Advanced Search
Article Contents

Using Analysis State to Construct a Forecast Error Covariance Matrix in Ensemble Kalman Filter Assimilation

Fund Project:

doi: 10.1007/s00376-012-2133-5

  • Correctly estimating the forecast error covariance matrix is a key step in any data assimilation scheme. If it is not correctly estimated, the assimilated states could be far from the true states. A popular method to address this problem is error covariance matrix inflation. That is, to multiply the forecast error covariance matrix by an appropriate factor. In this paper, analysis states are used to construct the forecast error covariance matrix and an adaptive estimation procedure associated with the error covariance matrix inflation technique is developed. The proposed assimilation scheme was tested on the Lorenz-96 model and 2D Shallow Water Equation model, both of which are associated with spatially correlated observational systems. The experiments showed that by introducing the proposed structure of the forecast error covariance matrix and applying its adaptive estimation procedure, the assimilation results were further improved.
    摘要: Correctly estimating the forecast error covariance matrix is a key step in any data assimilation scheme. If it is not correctly estimated, the assimilated states could be far from the true states. A popular method to address this problem is error covariance matrix inflation. That is, to multiply the forecast error covariance matrix by an appropriate factor. In this paper, analysis states are used to construct the forecast error covariance matrix and an adaptive estimation procedure associated with the error covariance matrix inflation technique is developed. The proposed assimilation scheme was tested on the Lorenz-96 model and 2D Shallow Water Equation model, both of which are associated with spatially correlated observational systems. The experiments showed that by introducing the proposed structure of the forecast error covariance matrix and applying its adaptive estimation procedure, the assimilation results were further improved.
  • 加载中
  • Anderson J. L.,S. L. Anderson, 1999: A Monte Carlo implementation of the non-linear filtering problem to produce ensemble assimilations and forecasts.Mon. Wea. Rev., 127, 2741-2758.
    Anderson J. L.,2007: An adaptive covariance inflation error correction algorithm for ensemble filters. Tellus A, 59, 210-224.
    Anderson J. L.,2009: Spatially and temporally varying adaptive covariance inflation for ensemble filters.Tellus A, 61, 72-83.
    Bai Y. L.,X. Li, 2011: Evolutionary algorithm-based error parameterization methods for data assimilation.Mon. Wea. Rev., 139, 2668-2685.
    Burgers G.,P. J. van Leeuwen,G. Evensen, 1998: Analysis scheme in the ensemble Kalman Filter.Mon. Wea. Rev., 126, 1719-1724.
    Butcher J. C.,2003: Numerical Methods for Ordinary Differential Equations. John Wiley & Sons, 440pp.
    Constantinescu M.,A. Sandu,T. Chai,G. R. Carmichael, 2007: Ensemble-based chemical data assimilation. I: General approach.Quart. J. Roy. Meteor. Soc., 133, 1229-1243.
    Dee D. P.,1995: On-line estimation of error covariance parameters for atmospheric data assimilation.Mon. Wea. Rev., 123, 1128-1145.
    Dee D. P.,A. M. da Silva, 1999: Maximum-likelihood estimation of forecast and observation error covariance parameters. Part I: Methodology.Mon. Wea. Rev., 127, 1822-1834.
    Dee D. P.,G. Gaspari,C. Redder,L. Rukhovets,A. M. da Silva, 1999: Maximum-likelihood estimation of forecast and observation error covariance parameters. Part II: Applications.Mon. Wea. Rev., 127, 1835-1849.
    Evensen G.,1994a: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics.J. Geophys. Res., 99, 10143-10162.
    Evensen G.,1994b: Inverse methods and data assimilation in nonlinear ocean models.Physica D, 77, 108-129.
    Evensen G.,2003: The ensemble Kalman filter: Theoretical formulation and practical implementation.Ocean Dynamics, 53, 343-367.
    Golub G. H.,C. F. Van Loan, 1996: Matrix Computations. The Johns Hopkins Univ, 694pp.
    Hamill T. M.,J. S. Whitaker, 2005: Accounting for the error due to unresolved scales in ensemble data assimilation: A comparison of different approaches.Mon. Wea. Rev., 133, 3132-3147.
    Houtekamer P. L.,H. L. Mitchell, 2001: A sequential ensemble Kalman filter for atmospheric data assimilation.Mon. Wea. Rev., 129, 123-137.
    Ide K.,P. Courtier,G. Michael,A. C. Lorenc, 1997: Unified notation for data assimilation: Operational, sequential and variational.J. Meteor. Soc. Japan, 75, 181-189.
    Lei L. L.,D. R. Stauffer, 2009: A hybrid ensemble Kalman filter approach to data assimilation in a two-dimensional shallow water model. Proc. 23rd Conference on Weather Analysis and Forecasting/19th Conference on Numerical Weather Prediction, Omaha, NE, American Meteorological Society, 9A.4.
    Li H.,E. Kalnay,T. Miyoshi, 2009: Simultaneous estimation of covariance inflation and observation errors within an ensemble Kalman filter.Quart. J. Roy. Meteor. Soc., 135, 523-533.
    Liang X.,X. G. Zheng,S. P. Zhang,G. C. Wu,Y. J. Dai,Y. Li, 2012: Maximum likelihood estimation of inflation factors on forecast error covariance matrix for ensemble Kalman filter assimilation.Quart. J. Roy. Meteor. Soc., 138, 263-273.
    Liu C. S.,Q. N. Xiao,B. Wang, 2008: An ensemble-based four-dimensional variational data assimilation scheme. Part I: Technical formulation and preliminary test.Mon. Wea. Rev., 136, 3363-3373.
    Lorenz E. N.,1996: Predictability—A problem partly solved. Proc. Seminar on Predictability, 40-58, Shinfield Park, Reading, Berkshire, U. K., ECMWF.
    Lorenz E. N.,K. A. Emanuel, 1998: Optimal sites for supplementary weather observations: Simulation with a small model.J. Atmos. Sci., 55, 399-414.
    Miyoshi T.,2011: The Gaussian approach to adaptive covariance inflation and its implementation with the local ensemble transform Kalman filter.Mon. Wea. Rev., 139, 1519-1535.
    Reichle R. H.,2008: Data assimilation methods in the earth sciences.Advances in Water Resources, 31, 1411-1418.
    Talagrand O.,1997: Assimilation of observations, an introduction.J. Meteor. Soc. Japan, 75(1B), 191-209.
    Tian X. J.,Z. H. Xie,A. G. Dai, 2011: A POD-based ensemble four-dimensional variational assimilation method.Tellus A, 63, 805-816.
    Tippett M. K.,J. L. Anderson,C. H. Bishop,T. M. Hamill,J. S. Whitaker, 2003: Notes and correspondence: Ensemble Square Root Filters.Mon. Wea. Rev., 131, 1485-1490.
    Wang X.,C.H. Bishop, 2003: A comparison of breeding and ensemble transform Kalman filter ensemble forecast schemes.J. Atmos. Sci., 60, 1140-1158.
    Zheng X. G.,2009: An adaptive estimation of forecast error covariance parameters for Kalman filtering data assimilation.Adv. Atmos. Sci., 26, 154-160, doi: 10.1007/s00376-009-0154-5.
  • [1] Xiaogu ZHENG, 2009: An Adaptive Estimation of Forecast Error Covariance Parameters for Kalman Filtering Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 154-160.  doi: 10.1007/s00376-009-0154-5
    [2] Fuqing ZHANG, Meng ZHANG, James A. HANSEN, 2009: Coupling Ensemble Kalman Filter with Four-dimensional Variational Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1-8.  doi: 10.1007/s00376-009-0001-8
    [3] LIU Zhengyu, WU Shu, ZHANG Shaoqing, LIU Yun, RONG Xinyao, , 2013: Ensemble Data Assimilation in a Simple Coupled Climate Model: The Role of Ocean-Atmosphere Interaction, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1235-1248.  doi: 10.1007/s00376-013-2268-z
    [4] Zhaoxia PU, Joshua HACKER, 2009: Ensemble-based Kalman Filters in Strongly Nonlinear Dynamics, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 373-380.  doi: 10.1007/s00376-009-0373-9
    [5] Jian YUE, Zhiyong MENG, Cheng-Ku YU, Lin-Wen CHENG, 2017: Impact of Coastal Radar Observability on the Forecast of the Track and Rainfall of Typhoon Morakot (2009) Using WRF-based Ensemble Kalman Filter Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 66-78.  doi: 10.1007/s00376-016-6028-8
    [6] Youmin TANG, Jaison AMBANDAN, Dake CHEN, , , 2014: Nonlinear Measurement Function in the Ensemble Kalman Filter, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 551-558.  doi: 10.1007/s00376-013-3117-9
    [7] Yong LI, Siming LI, Yao SHENG, Luheng WANG, 2018: Data Assimilation Method Based on the Constraints of Confidence Region, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 334-345.  doi: 10.1007/s00376-017-7045-y
    [8] Xinrong WU, Shaoqing ZHANG, Zhengyu LIU, 2016: Implementation of a One-Dimensional Enthalpy Sea-Ice Model in a Simple Pycnocline Prediction Model for Sea-Ice Data Assimilation Studies, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 193-207.  doi: 10.1007/s00376-015-5099-2
    [9] Banglin ZHANG, Vijay TALLAPRAGADA, Fuzhong WENG, Jason SIPPEL, Zaizhong MA, 2016: Estimation and Correction of Model Bias in the NASA/GMAO GEOS5 Data Assimilation System: Sequential Implementation, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 659-672.  doi: 10.1007/ s00376-015-5155-y
    [10] Lili LEI, Yangjinxi GE, Zhe-Min TAN, Yi ZHANG, Kekuan CHU, Xin QIU, Qifeng QIAN, 2022: Evaluation of a Regional Ensemble Data Assimilation System for Typhoon Prediction, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1816-1832.  doi: 10.1007/s00376-022-1444-4
    [11] FU Weiwei, ZHU Jiang, ZHOU Guangqing, WANG Huijun, 2005: A Comparison Study of Tropical Pacific Ocean State Estimation: Low-Resolution Assimilation vs. High-Resolution Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 212-219.  doi: 10.1007/BF02918510
    [12] FU Weiwei, ZHOU Guangqing, WANG Huijun, 2004: Ocean Data Assimilation with Background Error Covariance Derived from OGCM Outputs, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 181-192.  doi: 10.1007/BF02915704
    [13] LIU Juanjuan, WANG Bin, WANG Shudong, 2010: The Structure of Background-error Covariance in a Four-dimensional Variational Data Assimilation System: Single-point Experiment, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1303-1310.  doi: 10.1007/s00376-010-9067-6
    [14] Bohua Huang, James L. Kinter III, Paul S. Schopf, 2002: Ocean Data Assimilation Using Intermittent Analyses and Continuous Model Error Correction, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 965-992.  doi: 10.1007/s00376-002-0059-z
    [15] ZHANG Shuwen, LI Haorui, ZHANG Weidong, QIU Chongjian, LI Xin, 2005: Estimating the Soil Moisture Profile by Assimilating Near-Surface Observations with the Ensemble Kalman Filter (EnKF), ADVANCES IN ATMOSPHERIC SCIENCES, 22, 936-945.  doi: 10.1007/BF02918692
    [16] BAI Yulong, LI Xin, and HUANG Chunlin, 2013: Handling error propagation in sequential data assimilation using an evolutionary strategy, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1096-1105.  doi: 10.1007/s00376-012-2115-7
    [17] GAO Feng*, Peter P. CHILDS, Xiang-Yu HUANG, Neil A. JACOBS, and Jinzhong MIN, 2014: A Relocation-based Initialization Scheme to Improve Track-forecasting of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 27-36.  doi: 10.1007/s00376-013-2254-5
    [18] Tao SUN, Yaodeng CHEN, Deming MENG, Haiqin CHEN, 2021: Background Error Covariance Statistics of Hydrometeor Control Variables Based on Gaussian Transform, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 831-844.  doi: 10.1007/s00376-021-0271-3
    [19] Zhiyong MENG, Eugene E. CLOTHIAUX, 2022: Contributions of Fuqing ZHANG to Predictability, Data Assimilation, and Dynamics of High Impact Weather: A Tribute, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 676-683.  doi: 10.1007/s00376-021-1362-x
    [20] Linbin He, Weiyi Peng, Yu Zhang, Shiguang Miao, Siqi Chen, Jiajing Li, Duanzhou Shao, Xutao Zhang, 2024: Comparison of Adaptive Simulation Observation Experiments of the Heavy Rainfall in South China and Sichuan Basin, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-024-3114-1

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 21 June 2012
Manuscript revised: 17 September 2012
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Using Analysis State to Construct a Forecast Error Covariance Matrix in Ensemble Kalman Filter Assimilation

    Corresponding author: WU Guocan; 
    Corresponding author: ZHANG Shupeng, gcwu@bnu.edu.cn
  • 1. College of Global Change and Earth System Science, Beijing Normal University, Beijing 100875; 
  • 2. National Meteorological Information Center, China Meteorological Administration, Beijing 100081; 
  • 3. School of Mathematical Sciences, Beijing Normal University, Beijing 100875
Fund Project:  This work was supported by the National Program on Key Basic Research Project of China (Grant No. 2010CB950703), the National Natural Science foundation of China General Program (Grant No. 40975062) and the Young Scholars Fundation of Beijing Normal University (Grant No. 105502GK). The authors gratefully acknowledge the two anonymous reviewers for their constructive and relevant comments, which helped in improving the quality of this manuscript.

Abstract: Correctly estimating the forecast error covariance matrix is a key step in any data assimilation scheme. If it is not correctly estimated, the assimilated states could be far from the true states. A popular method to address this problem is error covariance matrix inflation. That is, to multiply the forecast error covariance matrix by an appropriate factor. In this paper, analysis states are used to construct the forecast error covariance matrix and an adaptive estimation procedure associated with the error covariance matrix inflation technique is developed. The proposed assimilation scheme was tested on the Lorenz-96 model and 2D Shallow Water Equation model, both of which are associated with spatially correlated observational systems. The experiments showed that by introducing the proposed structure of the forecast error covariance matrix and applying its adaptive estimation procedure, the assimilation results were further improved.

摘要: Correctly estimating the forecast error covariance matrix is a key step in any data assimilation scheme. If it is not correctly estimated, the assimilated states could be far from the true states. A popular method to address this problem is error covariance matrix inflation. That is, to multiply the forecast error covariance matrix by an appropriate factor. In this paper, analysis states are used to construct the forecast error covariance matrix and an adaptive estimation procedure associated with the error covariance matrix inflation technique is developed. The proposed assimilation scheme was tested on the Lorenz-96 model and 2D Shallow Water Equation model, both of which are associated with spatially correlated observational systems. The experiments showed that by introducing the proposed structure of the forecast error covariance matrix and applying its adaptive estimation procedure, the assimilation results were further improved.

1 Introduction
  • Data assimilation is a method to produce an optimal description of the system state based on model prediction and observations (Talagrand, 1997). Accurately estimating the forecast error covariance matrix is one of the most important steps in data assimilation. If it is correctly known, the analysis states can be generated by minimizing the objective function, which is a fairly technical aspect and can be accomplished with existing engineering solutions (Reichle, 2008).

    Ensemble Kalman filter (EnKF) is a popular data assimilation scheme, which has been widely used in atmospheric and oceanic fields since it was first introduced by Evensen (1994a, b). In EnKF, the forecast error covariance matrix is estimated as the sampling covariance matrix of the ensemble forecast states. However, past works on EnKF have found that the sampling errors in such estimation, resulting from the limited ensemble size, as well as the poor initial perturbations and model error, can generally lead to an underestimate of the forecast error covariance matrix and eventually result in filter divergence (e.g. Anderson and Anderson, 1999; Constantinescu et al., 2007).

    To compensate for this defect, using forecast error covariance matrix inflation to increase ensemble variance is increasingly important. One of the inflation techniques is additive inflation, in which a noise is added to the ensemble forecast states that samples the probability distribution of model error (Hamill and Whitaker, 2005). Another widely used forecast error covariance matrix inflation technique is multiplicative inflation; that is, to multiply the matrix by an appropriate factor.

    At the early state, the inflation factor is estimated empirically. (Wang and Bishop, 2003) developed an approach to optimize inflation factor. Their scheme was further improved by (Li et al., 2009). All these schemes are based on first moment estimation of squared observation-minus-forecast residual, which was first introduced to data assimilation by (Dee, 1995). (Zheng, 2009) and (Liang et al., 2012) proposed another approach to optimize the inflation factor, but using maximum likelihood estimation (MLE), which was introduced by Dee and colleagues (Dee and da Silva, 1999; Dee et al., 1999). Anderson (2007, 2009) also proposed an inflation method based on a Bayesian approach, and (Miyoshi, 2011) further simplified Anderson's inflation approach by making a number of additional simplifying assumptions. However, their study was limited to the case that observational error is spatially independent.

    In this paper, analysis states are used to construct the forecast error covariance matrix (which is different from the sampling covariance matrix of ensemble forecast states used in conventional EnKF). The methodology documented by (Liang et al., 2012) is further developed by introducing an adaptive procedure for estimating the proposed forecast error covariance matrix. By definition, the ensemble forecast error should be represented by ensemble forecast states minus true states rather than by ensemble forecast states minus the ensemble mean of the forecast states (Evensen, 2003). For the model with large error, the ensemble mean of forecast states can be far from the true states. Therefore, the sampling covariance matrix of the ensemble forecast states can be far from the true forecast error covariance matrix. As a result, the estimated analysis state cannot be accurate enough. In reality, it is impossible to know the true state exactly, but the analysis state is a better estimation of the true state than the forecast state. Therefore, in this paper, we propose to use the information feedback from analysis state to update the forecast error covariance matrix. In fact, our proposed forecast error covariance matrix is a combination of multiplicative and additive inflation. Bai and Li (2011) also used the feedback from the analysis state to improve assimilation, but in a different way.

    The remainder of the paper is organized as follows. Section 2 summarizes the EnKF scheme with the MLE inflation scheme of the forecast error covariance matrix (Zheng, 2009; Liang et al., 2012), as well as our proposed structure of the forecast error covariance matrix and its adaptive estimation procedure. Sections 3 and 4 respectively present the assimilation results from the low-order Lorenz model (Lorenz, 1996) and the more realistic 2D Shallow Water Equation (SWE) model with a larger correlated observational system. In particular, our approach is compared with two published forecast error covariance matrix inflation techniques: the MLE inflation scheme (Liang et al., 2012), and the first moment inflation scheme (Wang and Bishop, 2003). Finally, a discussion and conclusions are given in section 5.

2 Methodology
  • Using the notations of (Ide et al., 1997), a nonlinear discrete-time forecast and linear observational system is written as:

    xt,i= Mi-1(xa,i-1)+ηi, (1)

    Yo,i=HiXt,ii, (2)

    where i is the time step index; xt,i ={ xt,i (1), xt,i (2),…, xt,i (n)}T is the n-dimensional true state vector at the time step i; xa,,i-1 ={ xa,i-1 (1), xa,i-1 (2),…, xa,i-1 (n)}T is the n-dimensional analysis state vector, which is an estimate of xt,i-1; Mi is the nonlinear forecast operator, such as a weather forecast model; yo,I is the observational vector with dimension Pi; Hi (linear in this paper) is the measurement matrix of dimension Pi×n that maps the model states to the observed variables; ηi andεi are the forecast error and observational error vectors, which are assumed to be statistically independent of each other, time-uncorrelated, and have zero mean and covariance matrices i and Ri respectively. The goal of EnKF assimilation is to find a series of analysis states xa,i that is sufficiently close to the true state xt,i, using information provided by Mi and ya,i.

    It is well known that any EnKF assimilation scheme should include a forecast error inflation scheme; otherwise, the EnKF may diverge (Anderson and Anderson, 1999). In this paper, a procedure for estimating the multiplicative inflation factor of Pi is proposed based on the MLE principle (e.g. Dee and da Silva, 1999; Zheng, 2009; Liang et al., 2012). The basic filter algorithm in this paper uses perturbed observations (Burgers et al., 1998) without localization (Houtekamer and Mitchell, 2001). The estimation steps of this algorithm equipped with MLE inflation (Zheng, 2009; Liang et al., 2012) are as follows.

    Step 1: Calculate the perturbed forecast states

    Xf,I,j = Mi-1(xa,i-1,j) (3)

    where xa,i-1,j is the perturbed analysis states derived at the previous time step (1jm and m is the number of ensemble members). Define the forecast state xf,I to be the ensemble mean of xf,I,j.

    and then inflate it to the form of . The inflation factorλi is estimated by minimizing the objective function

    where is the observation-minus-forecast residual and Lj(λ) is the 2log-likelihood of dj. The relatively efficient calculations for the determinant and inverse in Eq. (5) are documented in Appendix A, as well as in (Liang et al., 2012).

    Step 3: Compute the perturbed analysis states

    where is the estimated inflation factor andε’i,j is normally distributed with mean zero and covariance matrix Ri. Then, define the estimate of the analysis state xa,i to be the ensemble mean of xa,I,j. Finally, set i=i+1 and return to Step 1.

  • By Eqs. (1) and (3), the ensemble forecast error is xf,i,j - xt,i. xf,i is an estimate of xt,i without knowing observations. The ensemble forecast error is initially estimated as xf,i,j - xt,i, which is used to construct the forecast error covariance matrix in section 2.1. However, owing to limited ensemble size and model error, xt,i can be biased. Therefore, xf,i,j - xt,i can be a biased estimate of xf,i,j - xt,i.

    In this paper, we propose to use observations for improving the estimation of ensemble forecast error. The idea is as follows: After analysis state xa,i is derived, it is generally a better estimate of xt,i than the forecast state xt,i. So xt,i in Eq. (4) is substituted by xa,i for updating the forecast error covariance matrix. This procedure is repeated iteratively until the corresponding objective function [Eq. (5)] converges. For the computational details, Step 2 in section 2.1 is modified to the following adaptive procedure:Step 2a: Use Step 2 in section 2.1 to inflate the initial forecast error covariance matrix to . Then, use Step 3 in section 2.1 to estimate the initial analysis state and set .

    Step 2b: Update the forecast error covariance matrix as

    Then, adjust the forecast error covariance matrices to , where is estimated by minimizing the following objective function

    If , where δ is a predetermined threshold to control the convergence of Eq. (8), then estimate the kth updated analysis state as

    set k=k+1 and return to Eq. (7); otherwise, accept as the estimated forecast error covariance matrix at the ith time step, and go to Step 3 in section 2.1.

    A flowchart of our proposed assimilation scheme is shown in Fig. 1. Moreover, our proposed forecast error covariance matrix [Eq. (7)] can be expressed as

    which is a multiplicatively inflated sampling error covariance matrix plus an additive inflation matrix (see Appendix B for the proof).

    Figure 1.  Flowchart of the proposed assimilation scheme.

  • For ideal models, the "true" state xt,i can be calculated. So, we can use the RMSE of the analysis state to evaluate the accuracy of the assimilation results. The RMSE at the ith step is defined as

    where ||.|| represents the Euclidean norm and n is the dimension of the state vector. In principle, a smaller RMSE means a better performance of the assimilation scheme.

3 Experiment on the Lorenz-96 model
  • In this section we apply our proposed data assimilation schemes to a nonlinear dynamical system with properties relevant to realistic forecast problems: the Lorenz-96 model (Lorenz, 1996) with model error and linear observational system. We evaluate the performances of the covariance inflation methods described in section 2 through the following experiments.

  • The Lorenz-96 model (Lorenz, 1996) is a strongly nonlinear dynamical system with quadratic nonlinearity, governed by the equation

    where k=1,2,…,40. To make Eq. (12) meaningful for all values of k, we define r-1= r39, r0= r40, r41= r1. The dynamics of Eq. (12) are "atmosphere-like" in that the three terms on the right-hand side consist of a nonlinear advection-like term, a damping term, and an external forcing term respectively. These terms can be thought of as some atmospheric quantity (e.g. zonal wind speed) distributed on a latitude circle.

    In our assimilation schemes, we set F=8. For this parameter setting, the leading Lyapunov exponent implies an error-doubling time of about eight time steps, and the fractal dimension of the attractor is 27.1 (Lorenz and Emanuel, 1998). The initial condition is chosen to be rk=F when k≠20 and r20 =1.001F. The initial ensemble perturbation states are generated by adding a white noise with standard deviation F/40 to the initial value. We solve Eq. (12) using a fourth-order Runge-Kutta time integration scheme (Butcher, 2003) with a time step of 0.05 nondimensional units to derive the true state. This is roughly equivalent to six hours in real time, assuming that the characteristic time scale of the dissipation in the atmosphere is five days (Lorenz, 1996).

    In this study, we assume the synthetic observations are generated at every model grid point by adding random noises which are multivariate-normally distributed with zero mean and covariance matrix Ri to the true state. In our experiments, the observational errors are spatially correlated, which may potentially be the case for remote sensing observations and radiances data. The leading-diagonal elements of Ri areσ20 =1 and the off-diagonal elements at site pair (j,k) are

    We would like to introduce model error in the Lorenz-96 model, because model error is inevitable in real dynamic systems. Thus, we chose different values of F in our assimilation schemes, while retaining F=8 when generating the "true" state. We assimilate observations every four time steps for 2000 steps and select the ensemble size as 30. The predetermined thresholdδto control the convergence of Eq. (8) is set to be one.

4 Results in the 2D SWE model
  • In this section we further test our proposed adaptive assimilation scheme using a larger model, the 2D SWE model (see Liang et al., 2012).

  • : The SWEs used in the present study are the equations of motion that can be used to describe the horizontal structure of an atmosphere. They describe the evolution of an incompressible fluid in response to gravitational and rotational accelerations. The barotropic nonlinear SWE consists of the following partial differential equations (Lei and Stauffer, 2009):

    Where μ is the fluid velocity in the first direction (or zonal velocity); νis the fluid velocity in the second direction (or meridional velocity); h is the fluid height (or depth); g is the acceleration due to gravity, which is set to 9.8 m s-2; f is the Coriolis coefficient associated with the Coriolis force, defined as the constant 10-4 s-1 using the f -plane approximation; and the diffusion coefficientκis given as 104 m2 s-1. Further, L and D are the length and width of the rectangular domain of integration, which are set to 500 km and 300 km respectively. The rectangular domain is divided into 50x31 dimensions using a uniform grid spacing of 10 km in the two directions. The time step is 30 seconds and the Lax-Wendroff method is used to integrate the model with respect to time. The periodic boundary condition is used at the first direction boundaries, and a free-slip rigid wall boundary condition (where μ and h are defined from the values one point inside the boundary) is used at the second direction boundaries. See (Liang et al., 2012) for a more detailed explanation.

    where H0, H1 and H2 are set to 50.0 m, 5.5 m and 3.325 m respectively. The initial ensemble height values are created by adding a white noise with unit variance to the initial height. The initial velocity field and the ensemble members are derived from the initial height fields with the geostrophic relation.

    The observational errors are specified asσ2oh =0.5 m2 s-2, σ2oh =0.5 m2 s-2, σ2oh=1.0 m2, and are spatially correlated with each other. The correlation coefficient between the grid points zk and zk’ is ρd(zk,zk’), where d(zk,zk’) is the distance between the two grid points and the parameter ρ is set to 0.5. Furthermore, the observational errors of different variables are assumed to be independent of each other. We use the diffusion coefficient k=5x10-4m2 s-1 to generate model error in our assimilation process. The ensemble size is 100. The predetermined threshold δ is also set to be one.

    Figure 4.  The same as Fig. 3, but for the SWE model.

  • We integrate SWE for 72 hours and the results are considered as the "true" states and have synthetic observations of μ, ν and h at randomly located 310 grid points for every 3 hours in a 48-hour assimilation period. A 24-hour forecast is taken after the 48-hour assimilation period. The time series of analysis RMSE of componentsμ, ν and h of the three assimilation schemes in the 72-hour period are plotted in Fig. 4. Since the first 24-hour period is treated as spin-up and the last 24-hour period is for validation of prediction, we only calculate the time-mean values between 25-hour and 48-hour. The results are shown in Table 2.

    Figure 4 and Table 2 show that the time-mean RMSE of EnKF with our proposed structure of forecast error covariance matrix is smaller than the EnKF with the MLE inflation scheme and with the first moment inflation scheme, especially for the component ν.

5 Discussion and conclusions
  • In all data assimilation schemes, the analysis state is estimated as a weighted average of model forecasts and observation. The weights are related to the corresponding error covariance matrices. Therefore, the performance of data assimilation depends crucially on how accurately the forecast and observational error covariance matrices can be estimated. An assimilation scheme that can estimate them more accurately should have a better assimilation result.

    In fact, the true forecast error should be represented as the ensemble forecast states minus the true state. Since in real problems the true state is not available, we use the ensemble mean of the forecast states instead. Consequently, the forecast error covariance matrix is represented as the sampling covariance matrix of the ensemble forecast states. However, if the model error is large, the ensemble mean of the forecast states may be far from the true state. In this case, the estimated forecast error covariance matrix will remain far from the truth, no matter which inflation technique is used.

    For verifying the aforementioned point, EnKF with error covariance matrix inflation was applied to the Lorenz-96 model and SWE model, but the forecast state xf,i in the forecast error covariance matrix [Eq. (4)] was substituted by the true state xt,i. The corresponding RMSEs are also shown in Figs. 2-4 and Tables 1-2. All the figures and tables show that the RMSE was very significantly reduced compared with the MLE inflation scheme and the first moment inflation scheme.

    However, since the true state xt,i is not known, we used the analysis state xa,i to substitute the forecast state xf,i, because xa,i is closer to xt,i than xf,i. To achieve this goal, a proposed structure of the forecast error covariance matrix and an adaptive estimation procedure for estimating the new structure have been introduced here to constantly improve the assimilation result. As shown in sections 3 and 4, the RMSEs of corresponding analysis states are indeed smaller than those of the other two schemes.

    The -2log-likelihood function of observation-minus-forecast seems a good objective function to quantify the goodness-of-fit of the forecast error covariance matrix. In most cases in this study, the number of iterations was only 3 or 4 and the objective function decreased sharply. In fact, as shown in Tables 1-2, the smaller -2log-likelihood always corresponded to a smaller RMSE of the analysis state. On the other hand, the improved forecast error covariance matrix indeed leads to the improvement of analysis state. All the figures and tables also show that analysis RMSE of EnKF with the MLE inflation scheme is smaller than that with the first moment inflation scheme. This may not be generally true, but could be case dependent.

    Similar to other EnKF assimilation with single parameter inflation schemes, this study also assumed the inflation factor is constant in space. Apparently, this is not the case in many practical applications, especially when the observations are unevenly distributed. Persistently applying the same inflation values that are reasonably large to address problems in densely observed areas to all state variables can systematically overinflate the ensemble variances in sparsely observed areas (Hamill,2005; Anderson,2009). Even if the adaptive procedure for estimation of the forecast error covariance matrix is applied, the problem may still exist to some extent. In the two case studies conducted in this paper, the observational systems were relatively evenly distributed.

    From the assimilation results shown in Fig. 2, the improvement is minor when there is no model error or model error is small (e.g. F=8,7,9). The improvement is more significant when model error is larger (e.g. F=8,12). Since the model error in numeric weather prediction (NWP) models is generally not very large (otherwise, the 4DVar scheme will not work), we suggest that the proposed scheme may not be able to significantly improve NWP data assimilation. However, we expect the proposed scheme may be more useful to improve hydrological data assimilation and land surface data assimilation, since the model errors in hydrological models and land surface models are generally larger.

    In future work, we will investigate how to modify the adaptive procedure to suit a system with unevenly distributed observations. We also plan to use our proposed scheme to improve the estimation of the forecast error covariance matrix in the data assimilation schemes developed by our colleagues (e.g. Liu et al., 2008; Tian et al., 2011; Bai and Li, 2011). In this way, we hope that data assimilation capabilities in China will be considerably improved.

  • Efficient calculations of and

    Provided that Ri-1 is inexpensive to obtain, : can be calculated using the Sherman-Morrison-Woodbury identity (Golub and van Loan, 1996), as is proposed in (Tippett et al., 2003):

    where I is the pi×pj identity matrix, and the n×m matrix Xf,i is the square root of :

    For ensemble filters, Xf,i can be obtained through the deviations of ensemble forecast states to their mean:

    In this way, only m×m matrices need to be inverted and thus the computational cost is significantly reduced.

    can be calculated efficiently using the singular value decomposition of the matrix :

    where Ri-1/2 is the square root of Ri-1, Ui and Vi are orthogonal matrices; Di is a pi×m matrix whose leading-diagonal elements are the singular values of Ri-1/2HiXf,i. Then, we have:

    where di,j is the jth diagonal element of Di.

  • In fact,

    (B1)

    Since xf,i is the ensemble mean forecast, we have

    (B2)

    and similarly

    (B3)

    So the last two terms of Eq. (B1) vanish. Therefore

    (B4)

    In this sense, what we proposed is a scheme combining a multiplicative inflation and an additive inflation.

  • Suppose that the forecast errors and observational errors are uncorrelated, the following relationship is satisfied:

    where <> represents the mathematical expectation operator. The vector di in this operator is regarded as a random vector.

    As is stated in (Wang and Bishop, 2003), for the global observational networks the number of independent scalar elements within Ri-1/2di is large, and thus distributes closely around its mean value . With this assumption, Eq. (C1) becomes:

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint