hurricane Karl (2010). Mon. Wea. Rev., 144, 14491468, doi:10.1175/MWR-D-15-0116.1.

    • Search Google Scholar
    • Export Citation
  • Potter, J. E., 1964: W matrix augmentation. MIT Instrumentation Laboratory, Memo. SGA 5-64, Massachusetts Institute of Technology, Cambridge, MA.

  • Press, W. H., S. A. Teukolsky, W. T. Vetterling, and B. P. Flannery, 1992: Numerical Recipes in FORTRAN: The Art of Scientific Computing. 2nd ed. Cambridge University Press, 933 pp.

  • Putnam, B. J., M. Xue, Y. Jung, N. Snook, and G. Zhang, 2014: The analysis and prediction of microphysical states and polarimetric radar variables in a mesoscale convective system using double-moment microphysics, multinetwork radar data, and the ensemble Kalman filter. Mon. Wea. Rev., 142, 141162, doi:10.1175/MWR-D-13-00042.1.

    • Search Google Scholar
    • Export Citation
  • Rabier, F., H. Järvinen, E. Klinker, J.-F. Mahfouf, and A. Simmons, 2000: The ECMWF operational implementation of four-dimensional variational assimilation. I: Experimental results with simplified physics. Quart. J. Roy. Meteor. Soc., 126, 11431170, doi:10.1002/qj.49712656415.

    • Search Google Scholar
    • Export Citation
  • Rodgers, C. D., 2000: Inverse Methods for Atmospheric Sounding. World Scientific, 240 pp.

  • Ruiz, J., and M. Pulido, 2015: Parameter estimation using ensemble-based data assimilation in the presence of model error. Mon. Wea. Rev., 143, 15681582, doi:10.1175/MWR-D-14-00017.1.

    • Search Google Scholar
    • Export Citation
  • Rutherford, I. D., 1976: An operational 3-dimensional multivariate statistical objective analysis scheme. Proc. JOC Study Group Conf. on Four-Dimensional Data Assimilation, Rep. 11, Paris, France, WMO/JCSU, 98121.

  • Ryzhkov, A. V., 2007: The impact of beam broadening on the quality of radar polarimetric data. J. Atmos. Oceanic Technol., 24, 729744, doi:10.1175/JTECH2003.1.

    • Search Google Scholar
    • Export Citation
  • Sakov, P., and L. Bertino, 2011: Relation between two common localisation methods for the EnKF. Comput. Geosci., 15, 225237, doi:10.1007/s10596-010-9202-6.

    • Search Google Scholar
    • Export Citation
  • Schraff, C., H. Reich, A. Rhodin, A. Schomburg, K. Stephan, A. Periáñez, and R. Potthast, 2016: Kilometre-scale ensemble data assimilation for the COSMO model (KENDA). Quart. J. Roy. Meteor. Soc., 142, 14531472, doi:10.1002/qj.2748.

    • Search Google Scholar
    • Export Citation
  • Schwartz, C. S., G. S. Romine, R. A. Sobash, K. R. Fossell, and M. L. Weisman, 2015: NCAR’s experimental real-time convection-allowing ensemble prediction system. Wea. Forecasting, 30, 16451654, doi:10.1175/WAF-D-15-0103.1.

    • Search Google Scholar
    • Export Citation
  • Shutts, G., 2005: A kinetic energy backscatter algorithm for use in ensemble prediction systems. Quart. J. Roy. Meteor. Soc., 131, 30793102, doi:10.1256/qj.04.106.

    • Search Google Scholar
    • Export Citation
  • Shutts, G., and A. Callado Pallares, 2011: Tracking down the origin of NWP model uncertainty: coarse-graining studies. Proc. ECMWF Workshop on Representing Model Uncertainty and Error in Numerical Weather and Climate Prediction Models, Shinfield Park, Reading, United Kingdom, ECMWF, 221231.

  • Simmons, A. J., 1996: The skill of 500 hPa height forecasts. Proc. ECMWF Seminar on Predictability, Shinfield Park, Reading, United Kingdom, ECMWF, 1968.

  • Sippel, J. A., F. Zhang, Y. Weng, L. Tian, G. M. Heymsfield, and S. A. Braun, 2014: Ensemble Kalman filter assimilation of HIWRAP observations of Hurricane Karl (2010) from the unmanned Global Hawk aircraft. Mon. Wea. Rev., 142, 45594580, doi:10.1175/MWR-D-14-00042.1.

    • Search Google Scholar
    • Export Citation
  • Sluka, T. C., S. G. Penny, E. Kalnay, and T. Miyoshi, 2016: Assimilating atmospheric observations into the ocean using strongly coupled ensemble data assimilation. Geophys. Res. Lett., 43, 752759, doi:10.1002/2015GL067238.

    • Search Google Scholar
    • Export Citation
  • Snook, N., M. Xue, and Y. Jung, 2011: Analysis of a tornadic mesoscale convective vortex based on ensemble Kalman filter assimilation of CASA X-band and WSR-88D radar data. Mon. Wea. Rev., 139, 34463468, doi:10.1175/MWR-D-10-05053.1.

    • Search Google Scholar
    • Export Citation
  • Snook, N., M. Xue, and Y. Jung, 2015: Multiscale EnKF assimilation of radar and conventional observations and ensemble forecasting for a tornadic mesoscale convective system. Mon. Wea. Rev., 143, 10351057, doi:10.1175/MWR-D-13-00262.1.

    • Search Google Scholar
    • Export Citation
  • Snyder, C., 2015: Introduction to the Kalman filter. Les Houches 2012: Advanced Data Assimilation for Geosciences, E. Blayo et al., Eds., Oxford University Press, 75–120.

  • Snyder, C., and F. Zhang, 2003: Assimilation of simulated Doppler radar observations with an ensemble Kalman filter. Mon. Wea. Rev., 131, 16631677, doi:10.1175//2555.1.

    • Search Google Scholar
    • Export Citation
  • Steinhoff, J., and D. Underhill, 1994: Modification of the Euler equations for “vorticity confinement�?: Application to the computation of interacting vortex rings. Phys. Fluids A Fluid Dyn., 6, 27382744, doi:10.1063/1.868164.

    • Search Google Scholar
    • Export Citation
  • Stewart, L. M., S. L. Dance, and N. K. Nichols, 2013: Data assimilation with correlated observation errors: Experiments with a 1-D shallow water model. Tellus, 65A, 19546, doi:10.3402/tellusa.v65i0.19546.

    • Search Google Scholar
    • Export Citation
  • Strohmaier, E., J. Dongarra, H. Simon, and M. Meuer, 2015: The Top500 list: Twenty years of insight into HPC performance. Accessed December 2015. [Available online at www.top500.org/lists/2015/11.]

  • Swinbank, R., and Coauthors, 2015: The TIGGE project and its achievements. Bull. Amer. Meteor. Soc., 97, 4967, doi:10.1175/BAMS-D-13-00191.1.

    • Search Google Scholar
    • Export Citation
  • Szunyogh, I., E. J. Kostelich, G. Gyarmati, E. Kalnay, B. R. Hunt, E. Ott, E. Satterfield, and J. A. Yorke, 2008: A local ensemble transform Kalman filter data assimilation system for the NCEP global model. Tellus, 60A, 113130, doi:10.1111/j.1600-0870.2007.00274.x.

    • Search Google Scholar
    • Export Citation
  • Tarantola, A., 1987: Inverse Problem Theory. Elsevier, 613 pp.

  • Tarantola, A., 2006: Popper, Bayes and the inverse problem. Nature Phys., 2, 492494, doi:10.1038/nphys375.

  • Tardif, R., G. J. Hakim, and C. Snyder, 2015: Coupled atmosphere–ocean data assimilation experiments with a low-order model and CMIP5 model data. Climate Dyn., 45, 14151427, doi:10.1007/s00382-014-2390-3.

    • Search Google Scholar
    • Export Citation
  • Temperton, C., and M. Roch, 1991: Implicit normal mode initialization for an operational regional model. Mon. Wea. Rev., 119, 667677, doi:10.1175/1520-0493(1991)119<0667:INMIFA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Thépaut, J.-N., and P. Courtier, 1991: Four-dimensional variational data assimilation using the adjoint of a multilevel primitive-equation model. Quart. J. Roy. Meteor. Soc., 117, 12251254, doi:10.1002/qj.49711750206.

    • Search Google Scholar
    • Export Citation
  • Thomas, S. J., J. P. Hacker, and J. L. Anderson, 2009: A robust formulation of the ensemble Kalman filter. Quart. J. Roy. Meteor. Soc., 135, 507521, doi:10.1002/qj.372.

    • Search Google Scholar
    • Export Citation
  • Thompson, T. E., L. J. Wicker, X. Wang, and C. Potvin, 2015: A comparison between the local ensemble transform Kalman filter and the ensemble square root filter for the assimilation of radar data in convective-scale models. Quart. J. Roy. Meteor. Soc., 141, 11631176, doi:10.1002/qj.2423.

    • Search Google Scholar
    • Export Citation
  • Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker, 2003: Ensemble square root filters. Mon. Wea. Rev., 131, 14851490, doi:10.1175/1520-0493(2003)131<1485:ESRF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Tong, M., and M. Xue, 2005: Ensemble Kalman filter assimilation of Doppler radar data with a compressible nonhydrostatic model: OSS experiments. Mon. Wea. Rev., 133, 17891807, doi:10.1175/MWR2898.1.

    • Search Google Scholar
    • Export Citation
  • Torn, R. D., and G. J. Hakim, 2008: Performance characteristics of a pseudo-operational ensemble Kalman filter. Mon. Wea. Rev., 136, 39473963, doi:10.1175/2008MWR2443.1.

    • Search Google Scholar
    • Export Citation
  • Torn, R. D., G. J. Hakim, and C. Snyder, 2006: Boundary conditions for limited-area ensemble Kalman filters. Mon. Wea. Rev., 134, 24902502, doi:10.1175/MWR3187.1.

    • Search Google Scholar
    • Export Citation
  • Toth, Z., and E. Kalnay, 1997: Ensemble forecasting at NCEP and the breeding method. Mon. Wea. Rev., 125, 32973319, doi:10.1175/1520-0493(1997)125<3297:EFANAT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Verlaan, M., and A. W. Heemink, 2001: Nonlinearity in data assimilation applications: A practical method for analysis. Mon. Wea. Rev., 129, 15781589, doi:10.1175/1520-0493(2001)129<1578:NIDAAA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Wang, S., M. Xue, A. Schenkman, and J. Min, 2013: An iterative ensemble square root filter and tests with simulated radar data for storm-scale data assimilation. Quart. J. Roy. Meteor. Soc., 139, 18881903, doi:10.1002/qj.2077.

    • Search Google Scholar
    • Export Citation
  • Wang, X., and T. Lei, 2014: GSI-based four-dimensional ensemble-variational (4DEnsVar) data assimilation: Formulation and single- resolution experiments with real data for NCEP global forecast system. Mon. Wea. Rev., 142, 33033325, doi:10.1175/MWR-D-13-00303.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., C. Snyder, and T. M. Hamill, 2007: On the theoretical equivalence of differently proposed ensemble-3DVAR hybrid analysis schemes. Mon. Wea. Rev., 135, 222227, doi:10.1175/MWR3282.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., D. M. Barker, C. Snyder, and T. M. Hamill, 2008: A hybrid ETKF–3DVAR data assimilation scheme for the WRF model. Part I: Observing system simulation experiment. Mon. Wea. Rev., 136, 51165131, doi:10.1175/2008MWR2444.1.

    • Search Google Scholar
    • Export Citation
  • Wang, X., D. Parrish, D. Kleist, and J. Whitaker, 2013: GSI 3DVar-based ensemble-variational hybrid data assimilation for NCEP global forecast system: Single-resolution experiments. Mon. Wea. Rev., 141, 40984117, doi:10.1175/MWR-D-12-00141.1.

    • Search Google Scholar
    • Export Citation
  • Wei, M., and Z. Toth, 2003: A new measure of ensemble performance: Perturbation versus error correlation analysis (PECA). Mon. Wea. Rev., 131, 15491565, doi:10.1175//1520-0493(2003)131<1549:ANMOEP>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Weng, Y., and F. Zhang, 2012: Assimilating airborne Doppler radar observations with an ensemble Kalman filter for convection-permitting hurricane initialization and prediction: Katrina (2005). Mon. Wea. Rev., 140, 841859, doi:10.1175/2011MWR3602.1.

    • Search Google Scholar
    • Export Citation
  • Weng, Y., and F. Zhang, 2016: Advances in convection-permitting tropical cyclone analysis and prediction through EnKF assimilation of reconnaissance aircraft observations. J. Meteor. Soc. Japan, 94, 345358, doi:10.2151/jmsj.2016-018.

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and T. M. Hamill, 2002: Ensemble data assimilation without perturbed observations. Mon. Wea. Rev., 130, 19131924, doi:10.1175/1520-0493(2002)130<1913:EDAWPO>2.0.CO;2; Corrigendum, 134, 1722, doi:10.1175/MWR3156.1

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., and T. M. Hamill, 2012: Evaluating methods to account for system errors in ensemble data assimilation. Mon. Wea. Rev., 140, 30783089, doi:10.1175/MWR-D-11-00276.1.

    • Search Google Scholar
    • Export Citation
  • Whitaker, J. S., T. M. Hamill, X. Wei, Y. Song, and Z. Toth, 2008: Ensemble data assimilation with the NCEP global forecast system. Mon. Wea. Rev., 136, 463482, doi:10.1175/2007MWR2018.1.

    • Search Google Scholar
    • Export Citation
  • Xue, M., M. Tong, and K. K. Droegemeier, 2006: An OSSE framework based on the ensemble square root Kalman filter for evaluating the impact of data from radar networks on thunderstorm analysis and forecasting. J. Atmos. Oceanic Technol., 23, 4666, doi:10.1175/JTECH1835.1.

    • Search Google Scholar
    • Export Citation
  • Xue, M., J. Schleif, F. Kong, K. W. Thomas, Y. Wang, and K. Zhu, 2013: Track and intensity forecasting of hurricanes: Impact of convection-permitting resolution and global ensemble Kalman filter analysis on 2010 Atlantic season forecasts. Wea. Forecasting, 28, 13661384, doi:10.1175/WAF-D-12-00063.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, B. Hunt, and N. E. Bowler, 2009: Weight interpolation for efficient data assimilation with the Local Ensemble Transform Kalman Filter. Quart. J. Roy. Meteor. Soc., 135, 251262, doi:10.1002/qj.353.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and B. Hunt, 2012a: Handling nonlinearity in an ensemble Kalman filter: Experiments with the three-variable Lorenz model. Mon. Wea. Rev., 140, 26282646, doi:10.1175/MWR-D-11-00313.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and T. Miyoshi, 2012b: Accelerating the EnKF spinup for typhoon assimilation and prediction. Wea. Forecasting, 27, 878897, doi:10.1175/WAF-D-11-00153.1.

    • Search Google Scholar
    • Export Citation
  • Yang, S.-C., E. Kalnay, and T. Enomoto, 2015: Ensemble singular vectors and their use as additive inflation in EnKF. Tellus, 67A, 26536, doi:10.3402/tellusa.v67.26536.

    • Search Google Scholar
    • Export Citation
  • Ying, Y., and F. Zhang, 2015: An adaptive covariance relaxation method for ensemble data assimilation. Quart. J. Roy. Meteor. Soc., 141, 28982906, doi:10.1002/qj.2576.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., and Y. Weng, 2015: Predicting hurricane intensity and associated hazards: A five-year real-time forecast experiment with assimilation of airborne Doppler radar observations. Bull. Amer. Meteor. Soc., 96, 2533, doi:10.1175/BAMS-D-13-00231.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., C. Snyder, and J. Sun, 2004: Impacts of initial estimate and observation availability on convective-scale data assimilation with an ensemble Kalman filter. Mon. Wea. Rev., 132, 12381253, doi:10.1175/1520-0493(2004)132<1238:IOIEAO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., Y. Weng, J. A. Sippel, Z. Meng, and C. H. Bishop, 2009a: Cloud-resolving hurricane initialization and prediction through assimilation of Doppler radar observations with an ensemble Kalman filter. Mon. Wea. Rev., 137, 21052125, doi:10.1175/2009MWR2645.1.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., M. Zhang, and J. A. Hansen, 2009b: Coupling ensemble Kalman filter with four-dimensional variational data assimilation. Adv. Atmos. Sci., 26, 18, doi:10.1007/s00376-009-0001-8.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., Y. Weng, J. F. Gamache, and F. D. Marks, 2011: Performance of convection-permitting hurricane initialization and prediction during 2008–2010 with ensemble data assimilation of inner-core airborne Doppler radar observations. Geophys. Res. Lett., 38, L15810, doi:10.1029/2011GL048469.

    • Search Google Scholar
    • Export Citation
  • Zhang, F., M. Minamide, and E. E. Clothiaux, 2016: Potential impacts of assimilating all-sky infrared satellite radiances from GOES-R on convection-permitting analysis and prediction of tropical cyclones. Geophys. Res. Lett., 43, 29542963, doi:10.1002/2016GL068468.

    • Search Google Scholar
    • Export Citation
  • Zhang, M., and F. Zhang, 2012: E4DVar: Coupling an ensemble Kalman filter with four-dimensional variational data assimilation in a limited-area weather prediction model. Mon. Wea. Rev., 140, 587600, doi:10.1175/MWR-D-11-00023.1.

    • Search Google Scholar
    • Export Citation
  • Zhen, Y., and F. Zhang, 2014: A probabilistic approach to adaptive covariance localization for serial ensemble square root filters. Mon. Wea. Rev., 142, 44994518, doi:10.1175/MWR-D-13-00390.1.

    • Search Google Scholar
    • Export Citation
  • Zhou, Y., D. McLaughlin, D. Entekhabi, and G.-H. Crystal Ng, 2008: An ensemble multiscale filter for large nonlinear data assimilation problems. Mon. Wea. Rev., 136, 678698, doi:10.1175/2007MWR2064.1.

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

    The general recipe for the application of the Monte Carlo procedure. The best estimates are plotted with the large open circles. The black box is a procedure to transform input into output. To obtain information on the uncertainty in the output, the procedure is executed once for each member in the input ensemble. The individual estimates in each ensemble are indicated by the small filled circles.

  • Fig. 2.

    The average number of cores for the top 10 computers of the top500 list (Strohmaier et al. 2015). For each year, the November list has been used.

  • Fig. 3.

    Schematic diagram to dramatize the difficulty of localization within the assimilation window. An observation is made in area A and valid at the beginning of time window. During the time window a Lagrangian particle can move to area B, at the central analysis time, and to area C at the final time. At the central time, one may want to have the impact of the observation localized in the circle around area B and at the final time in the circle around area C. The assumption is that the localizing area is very small compared to the distance the Lagrangian particle can cover during the assimilation window.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 18196 8108 3641
PDF Downloads 6377 1345 159