Advanced Search
Article Contents

Comparison of Nonlinear Local Lyapunov Vectors with Bred Vectors, Random Perturbations and Ensemble Transform Kalman Filter Strategies in a Barotropic Model


doi: 10.1007/s00376-016-6003-4

  • The breeding method has been widely used to generate ensemble perturbations in ensemble forecasting due to its simple concept and low computational cost. This method produces the fastest growing perturbation modes to catch the growing components in analysis errors. However, the bred vectors (BVs) are evolved on the same dynamical flow, which may increase the dependence of perturbations. In contrast, the nonlinear local Lyapunov vector (NLLV) scheme generates flow-dependent perturbations as in the breeding method, but regularly conducts the Gram-Schmidt reorthonormalization processes on the perturbations. The resulting NLLVs span the fast-growing perturbation subspace efficiently, and thus may grasp more components in analysis errors than the BVs. In this paper, the NLLVs are employed to generate initial ensemble perturbations in a barotropic quasi-geostrophic model. The performances of the ensemble forecasts of the NLLV method are systematically compared to those of the random perturbation (RP) technique, and the BV method, as well as its improved version——the ensemble transform Kalman filter (ETKF) method. The results demonstrate that the RP technique has the worst performance in ensemble forecasts, which indicates the importance of a flow-dependent initialization scheme. The ensemble perturbation subspaces of the NLLV and ETKF methods are preliminarily shown to catch similar components of analysis errors, which exceed that of the BVs. However, the NLLV scheme demonstrates slightly higher ensemble forecast skill than the ETKF scheme. In addition, the NLLV scheme involves a significantly simpler algorithm and less computation time than the ETKF method, and both demonstrate better ensemble forecast skill than the BV scheme.
  • 加载中
  • Anderson J. L., S. Anderson, 1999: A Monte Carlo implementation of the nonlinear filtering problem to produce ensemble assimilations and forecasts. Mon. Wea. Rev., 127, 2741- 2758.10.1175/1520-0493(1999)1272.0.CO;27055e7f275587995641eee2766ef174fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999MWRv..127.2741Ahttp://adsabs.harvard.edu/abs/1999MWRv..127.2741ANot Available
    Barkmeijer J., 1992: Local error growth in a barotropic model. Tellus A, 44, 314- 323.10.1034/j.1600-0870.1992.t01-3-00003.x8649328541d811287afd9fe2b6c14f1bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1034%2Fj.1600-0870.1992.t01-3-00003.x%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1034/j.1600-0870.1992.t01-3-00003.x/abstractABSTRACT The concept of local error growth is considered within the context of the non-viscous barotropic vorticity equation. Using the adjoint equation of the tangent linear barotropic vorticity equation, we determine error patterns in the initial state that result in the maximum linear error growth at some pre-chosen geographical area and prediction time. For some solutions of the nonlinear barotropic vorticity equation, we investigate to what extent error growth depends on the geographical position of the area considered. It appears that for short prediction times (<3days), the barotropic growth of kinetic energy of initial perturbations is exponential plus a substantial transient effect due to the non-orthogonality of the eigenmodes of the tangent linear operator with respect to the kinetic energy inner product. The latter mechanism initially causes very rapid perturbation growth, even for neutral flows.
    Benettin G., L. Galgani, A. Giorgilli, and J. M. Strelcyn, 1980: Lyapunov characteristic exponents for smooth dynamical systems and for Hamiltonian systems; a method for computing all of them. Part 1: Theory. Meccanica, 15, 9- 20.10.1007/BF02128237db41e98c-0da1-4c52-b2cf-23ee28e503d5c375b5d107dc0d1180d8ef1a5a457fdchttp%3A%2F%2Flink.springer.com%2F10.1007%2Fbf02128237refpaperuri:(3cc90f68f01107a05a62349e29a08786)http://link.springer.com/10.1007/bf02128237Not Available
    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, 420- 436.24610b9bb142a4812e0dd6e3b446b34chttp%3A%2F%2Fnsr.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.1175%2F1520-0493%282001%291292.0.CO%3B2%26link_type%3DDOIhttp://nsr.oxfordjournals.org/external-ref?access_num=10.1175/1520-0493(2001)1292.0.CO;2&amp;link_type=DOI
    Bowler N. E., 2006: Comparison of error breeding, singular vectors, random perturbations and ensemble Kalman filter perturbation strategies on a simple model. Tellus A, 58, 538- 548.10.1111/j.1600-0870.2006.00197.x5ce8a377-8e2f-4e3d-8104-0ece17c58e4dd598fbc96ab13e338f9c45e1fc026536http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1600-0870.2006.00197.x%2Fcitedbyrefpaperuri:(4a9b947e82663979db74e48166973349)http://onlinelibrary.wiley.com/doi/10.1111/j.1600-0870.2006.00197.x/citedbyABSTRACT An experiment has been performed, using a simple chaotic model, to compare different ensemble perturbation strategies. The model used is a 300 variable Lorenz 95 model which displays many of the characteristics of atmospheric numerical weather prediction models. Twenty member ensembles were generated using five perturbation strategies, error breeding, singular vectors, random perturbations (RPs), the Ensemble Kalman Filter (EnKF) and the Ensemble Transform Kalman Filter (ETKF). Based on normal verification methods, such as rank histograms and spread of the perturbations, the RPs method performs as well as any other methodhis illustrates the limitations of using a simple model. Consideration of the quality of the background error information provided by the ensemble gives a better assessment of the ensemble skill. This measure indicates that the EnKF performs best, with the ETKF combined with RPs being the next most skillful. It was found that neither the ETKF, error breeding nor singular vectors provided useful background information on their own.Central to the success of the EnKF is the localization of the background error covariance which removes spurious long-range correlations within the ensemble. Computationally efficient versions of the EnKF (such as the ETKF) cannot accommodate covariance localization and their performance is seen to suffer. Applying the ETKF to a series of local domains has been tested, which allows covariance localization whilst remaining computational efficient, and this has been found to be nearly as effective as the EnKF with covariance localization.
    Buizza R., P. L. Houtekamer, Z. Toth, G. Pellerin, M. Z, Wei, and Y. J. Zhu, 2005: A comparison of the ECMWF, MSC, and NCEP global ensemble prediction systems. Mon. Wea. Rev., 133, 1076- 1097.d6ae0c6ecff394896dd321deb3e191abhttp%3A%2F%2Fbioscience.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.1175%2FMWR2905.1%26link_type%3DDOIhttp://bioscience.oxfordjournals.org/external-ref?access_num=10.1175/MWR2905.1&amp;link_type=DOI
    Carrassi A., A. Trevisan, and F. Uboldi, 2007: Adaptive observations and assimilation in the unstable subspace by breeding on the data assimilation system. Tellus A, 59( 2), 101- 113.10.1111/j.1600-0870.2006.00210.x8187068327f4c0d19c1019f197958bf6http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1600-0870.2006.00210.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/j.1600-0870.2006.00210.x/citedbyResults of targeting and assimilation experiments in a quasi-geostrophic atmospheric model are presented and discussed. The basic idea is to assimilate observations in the unstable subspace (AUS) of the data-assimilation system. The unstable subspace is estimated by breeding on the data-assimilation system (BDAS). The analysis update has the same local structure as the observationally fozrced bred modes.
    Cheung K. K. W., J. C. L. Chan, 1999: Ensemble forecasting of tropical cyclone motion using a barotropic model. Part I: perturbations of the environment. Mon. Wea. Rev., 127, 1229- 1243.
    Descamps L., O. Talagrand, 2007: On some aspects of the definition of initial conditions for ensemble prediction. Mon. Wea. Rev., 135, 3260- 3272.10.1175/MWR3452.1fe879b2caf3a0f47038475377af0a3a5http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007MWRv..135.3260Dhttp://adsabs.harvard.edu/abs/2007MWRv..135.3260DFour methods for initialization of ensemble forecasts are systematically compared, namely the methods of singular vectors (SV) and bred modes (BM), as well as the ensemble Kalman filter (EnKF) and the ensemble transform Kalman filter (ETKF). The comparison is done on synthetic data with two models of the flow, namely, a low-order model introduced by Lorenz and a three-level quasigeostrophic atmospheric model. For the latter, both cases of a perfect and an imperfect model are considered. The performance of the various initialization methods is assessed in terms of the statistical reliability and resolution of the ensuing predictions. The relative performance of the four methods, which is statistically significant to a range of about 6 days, is in the order EnKF > ETKF > BM > SV. The difference between the former two methods and the latter two is on the whole more significant than the differences between EnKF and ETKF, or between BM and SV separately. The general conclusion is that, if the quality of ensemble predictions is assessed by the degree to which the predicted ensembles statistically sample the uncertainty on the future state of the flow, the best initial ensembles are those that best statistically sample the uncertainty on the present state of the flow.
    Ding R. Q., J. P. Li, 2007: Nonlinear finite-time Lyapunov exponent and predictability. Physics Letters A, 364, 396- 400.10.1016/j.physleta.2006.11.094c52a97486acf5c7f7ec380454abc48ffhttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS037596010601944Xhttp://www.sciencedirect.com/science/article/pii/S037596010601944XABSTRACT In this Letter, we introduce a definition of the nonlinear finite-time Lyapunov exponent (FTLE), which is a nonlinear generalization to the existing local or finite-time Lyapunov exponents. With the nonlinear FTLE and its derivatives, the limit of dynamic predictability in large classes of chaotic systems can be efficiently and quantitatively determined.
    Ding R. Q., J. P. Li, and K.-J. Ha, 2008: Trends and interdecadal changes of weather predictability during 1950s-1990s. J. Geophys. Res., 113,D24112, doi: 10.1029/2008JD010404.10.1029/2008JD0104044e99cb4e725f7eaa1cf0c879ab0d4cf8http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008JD010404%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2008JD010404/citedby[1] To study the atmospheric predictability from the view of nonlinear error growth dynamics, a new approach using the nonlinear local Lyapunov exponent (NLLE) is introduced by the authors recently. In this paper, the trends and interdecadal changes of weather predictability limit (WPL) during 1950s1990s are investigated by employing the NLLE approach. The results show that there exist significant trend changes for WPL over most of the globe. At three different pressure levels in the troposphere (850, 500, and 200 hPa), spatial distribution patterns of linear trend coefficients of WPL are similar. Significant decreasing trends in WPL could be found in the most regions of the northern midlatitudes and Africa, while significant increasing trends in WPL lie in the most regions of the tropical Pacific and southern mid-high latitudes. In the lower stratosphere (50 hPa), the WPL in the whole tropics all shows significant increasing trends, while it displays significant decreasing trends in the most regions of the Antarctic and northern mid-high latitudes. By examining the temporal variations of WPL in detail, we find that the interdecadal changes of WPL in most regions at different levels mainly happen in the 1970s, which is consistent with the significant climate shift occurring in the late 1970s. Trends and interdecadal changes of WPL are found to be well related to those of atmospheric persistence, which in turn are linked to the changes of atmospheric internal dynamics. Further analysis indicates that the changes of atmospheric static stability due to global warming might be one of main causes responsible for the trends and interdecadal changes of atmospheric persistence and predictability in the southern and northern mid-high latitudes. The increased sea surface temperature (SST) variability exerts a stronger external forcing on the tropical Pacific atmosphere that tends to enhance the persistence of tropical Pacific atmosphere. This process appears to be responsible for the increase of atmospheric predictability and persistence in the tropical Pacific since the late 1970s.
    Ding R. Q., J. P. Li, and K.-H. Seo, 2010: Predictability of the Madden-Julian oscillation estimated using observational data. Mon. Wea. Rev., 138, 1004- 1013.10.1175/2009MWR3082.1a63f25a7340ba12998736beda92a0aa8http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2010MWRv..138.1004Dhttp://adsabs.harvard.edu/abs/2010MWRv..138.1004DNot Available
    Ding R. Q., J. P. Li, and K.-H. Seo, 2011: Estimate of the predictability of boreal summer and winter intraseasonal oscillations from observations. Mon. Wea. Rev., 139, 2421- 2438.10.1175/2011MWR3571.1cfa82d458df50569729aae32846f7285http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2011MWRv..139.2421Dhttp://adsabs.harvard.edu/abs/2011MWRv..139.2421DNot Available
    Duan W. S., Z. H. Huo, 2016: An approach to generating mutually independent initial perturbations for ensemble forecasts: orthogonal conditional nonlinear optimal perturbations. J. Atmos. Sci.,73, 997-1014, doi: 10.1175/JAS-D-15-0138.1.
    Durran D. R., M. Gingrich, 2014: Atmospheric predictability: Why butterflies are not of practical importance.J. Atmos. Sci., 71, 2476- 2488.10.1175/JAS-D-14-0007.19622df5f1e1f43a0700833fd52981251http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014JAtS...71.2476Dhttp://adsabs.harvard.edu/abs/2014JAtS...71.2476DNot Available
    Epstein E. S., 1969: Stochastic dynamic prediction. Tellus, 21, 739- 759.10.3402/tellusa.v21i6.1014346ad0ed3-01d9-4bec-a167-43471f2ad241283fb6357090d5876a15eea2b10f272bhttp://dx.doi.org/10.1111/j.2153-3490.1969.tb00483.xhttp://dx.doi.org/10.1111/j.2153-3490.1969.tb00483.xThe question of analysis to obtain the initial stochastic statement of the atmospheric state is considered and one finds here too a promise of significant advantages over present deterministic methods. It is shown how the stochastic method can be used to assess the value of new or improved data by considering their influence on the decrease in the uncertainty of the forecast. Comparisons among physical-numerical models are also made more effectively by applying stochastic methods. Finally the implications of stochastic dynamic prediction on the question of predictability are briefly considered, with the conclusion that some earlier estimates have been too pessimistic.
    Evensen G., 2003: The ensemble Kalman filter: theoretical formulation and practical implementation. Ocean Dynamics, 53, 343- 367.10.1007/s10236-003-0036-9fe06c7adae408eb1888b539a799f9cdahttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs10236-003-0036-9http://link.springer.com/10.1007/s10236-003-0036-9The purpose of this paper is to provide a comprehensive presentation and interpretation of the Ensemble Kalman Filter (EnKF) and its numerical implementation. The EnKF has a large user group, and numerous publications have discussed applications and theoretical aspects of it. This paper reviews the important results from these studies and also presents new ideas and alternative interpretations which further explain the success of the EnKF. In addition to providing the theoretical framework needed for using the EnKF, there is also a focus on the algorithmic formulation and optimal numerical implementation. A program listing is given for some of the key subroutines. The paper also touches upon specific issues such as the use of nonlinear measurements, in situ profiles of temperature and salinity, and data which are available with high frequency in time. An ensemble based optimal interpolation (EnOI) scheme is presented as a cost-effective approach which may serve as an alternative to the EnKF in some applications. A fairly extensive discussion is devoted to the use of time correlated model errors and the estimation of model bias.
    Evensen G., 2004: Sampling strategies and square root analysis schemes for the EnKF. Ocean Dynamics, 54, 539- 560.10.1007/s10236-004-0099-22cc4dfccac3482af78528fe442c005bahttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs10236-004-0099-2http://link.springer.com/10.1007/s10236-004-0099-2The purpose of this paper is to examine how different sampling strategies and implementations of the analysis scheme influence the quality of the results in the EnKF. It is shown that by selecting the initial ensemble, the model noise and the measurement perturbations wisely, it is possible to achieve a significant improvement in the EnKF results, using the same number of members in the ensemble. The results are also compared with a square root implementation of the EnKF analysis scheme where the analyzed ensemble is computed without the perturbation of measurements. It is shown that the measurement perturbations introduce sampling errors which can be reduced using improved sampling schemes in the standard EnKF or fully eliminated when the square root analysis algorithm is used. Further, a new computationally efficient square root algorithm is proposed which allows for the use of a low-rank representation of the measurement error covariance matrix. It is shown that this algorithm in fact solves the full problem at a low cost without introducing any new approximations.
    Feng J., R. Q. Ding, D. Q. Liu, and J. P. Li, 2014: The application of nonlinear local Lyapunov vectors to ensemble predictions in the Lorenz systems.J. Atmos. Sci., 71( 10), 3554- 3567.10.1175/JAS-D-13-0270.1fc4c34ac6a236575e4326e2c3b97a50chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014EGUGA..1616378Fhttp://adsabs.harvard.edu/abs/2014EGUGA..1616378FNonlinear local Lyapunov vectors (NLLVs) are developed to indicate orthogonal directions in phase space with different perturbation growth rates. In particular, the first few NLLVs are considered to be an appropriate orthogonal basis for the fast-growing subspace. In this paper, the NLLV method is used to generate initial perturbations and implement ensemble forecasts in simple nonlinear models (the Lorenz63 and Lorenz96 models) to explore the validity of the NLLV method. The performance of the NLLV method is compared comprehensively and systematically with other methods such as the bred vector (BV) and the random perturbation (Monte Carlo) methods. In experiments using the Lorenz63 model, the leading NLLV (LNLLV) captured a more precise direction, and with a faster growth rate, than any individual bred vector. It may be the larger projection on fastest-growing analysis errors that causes the improved performance of the new method. Regarding the Lorenz96 model, two practical measures, namely the spread/skill relationship and the Brier score, were used to assess the reliability and resolution of these ensemble schemes. Overall, the ensemble spread of NLLVs is more consistent with the errors of the ensemble mean, which indicates the better performance of NLLVs in simulating the evolution of analysis errors. In addition, the NLLVs perform significantly better than the BVs in terms of reliability and the random perturbations in resolution.
    Fraedrich K., 1987: Estimating weather and climate predictability on attractors.J. Atmos. Sci., 44, 722- 728.10.1175/1520-0469(1987)044<0722:EWACPO>2.0.CO;20564d40859f80495b7f3f603fd36b4cehttp%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D888056http://www.ams.org/mathscinet-getitem?mr=888056Abstract Predictability is deduced from phase space trajectories of weather and climate variables which evolve on attractors (local surface pressure and a δ 18 O-record). Predictability can be defined by the divergence of initially close pieces of trajectories and estimated by the cumulative distance distributions of expanding pairs of points on the single variable trajectory. The e -folding expansion rates characterize predictability tune scales. As a first estimate one obtains a predictability time scale of about two weeks for the weather variable and 10–15 thousand years for the climate variable.
    Gaspari G., S. E. Cohn, 1999: Construction of correlation functions in two and three dimensions.Quart. J. Roy. Meteor. Soc., 125, 723- 757.10.1002/qj.497125554176bd81eb7b6838de7499add8663f840e7http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49712555417%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49712555417/fullAbstract This article focuses on the construction, directly in physical space, of simply parametrized covariance functions for data-assimilation applications. A self-contained, rigorous mathematical summary of relevant topics from correlation theory is provided as a foundation for this construction. Covariance and correlation functions are defined, and common notions of homogeneity and isotropy are clarified. Classical results are stated, and proven where instructive. Included are smoothness properties relevant to multivariate statistical-analysis algorithms where wind/wind and wind/mass correlation models are obtained by differentiating the correlation model of a mass variable. the Convolution Theorem is introduced as the primary tool used to construct classes of covariance and cross-covariance functions on three-dimensional Euclidean space R 3 . Among these are classes of compactly supported functions that restrict to covariance and cross-covariance functions on the unit sphere S 2 , and that vanish identically on subsets of positive measure on S 2 . It is shown that these covariance and cross-covariance functions on S 2 , referred to as being space-limited , cannot be obtained using truncated spectral expansions. Compactly supported and space-limited covariance functions determine sparse covariance matrices when evaluated on a grid, thereby easing computational burdens in atmospheric data-analysis algorithms. Convolution integrals leading to practical examples of compactly supported covariance and cross-covariance functions on R 3 are reduced and evaluated. More specifically, suppose that gi and gj are radially symmetric functions defined on R 3 such that gi (x) = 0 for |x| > di and gj (x) = 0 for |xv > dj , O < di,dj ≦, where |. | denotes Euclidean distance in R 3 . the parameters di and dj are ‘cut-off’ distances. Closed-form expressions are determined for classes of convolution cross-covariance functions Cij (x,y) := ( gi * gj )(x-y), i ≠ j , and convolution covariance functions Cii (x,y) := ( gi * gi )(x-y), vanishing for |x - y| > di + dj and |x - y| > 2 di , respectively, Additional covariance functions on R 3 are constructed using convolutions over the real numbers R , rather than R 3 . Families of compactly supported approximants to standard second- and third-order autoregressive functions are constructed as illustrative examples. Compactly supported covariance functions of the form C (x,y) := Co (|x - y|), x,y ∈ R 3 , where the functions Co ( r ) for r ∈ R are 5th-order piecewise rational functions, are also constructed. These functions are used to develop space-limited product covariance functions B (x, y) C (x, y), x, y ∈ S 2 , approximating given covariance functions B (x, y) supported on all of S 2 × S 2 .
    Houtekamer P. L., H. L. Mitchell, 1998: Data assimilation using an ensemble Kalman filter technique. Mon. Wea. Rev., 126, 796- 811.10.1175/1520-0493(1998)126<0796:DAUAEK>2.0.CO;2d3f8c1aa877d5eac5d73e53c0fdac200http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1998MWRv..126..796Hhttp://adsabs.harvard.edu/abs/1998MWRv..126..796HNot Available
    Houtekamer P. L., H. L. Mitchell, 2001: A sequential ensemble Kalman filter for atmospheric data assimilation. Mon. Wea. Rev., 129, 123- 137.20195ce27846d743db66fe3982faf1b7http%3A%2F%2Fbiomet.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.1175%2F1520-0493%282001%291292.0.CO%3B2%26link_type%3DDOIhttp://biomet.oxfordjournals.org/external-ref?access_num=10.1175/1520-0493(2001)1292.0.CO;2&amp;link_type=DOI
    Kalnay E., 2002: Atmospheric Modeling,Data Assimilation and Predictability. Cambridge University Press, 221 pp.10.1198/tech.2005.s32617d7fbb1594d0cade45315e4ab692760http%3A%2F%2Fwww.jstor.org%2Fstable%2F25471085http://www.jstor.org/stable/25471085CiteSeerX - Document Details (Isaac Councill, Lee Giles, Pradeep Teregowda): Numerical weather prediction (NWP) now provides major guidance in our daily weather forecast. The accuracy of NWP models has improved steadily since the first successful experiment made by Charney, Fj!rtoft and von Neuman (1950). During the past 50 years, a large number of technical papers and reports have been devoted to NWP, but the number of textbooks dealing with the subject has been very small, the latest being the 1980 book by Haltiner & Williams, which was dedicated to descriptions of the atmospheric dynamics and numerical methods for atmospheric modeling. However, in the intervening years much impressive progress has been made in all aspects of NWP, including the success in model initialization and ensemble forecasts. Eugenia Kalnay recent book covers for the first time in the long history of NWP, not only methods for numerical modeling, but also the important related areas of data assimilation and predictability. It incorporates all aspects of environmental computer modeling including an historical overview of NWP, equations of motion and their approximations, a modern description of the methods to determine the initial conditions using weather observations and a clear discussion of chaos in dynamic systems and how these concepts can be
    Lawrence A. R., M. Leutbecher, and T. N. Palmer, 2009: The characteristics of Hessian singular vectors using an advanced data assimilation scheme.Quart. J. Roy. Meteor. Soc., 135, 1117- 1132.10.1002/qj.447eb65b6c81077bf94a8119707b7521a83http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.447%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/qj.447/citedbyNot Available
    Leith C. E., 1974: Theoretical skill of Monte Carlo forecasts. Mon. Wea. Rev., 102, 409- 418.10.1175/1520-0493(1974)102<0409:TSOMCF>2.0.CO;286f6231bdfd8ee5b602d3893586fda21http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1974MWRv..102..409Lhttp://adsabs.harvard.edu/abs/1974MWRv..102..409LNot Available
    Li J. P., J. F. Chou, 1997: Existence of the atmosphere attractor. Science in China Series D: Earth Sciences, 40( 3), 215- 220.10.1007/BF028783812190ea244556cc025d10cc855908ad41http%3A%2F%2Flink.springer.com%2F10.1007%2FBF02878381http://www.cnki.com.cn/Article/CJFDTotal-JDXG199702012.htm The global asymptotic behavior of solutions for the equations of large-scale atmospheric motion with the non-stationary external forcing is studied in the infinite-dimensional Hilbert space. Based on the properties of operators of the equations, some energy inequalities and the uniqueness theorem of solutions are obtained. On the assumption that external forces are bounded, the exsitence of the global absorbing set and the atmosphere attractor is proved, and the characteristics of the decay of effect of initial field and the adjustment to the external forcing are revealed. The physical sense of the results is discussed and some ideas about climatic numerical forecast are elucidated.
    Li J. P., S. H. Wang, 2008: Some mathematical and numerical issues in geophysical fluid dynamics and climate dynamics. Communications in Computational Physics, 3, 759- 793.10.1007/978-1-4614-8963-4_59057e2d6-02ce-45d3-924c-e5444189700fa48dc7c8a5503b67727eff1b53683535http%3A%2F%2Fsourcedb.cas.cn%2Fsourcedb_iap_cas%2Fen%2Fpapers%2F200908%2Ft20090831_2459201.htmlrefpaperuri:(3061fd2c074e4ad9caba9a4235258fb0)http://sourcedb.cas.cn/sourcedb_iap_cas/en/papers/200908/t20090831_2459201.htmlIn this article, we address both recent advances and open questions in some mathematical and computational issues in geophysical fluid dynamics (GFD) and climate dynamics. The main focus is on 1) the primitive equations (PEs) models and their related mathematical and computational issues, 2) climate variability, predictability and successive bifurcation, and 3) a new dynamical systems theory and its applications to GFD and climate dynamics.
    Li J. P., R. Q. Ding, 2011: Temporal-spatial distribution of atmospheric predictability limit by local dynamical analogs.Mon. Wea. Rev., 139, 3265- 3283.10.1175/MWR-D-10-05020.194414680e18eb59458f83a2bb8f359achttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2011mwrv..139.3265lhttp://adsabs.harvard.edu/abs/2011mwrv..139.3265lAbstract To quantify the predictability limit of a chaotic system, the authors recently developed a method using the nonlinear local Lyapunov exponent (NLLE). The NLLE method provides a measure of local predictability limit of chaotic systems and is intended to supplement existing predictability methods. To apply the NLLE in studies of actual atmospheric predictability, an algorithm based on local dynamical analogs is devised to enable the estimation of the NLLE and its derivatives using experimental or observational data. Two examples are given to illustrate the effectiveness of the algorithm, involving the Lorenz63 three-variable model and the Lorenz96 forty-variable model; they reveal that the algorithm is applicable in estimating the NLLE of a chaotic system from its experimental time series. On this basis, the NLLE method is used to investigate temporalpatial distributions of predictability limits of the daily geopotential height and wind fields. The limit of atmospheric predictability varies widely with region, altitude, and season. The predictability limits of the daily geopotential height and wind fields are generally less than 3 weeks in the troposphere, whereas they are approximately 1 month in the lower stratosphere, revealing a potential predictability source for forecasting weather from the stratosphere. Further work is required to examine broader applications of the NLLE method in predictability studies of the atmosphere, ocean, and other systems.
    Li J. P., R. Q. Ding, 2013: Temporal-spatial distribution of the predictability limit of monthly sea surface temperature in the global oceans. Int. J. Climatol., 33, 1936- 1947.10.1002/joc.3562d591093830f02d95321208ab398e296ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.3562%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/joc.3562/pdfNot Available
    Li J. P., R. Q. Ding, and B. H. Chen, 2006: Review and Prospect on the Predictability Study of the Atmosphere. Review and Prospects of the Developments of Atmosphere Sciences in Early 21st Century. China Meteorology Press, 96- 104. (in Chinese)
    Lorenz E. N., 1965: A study of the predictability of a 28-variable atmospheric model. Tellus, 17, 321- 333.10.1111/j.2153-3490.1965.tb01424.x5290b716f057ec79a3c6ace9c6fd2bd7http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1965.tb01424.x%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1965.tb01424.x/abstractNot Available
    Magnusson L., E. Källèn, and J. Nycander, 2008: Initial state perturbations in ensemble forecasting. Nonlinear Processes in Geophysics, 15, 751- 759.10.5194/npg-15-751-2008c6dae7ce4c145a3213c8d22e0ca31101http%3A%2F%2Fwww.oalib.com%2Fpaper%2F1377277http://www.oalib.com/paper/1377277Due to the chaotic nature of atmospheric dynamics, numerical weather prediction systems are sensitive to errors in the initial conditions. To estimate the forecast uncertainty, forecast centres produce ensemble forecasts based on perturbed initial conditions. How to optimally perturb the initial conditions remains an open question and different methods are in use. One is the singular vector (SV) method, adapted by ECMWF, and another is the breeding vector (BV) method (previously used by NCEP). In this study we compare the two methods with a modified version of breeding vectors in a low-order dynamical system (Lorenz-63). We calculate the Empirical Orthogonal Functions (EOF) of the subspace spanned by the breeding vectors to obtain an orthogonal set of initial perturbations for the model. We will also use Normal Mode perturbations. Evaluating the results, we focus on the fastest growth of a perturbation. The results show a large improvement for the BV-EOF perturbations compared to the non-orthogonalised BV. The BV-EOF technique also shows a larger perturbation growth than the SVs of this system, except for short time-scales. The highest growth rate is found for the second BV-EOF for the long-time scale. The differences between orthogonal and non-orthogonal breeding vectors are also investigated using the ECMWF IFS-model. These results confirm the results from the Loernz-63 model regarding the dependency on orthogonalisation.
    Magnusson L., J. Nycand er, and E. Källèn, 2009: Flow-dependent versus flow-independent initial perturbations for ensemble prediction. Tellus A, 61( 3), 194- 209.10.1111/j.1600-0870.2008.00385.x218f02584b9ae2a70174ff58b538b89fhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1600-0870.2008.00385.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/j.1600-0870.2008.00385.x/citedbyEnsemble prediction relies on a faithful representation of initial uncertainties in a forecasting system. Early research on initial perturbation methods tested random perturbations by adding `white noise' to the analysis. Here, an alternative kind of random perturbations is introduced by using the difference between two randomly chosen atmospheric states (i.e. analyses). It yields perturbations (random field, RF, perturbations) in approximate flow balance.
    Molteni F., R. Buizza, T. N. Palmer, and T. Petroliagis, 1996: The new ECMWF ensemble prediction system: methodology and validation.Quart. J. Roy. Meteor. Soc., 122, 73- 119.10.1002/qj.49712252905facc8001-a3a5-48a8-94e5-fa14103400a306e58ba7360249148550625e940bddcehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49712252905%2Ffullrefpaperuri:(d38acaf5dc0dec476fd404c53c2f7bb0)http://onlinelibrary.wiley.com/doi/10.1002/qj.49712252905/fullThe ECMWF ensemble prediction system: methodology and validation. MOLTENI F. Quart. J. Roy. Meteor. Soc. 122, 73-119, 1996
    Mu M., Z. Y. Zhang, 2006: Conditional nonlinear optimal perturbations of a two-dimensional quasigeostrophic model.J. Atmos. Sci., 63, 1587- 1604.10.1002/qj.256bc158660b9c89cbf8721e7e8ca0260dehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2006JAtS...63.1587Mhttp://adsabs.harvard.edu/abs/2006JAtS...63.1587MCiteSeerX - Scientific documents that cite the following paper: Conditional nonlinear optimal perturbations of a two-dimensional quasigeostrophic model
    Mu M., Z. N. Jiang, 2008: A new approach to the generation of initial perturbations for ensemble prediction: Conditional nonlinear optimal perturbation. Chinese Science Bulletin, 53, 2062- 2068.10.1007/s11434-008-0272-y4b262054db4a6d12088d0a3f072c4570http%3A%2F%2Flink.springer.com%2F10.1007%2Fs11434-008-0272-yhttp://www.cnki.com.cn/Article/CJFDTotal-JXTW200813024.htm
    Murphy A. H., 1973: A new vector partition of the probability score. J. Appl. Meteor., 12, 595- 600.98fc6802469e1f61885d283bfbac4f0chttp%3A%2F%2Fwww.emeraldinsight.com%2Fservlet%2Flinkout%3Fsuffix%3Db28%26dbid%3D16%26doi%3D10.1108%252FEJM-05-2012-0288%26key%3D10.1175%252F1520-0450%281973%290122.0.CO%253B2http://xueshu.baidu.com/s?wd=paperuri%3A%28e7073650750802c7824e422f3b610106%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.emeraldinsight.com%2Fservlet%2Flinkout%3Fsuffix%3Db28%26dbid%3D16%26doi%3D10.1108%252FEJM-05-2012-0288%26key%3D10.1175%252F1520-0450%281973%290122.0.CO%253B2&ie=utf-8&sc_us=10008998936017562578
    Palatella L., A. Trevisan, 2015: Interaction of Lyapunov vectors in the formulation of the nonlinear extension of the Kalman filter. Physical Review E, 91, 042905.10.1103/PhysRevE.91.042905c199aee48dac355c6321809562dd75bchttp%3A%2F%2Feuropepmc.org%2Fabstract%2FMED%2F25974560http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM25974560When applied to strongly nonlinear chaotic dynamics the extended Kalman filter (EKF) is prone to divergence due to the difficulty of correctly forecasting the forecast error probability density function. In operational forecasting applications ensemble Kalman filters circumvent this problem with empirical procedures such as covariance inflation. This paper presents an extension of the EKF that includes nonlinear terms in the evolution of the forecast error estimate. This is achieved starting from a particular square-root implementation of the EKF with assimilation confined in the unstable subspace (EKF-AUS), that is, the span of the Lyapunov vectors with non-negative exponents. When the error evolution is nonlinear, the space where it is confined is no more restricted to the unstable and neutral subspace causing filter divergence. The algorithm presented here, denominated EKF-AUS-NL, includes the nonlinear terms in the error dynamics: These result from the nonlinear interaction among the leading Lyapunov vectors and account for all directions where the error growth may take place. Numerical results show that with the nonlinear terms included, filter divergence can be avoided. We test the algorithm on the Lorenz96 model, showing very promising results.
    Pe\na M., Z. Toth, 2014: Estimation of analysis and forecast error variances. Tellus A, 66, 21767.10.3402/tellusa.v66.21767dcbd6bb39469335359a2cf0c8f32515fhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F269711012_Estimation_of_analysis_and_forecast_error_varianceshttp://www.researchgate.net/publication/269711012_Estimation_of_analysis_and_forecast_error_variancesAccurate estimates of error variances in numerical analyses and forecasts (i.e. difference between analysis or forecast fields and nature on the resolved scales) are critical for the evaluation of forecasting systems, the tuning of data assimilation (DA) systems and the proper initialisation of ensemble forecasts. Errors in observations and the difficulty in their estimation, the fact that estimates of analysis errors derived via DA schemes, are influenced by the same assumptions as those used to create the analysis fields themselves, and the presumed but unknown correlation between analysis and forecast errors make the problem difficult. In this paper, an approach is introduced for the unbiased estimation of analysis and forecast errors. The method is independent of any assumption or tuning parameter used in DA schemes. The method combines information from differences between forecast and analysis fields (erceived forecast errors) with prior knowledge regarding the time evolution of (1) forecast error variance and (2) correlation between errors in analyses and forecasts. The quality of the error estimates, given the validity of the prior relationships, depends on the sample size of independent measurements of perceived errors. In a simulated forecast environment, the method is demonstrated to reproduce the true analysis and forecast error within predicted error bounds. The method is then applied to forecasts from four leading numerical weather prediction centres to assess the performance of their corresponding DA and modelling systems. Error variance estimates are qualitatively consistent with earlier studies regarding the performance of the forecast systems compared. The estimated correlation between forecast and analysis errors is found to be a useful diagnostic of the performance of observing and DA systems. In case of significant model-related errors, a methodology to decompose initial value and model-related forecast errors is also proposed and successfully demonstrated. Keywords: uncertainty of analysis, forecast verification, estimation methods, data assimilation, ensemble forecasts (Published: 24 November 2014) Citation: Tellus A 2014, 66 , 21767, http://dx.doi.org/10.3402/tellusa.v66.21767
    Stephenson D. B., C. A. S. Coelho, and I. T. Jolliffe, 2008: Two extra components in the Brier Score decomposition. Mon. Wea. Rev., 23, 752- 757.10.1175/2007WAF2006116.189e970acdb5dba5c19ac64d274147bc1http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2008WtFor..23..752Shttp://adsabs.harvard.edu/abs/2008WtFor..23..752SAbstract The Brier score is widely used for the verification of probability forecasts. It also forms the basis of other frequently used probability scores such as the rank probability score. By conditioning (stratifying) on the issued forecast probabilities, the Brier score can be decomposed into the sum of three components: uncertainty, reliability, and resolution. This Brier score decomposition can provide useful information to the forecast provider about how the forecasts can be improved. Rather than stratify on all values of issued probability, it is common practice to calculate the Brier score components by first partitioning the issued probabilities into a small set of bins. This note shows that for such a procedure, an additional two within-bin components are needed in addition to the three traditional components of the Brier score. The two new components can be combined with the resolution component to make a generalized resolution component that is less sensitive to choice of bin width than is the traditional resolution component. The difference between the generalized resolution term and the conventional resolution term also quantifies how forecast skill is degraded when issuing categorized probabilities to users. The ideas are illustrated using an example of multimodel ensemble seasonal forecasts of equatorial sea surface temperatures.
    Toth Z., E. Kalnay, 1993: Ensemble forecasting at NMC: The generation of perturbations. Bull. Amer. Meteor. Soc., 74, 2317- 2330.10.1175/1520-0477(1993)074<2317:EFANTG>2.0.CO;2cbe63b780b83c7603b5d4d22f0d6e97fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1993BAMS...74.2317Thttp://adsabs.harvard.edu/abs/1993BAMS...74.2317TOn 7 December 1992, The National Meteorological Center (NMC) started operational ensemble forecasting. The ensemble forecast configuration implemented provides 14 independent forecasts every day verifying on days 1-10. In this paper we briefly review existing methods for creating perturbations for ensemble forecasting. We point out that a regular analysis cycle is a "breeding ground" for fast-growing modes. Based on this observation, we devise a simple and inexpensive method to generate growing modes of the atmosphere.The new method, "breeding of growing modes", or BGM, consists of one additional, perturbed short-range forecast, introduced on top of the regular analysis in an analysis cycle. The difference between the control and perturbed six-hour (first guess) forecast is scaled back to the size of the initial perturbation and then reintroduced onto the new atmospheric analysis. Thus, the perturbation evolves along with the time dependent analysis fields, ensuring that after a few days of cycling the perturbation field consists of a superposition of fast-growing modes corresponding to the contemporaneous atmosphere, akin to local Lyapunov vectors.The breeding cycle has been designed to model how the growing errors are "bred" and maintained in a conventional analysis cycle through the successive use of short-range forecasts. The bred modes should thus offer a good estimate of possible growing error fields in the analysis. Results from extensive experiments indicate that ensembles of just two BGM forecasts achieve better results than much larger random Monte Cado or lagged average forecast (LAF) ensembles. Therefore, the operational ensemble configuration at NMC is based on the BGM method to generate efficient initial perturbations.The only two methods explicitly designed to generate perturbations that contain fast-growing modes corresponding to the evolving atmosphere are the BGM and the method of Lorenz, which is based on the singular modes of the linear tangent model. This method has been adopted operationally at The European Centre for Medium-Range Forecasts (ECMWF) for ensemble forecasting. Both the BGM and the ECMWF methods seem promising, but since it has not yet been possible to compare in detail their operational performance we limit ourselves to pointing out some of their similarities and differences.
    Toth Z., E. Kalnay, 1997: Ensemble forecasting at NCEP and the breeding method. Mon. Wea. Rev., 125, 3297- 3319.10.1175/1520-0493(1997)125<3297:EFANAT>2.0.CO;2d77eecdcf56bec983b860925dfc56c2dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1997MWRv..125.3297Thttp://adsabs.harvard.edu/abs/1997MWRv..125.3297TPurpose: The aim of this study was to evaluate the effects of curing mode and viscosity on the biaxial flexural strength (FS) and modulus (FM) of dual resin cements. Methods: Eight experimental groups were created (n=12) according to the dual-cured resin cements (Nexus 2/Kerr Corp. and Variolink II/IvoclarVivadent), curing modes (dual or self-cure), and viscosities (low and high). Forty-eight cement discs of each product (0.5 mm thick by 6.0 mm diameter) were fabricated. Half specimens were light--activated for 40 seconds and half were allowed to self-cure. After 10 days, the biaxial flexure test was performed using a universal testing machine (1.27 mm/min, Instron 5844). Data were statistically analyzed by three-way ANOVA and Tukey's test (5%). Results: Light-activation increased FS and FM of resin cements at both viscosities in comparison with self-curing mode. The high viscosity version of light-activated resin cements exhibited higher FS than low viscosity versions. The viscosity of resin and the type of cement did not influence the FM. Light-activation of dual-polymerizing resin cements provided higher FS and FM for both resin cements and viscosities. Conclusion: The use of different resin cements with different viscosities may change the biomechanical behavior of these luting materials.
    Trevisan A., R. Legnani, 1995: Transient error growth and local predictability: A study in the Lorenz system. Tellus A, 47, 103- 117.10.1034/j.1600-0870.1995.00006.xaa6c4400de03ba3d68d42ff0f63180e3http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1034%2Fj.1600-0870.1995.00006.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1034/j.1600-0870.1995.00006.x/citedbyLorenz's three-variable convective model is used as a prototypical chaotic system in order to develop concepts related to finite time local predictability. Local predictability measures can be represented by global measures only if the instability properties of the attractor are homogeneous in phase space. More precisely, there are two sources of variability of predictability in chaotic attractors. The first depends on the direction of the initial error vector, and its dependence is limited to an initial transient period. If the attractor has homogeneous predictability properties, this is the only source of variability of error growth rate and, after the transient has elapsed, all initial perturbations grow at the same rate, given by the first (global) Lyapunov exponent. The second is related to the local instability properties in phase space. If the predictability properties of the attractor are not homogeneous, this additional source of variability affects both the transient and post-transient phases of error growth. After the transient phase all initial perturbations of a particular initial condition grow at the same rate, given in this case by the first local Lyapunov exponent. We consider various currently used indexes to quantify finite time local predictability. The probability distributions of the different indexes are examined during and after the transient phase. By comparing their statistics it is possible to discriminate the relative importance of the two sources of variability of predictability and to determine the most appropriate measure of predictability for a given forecast time. It is found that a necessary premise for choosing a relevant local predictability index for a specific system is the study of the characteristics of its transient. The consequences for the problem of forecasting forecast skill in operational models are discussed.
    Trevisan A., F. Uboldi, 2004: Assimilation of standard and targeted observations within the unstable subspace of the observation-analysis-forecast cycle system.J. Atmos. Sci., 61, 103- 113.10.1175/1520-0469(2004)061<0103:AOSATO>2.0.CO;2b93afd432c2099523d9e3e9221a7221ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004jats...61..103thttp://adsabs.harvard.edu/abs/2004jats...61..103tIn this paper it is shown that the flow-dependent instabilities that develop within an observation analysis forecast (OAF) cycle and that are responsible for the background error can be exploited in a very simple way to assimilate observations. The basic idea is that, in order to minimize the analysis and forecast errors, the analysis increment must be confined to the unstable subspace of the OAF cycle solution. The analysis solution here formally coincides with that of the classical three-dimensional variational solution with the background error covariance matrix estimated in the unstable subspace.The unstable directions of the OAF system solution are obtained by breeding initially random perturbations of the analysis but letting the perturbed trajectories undergo the same process as the control solution, including assimilation of all the available observations. The unstable vectors are then used both to target observations and for the assimilation design.The approach is demonstrated in an idealized environment using a simple model, simulated standard observations over land with a single adaptive observation over the ocean. In the application a simplified form is adopted of the analysis solution and a single unstable vector at each analysis time whose amplitude is determined by means of the adaptive observation. The remarkable reduction of the analysis and forecast error obtained by this simple method suggests that only a few accurately placed observations are sufficient to control the local instabilities that take place along the cycle.The stability of the system, with or without forcing by observations, is studied and the growth rate of the leading instability of the different control solutions is estimated. Whereas the model has more than one positive Lyapunov exponent, the solution of the OAF scheme that includes the adaptive observation is stable. It is suggested that a negative exponent can be considered a necessary condition for the convergence of a particular OAF solution to the truth, and that the estimate of the degree of stability of the control trajectory can be used as a simple criterion to evaluate the efficiency of data assimilation and observation strategies.The present findings are in line with previous quantative observability results with more realistic models and with recent studies that indicate a local low dimensionality of the unstable subspace.
    Trevisan A., L. Palatella, 2011: Chaos and weather forecasting: the role of the unstable subspace in predictability and state estimation problems. International Journal of Bifurcation and Chaos, 21( 12), 3389- 3415.10.1142/S0218127411030635e7f10c82-afec-4421-9d02-b349c29928ebWOS:000300016000002fc774617d6d0c74e56d2eb42e1e6bb35http%3A%2F%2Fwww.worldscientific.com%2Fdoi%2Fabs%2F10.1142%2FS0218127411030635http://www.worldscientific.com/doi/abs/10.1142/S0218127411030635In the first part of this paper, we review some important results on atmospheric predictability, from the pioneering work of Lorenz to recent results with operational forecasting models. Particular relevance is given to the connection between atmospheric predictability and the theory of Lyapunov exponents and vectors. In the second part, we briefly review the foundations of data assimilation methods and then we discuss recent results regarding the application of the tools typical of chaotic systems theory described in the first part to well established data assimilation algorithms, the Extended Kalman Filter (EKF) and Four Dimensional Variational Assimilation (4DVar). In particular, the Assimilation in the Unstable Space (AUS), specifically developed for application to chaotic systems, is described in detail.
    Vannitsem S., C. Nicolis, 1997: Lyapunov vectors and error growth patterns in a T21L3 quasigeostrophic model.J. Atmos. Sci., 54, 347- 361.10.1175/1520-0469(1997)054<0347:LVAEGP>2.0.CO;2c72165e612a8b150a2b386e2a5a169cdhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1997JAtS...54..347Vhttp://adsabs.harvard.edu/abs/1997JAtS...54..347VAbstract The authors report a systematic study on the short and intermediate time predictability properties of a quasigeostrophic T21L3 model in which emphasis is placed on the role of the Lyapunov vectors in the growth patterns of generic initial error fields. It is found that under scale-independent small-amplitude initial perturbations the evolution of the mean error is intimately related to the spectral distribution of the Lyapunov vectors. In the case of perturbations at a particular scale of motion the picture turns out to be more involved, particularly as far as mean error growth over all wavenumbers is concerned, and must appeal to coupling mechanisms between different scales. The role of the norm used for the measure of the mean error growth and the specific predictability properties at different vertical levels of the model are also analyzed.
    Wang X., C. H. Bishop, 2003: A comparison of breeding and ensemble transform Kalman filter ensemble forecast schemes.J. Atmos. Sci., 60, 1140- 1158.10.1175/1520-0469(2003)0602.0.CO;2e944df1a5d6726e5668650fe360597c4http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2003EAEJA.....8087Whttp://adsabs.harvard.edu/abs/2003EAEJA.....8087WAbstract The ensemble transform Kalman filter (ETKF) ensemble forecast scheme is introduced and compared with both a simple and a masked breeding scheme. Instead of directly multiplying each forecast perturbation with a constant or regional rescaling factor as in the simple form of breeding and the masked breeding schemes, the ETKF transforms forecast perturbations into analysis perturbations by multiplying by a transformation matrix. This matrix is chosen to ensure that the ensemble-based analysis error covariance matrix would be equal to the true analysis error covariance if the covariance matrix of the raw forecast perturbations were equal to the true forecast error covariance matrix and the data assimilation scheme were optimal. For small ensembles (100), the computational expense of the ETKF ensemble generation is only slightly greater than that of the masked breeding scheme. Version 3 of the Community Climate Model (CCM3) developed at National Center for Atmospheric Research (NCAR) is used to test and compare these ensemble generation schemes. The NCEPCAR reanalysis data for the boreal summer in 2000 are used for the initialization of the control forecast and the verifications of the ensemble forecasts. The ETKF and masked breeding ensemble variances at the analysis time show reasonable correspondences between variance and observational density. Examination of eigenvalue spectra of ensemble covariance matrices demonstrates that while the ETKF maintains comparable amounts of variance in all orthogonal and uncorrelated directions spanning its ensemble perturbation subspace, both breeding techniques maintain variance in few directions. The growth of the linear combination of ensemble perturbations that maximizes energy growth is computed for each of the ensemble subspaces. The ETKF maximal amplification is found to significantly exceed that of the breeding techniques. The ETKF ensemble mean has lower root-mean-square errors than the mean of the breeding ensemble. New methods to measure the precision of the ensemble-estimated forecast error variance are presented. All of the methods indicate that the ETKF estimates of forecast error variance are considerably more accurate than those of the breeding techniques.
    Wei M. Z., Z. Toth, R. Wobus, Y. J. Zhu, C. H. Bishop, and X. G. Wang, 2006: Ensemble transform Kalman filter-based ensemble perturbations in an operational global prediction system at NCEP. Tellus A, 58, 28- 44.10.1111/j.1600-0870.2006.00159.x9327d5ee3404291fa0c4ad6c9a0f1523http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1600-0870.2006.00159.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/j.1600-0870.2006.00159.x/citedbyThe initial perturbations used for the operational global ensemble prediction system of the National Centers for Environmental Prediction are generated through the breeding method with a regional rescaling mechanism. Limitations of the system include the use of a climatologically fixed estimate of the analysis error variance and the lack of an orthogonalization in the breeding procedure. The Ensemble Transform Kalman Filter (ETKF) method is a natural extension of the concept of breeding and, as shown by Wang and Bishop, can be used to generate ensemble perturbations that can potentially ameliorate these shortcomings. In the present paper, a spherical simplex 10-member ETKF ensemble, using the actual distribution and error characteristics of real-time observations and an innovation-based inflation, is tested and compared with a 5-pair breeding ensemble in an operational environment.
    Wei M. Z., Z. Toth, R. Wobus, and Y. J. Zhu, 2008: Initial perturbations based on the ensemble transform (ET) technique in the NCEP global operational forecast system. Tellus A, 60, 62- 79.10.1111/j.1600-0870.2007.00273.xc630b34e2d19a77224c2d239c5c9c31dhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1600-0870.2007.00273.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/j.1600-0870.2007.00273.x/citedbyABSTRACT Since modern data assimilation (DA) involves the repetitive use of dynamical forecasts, errors in analyses share characteristics of those in short-range forecasts. Initial conditions for an ensemble prediction/forecast system (EPS or EFS) are expected to sample uncertainty in the analysis field. Ensemble forecasts with such initial conditions can therefore (a) be fed back to DA to reduce analysis uncertainty, as well as (b) sample forecast uncertainty related to initial conditions. Optimum performance of both DA and EFS requires a careful choice of initial ensemble perturbations.A can be improved with an EFS that represents the dynamically conditioned part of forecast error covariance as accurately as possible, while an EFS can be improved by initial perturbations reflecting analysis error variance. Initial perturbation generation schemes that dynamically cycle ensemble perturbations reminiscent to how forecast errors are cycled in DA schemes may offer consistency between DA and EFS, and good performance for both. In this paper, we introduce an EFS based on the initial perturbations that are generated by the Ensemble Transform (ET) and ET with rescaling (ETR) methods to achieve this goal. Both ET and ETR are generalizations of the breeding method (BM).he results from ensemble systems based on BM, ET, ETR and the Ensemble Transform Kalman Filter (ETKF) method are experimentally compared in the context of ensemble forecast performance. Initial perturbations are centred around a 3D-VAR analysis, with a variance equal to that of estimated analysis errors. Of the four methods, the ETR method performed best in most probabilistic scores and in terms of the forecast error explained by the perturbations. All methods display very high time consistency between the analysis and forecast perturbations. It is expected that DA performance can be improved by the use of forecast error covariance from a dynamically cycled ensemble either with a variational DA approach (coupled with an ETR generation scheme), or with an ETKF-type DA scheme.
    Wolf A., J. B. Swift, H. L. Swinney, and J. A. Vastano, 1985: Determining Lyapunov exponents from a time series. Physica D, 16, 285- 317.10.1016/0167-2789(85)90011-9baaa5d4fc1835d289fb01a91524965bahttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2F0167278985900119http://www.sciencedirect.com/science/article/pii/0167278985900119We present the first algorithms that allow the estimation of non-negative Lyapunov exponents from an experimental time series. Lyapunov exponents, which provide a qualitative and quantitative characterization of dynamical behavior, are related to the exponentially fast divergence or convergence of nearby orbits in phase space. A system with one or more positive Lyapunov exponents is defined to be chaotic. Our method is rooted conceptually in a previously developed technique that could only be applied to analytically defined model systems: we monitor the long-term growth rate of small volume elements in an attractor. The method is tested on model systems with known Lyapunov spectra, and applied to data for the Belousov-Zhabotinskii reaction and Couette-Taylor flow.
    Yoden S., M. Nomura, 1993: Finite-time Lyapunov stability analysis and its application to atmospheric predictability.J. Atmos. Sci., 50, 1531- 1543.10.1175/1520-0469(1993)0502.0.CO;2acce085a99cc527702f019f7391a3174http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D1220381http://www.ams.org/mathscinet-getitem?mr=1220381Finite-time Lyapunov stability analysis is reviewed and applied to a low-order spectral model of barotropic flow in a midlatitude channel. The tangent linear equations of the model are used to investigate the growth of small perturbations superposed on a reference solution for a prescribed time interval. Three types of reference solutions of the model, stationary, periodic, and chaotic, are investigated to demonstrate usefulness of the analysis in the study of the atmospheric predictability problem.The finite-time Lyapunov exponents, which give the growth rate of small perturbations, depend upon the reference solution as well as the prescribed time interval. The finite-time Lyapunov vector corresponding to the largest Lyapunov exponent gives the streamfunction field of the fastest growing perturbation for the time interval. In the case of the chaotic reference solution, the streamfunction field has large amplitudes in limited areas for a small time interval. The areas of the large perturbation growth have some relation to the reference streamfunction field.A possible application of the finite-time Lyapunov exponents and vectors to the atmospheric predictability problem is discussed. These quantities might be used as several forecast measures of the time-dependent predictability in numerical weather predictions.
    Zhang J., W. S. Duan, and X. F. Zhi, 2015: Using CMIP5 model outputs to investigate the initial errors that cause the "spring predictability barrier" for El Niño events. Science China Earth Sciences, 58( 6), 685- 696.10.1007/s11430-014-4994-1d57fadd1-363a-49e9-91d4-df3dc995c4d63f8214c469d75bc5ff9d90f96d0f82e1http%3A%2F%2Flink.springer.com%2F10.1007%2Fs11430-014-4994-1refpaperuri:(60b1f6a9262208ffa3e1e125bf94f4b1)http://www.cnki.com.cn/Article/CJFDTotal-JDXG201505003.htmMost ocean-atmosphere coupled models have difficulty in predicting the El Ni-Southern Oscillation(ENSO) when starting from the boreal spring season. However, the cause of this spring predictability barrier(SPB) phenomenon remains elusive. We investigated the spatial characteristics of optimal initial errors that cause a significant SPB for El Ni events by using the monthly mean data of the pre-industrial(PI) control runs from several models in CMIP5 experiments. The results indicated that the SPB-related optimal initial errors often present an SST pattern with positive errors in the central-eastern equatorial Pacific, and a subsurface temperature pattern with positive errors in the upper layers of the eastern equatorial Pacific, and negative errors in the lower layers of the western equatorial Pacific. The SPB-related optimal initial errors exhibit a typical La Ni--like evolving mode, ultimately causing a large but negative prediction error of the Ni-3.4 SST anomalies for El Ni events. The negative prediction errors were found to originate from the lower layers of the western equatorial Pacific and then grow to be large in the eastern equatorial Pacific. It is therefore reasonable to suggest that the El Ni predictions may be most sensitive to the initial errors of temperature in the subsurface layers of the western equatorial Pacific and the Ni-3.4 region, thus possibly representing sensitive areas for adaptive observation. That is, if additional observations were to be preferentially deployed in these two regions, it might be possible to avoid large prediction errors for El Ni and generate a better forecast than one based on additional observations targeted elsewhere. Moreover, we also confirmed that the SPB-related optimal initial errors bear a strong resemblance to the optimal precursory disturbance for El Ni and La Ni events. This indicated that improvement of the observation network by additional observations in the identified sensitive areas would also be helpful in detecting the signals provided by the precursory disturbance, which may greatly improve the ENSO prediction skill.
    Ziehmann C., L. A. Smith, and J. Kurths, 2000: Localized Lyapunov exponents and the prediction of predictability. Physics Letters A, 271, 237- 251.10.1016/S0375-9601(00)00336-4aef304e3cb647d0a59ceaec1a98852abhttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0375960100003364http://www.sciencedirect.com/science/article/pii/S0375960100003364Every forecast should include an estimate of its likely accuracy, a current measure of predictability. Two distinct types of localized Lyapunov exponents based on infinitesimal uncertainty dynamics are investigated to reflect this predictability. Regions of high predictability within which any initial uncertainty will decrease are proven to exist in two common chaotic systems; potential implications of these regions are considered. The relevance of these results for finite size uncertainties is discussed and illustrated numerically.
  • [1] Jun Kyung KAY, Hyun Mee KIM, Young-Youn PARK, Joohyung SON, 2013: Effect of Doubling the Ensemble Size on the Performance of Ensemble Prediction in the Warm Season Using MOGREPS Implemented at the KMA, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1287-1302.  doi: 10.1007/s00376-012-2083-y
    [2] Jie FENG, Jianping LI, Jing ZHANG, Deqiang LIU, Ruiqiang DING, 2019: The Relationship between Deterministic and Ensemble Mean Forecast Errors Revealed by Global and Local Attractor Radii, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 271-278.  doi: 10.1007/s00376-018-8123-5
    [3] Xiaqiong ZHOU, Johnny C. L. CHEN, 2006: Ensemble Forecasting of Tropical Cyclone Motion Using a Baroclinic Model, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 342-354.  doi: 10.1007/s00376-006-0342-5
    [4] Ruiqiang DING, Baojia LIU, Bin GU, Jianping LI, Xuan LI, 2019: Predictability of Ensemble Forecasting Estimated Using the Kullback-Leibler Divergence in the Lorenz Model, ADVANCES IN ATMOSPHERIC SCIENCES, , 837-846.  doi: 10.1007/s00376-019-9034-9
    [5] 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
    [6] Xubin ZHANG, 2022: Impacts of New Implementing Strategies for Surface and Model Physics Perturbations in TREPS on Forecasts of Landfalling Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1833-1858.  doi: 10.1007/s00376-021-1222-8
    [7] Wansuo DUAN, Lichao YANG, Mu MU, Bin WANG, Xueshun SHEN, Zhiyong MENG, Ruiqiang DING, 2023: Recent Advances in China on the Predictability of Weather and Climate, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1521-1547.  doi: 10.1007/s00376-023-2334-0
    [8] Xin LIU, Jing CHEN, Yongzhu LIU, Zhenhua HUO, Zhizhen XU, Fajing CHEN, Jing WANG, Yanan MA, Yumeng HAN, 2024: An Initial Perturbation Method for the Multiscale Singular Vector in Global Ensemble Prediction, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 545-563.  doi: 10.1007/s00376-023-3035-4
    [9] WAN Liying, ZHU Jiang, WANG Hui, YAN Changxiang, Laurent BERTINO, 2009: A ``Dressed" Ensemble Kalman Filter Using the Hybrid Coordinate Ocean Model in the Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1042-1052.  doi: 10.1007/s00376-009-7208-6
    [10] ZHANG Shuwen, LI Deqin, QIU Chongjian, 2011: A Multimodel Ensemble-based Kalman Filter for the Retrieval of Soil Moisture Profiles, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 195-206.  doi: 10.1007/s00376-010-9200-6
    [11] 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
    [12] LI Shan, RONG Xingyao, LIU Yun, LIU Zhengyu, Klaus FRAEDRICH, 2013: Dynamic Analogue Initialization for Ensemble Forecasting, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1406-1420.  doi: 10.1007/s00376-012-2244-z
    [13] 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
    [14] ZHENG Xiaogu, WU Guocan, ZHANG Shupeng, LIANG Xiao, DAI Yongjiu, LI Yong, , 2013: Using Analysis State to Construct a Forecast Error Covariance Matrix in Ensemble Kalman Filter Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1303-1312.  doi: 10.1007/s00376-012-2133-5
    [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] CUI Limei, SUN Jianhua, QI Linlin, LEI Ting, 2011: Application of ATOVS Radiance-Bias Correction to Typhoon Track Prediction with Ensemble Kalman Filter Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 178-186.  doi: 10.1007/s00376-010-9145-9
    [17] Kefeng ZHU, Ming XUE, Yujie PAN, Ming HU, Stanley G. BENJAMIN, Stephen S. WEYGANDT, Haidao LIN, 2019: The Impact of Satellite Radiance Data Assimilation within a Frequently Updated Regional Forecast System Using a GSI-based Ensemble Kalman Filter, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1308-1326.  doi: 10.1007/s00376-019-9011-3
    [18] 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
    [19] Yu ZHANG, Yuanfu XIE, Hongli WANG, Dehui CHEN, Zoltan TOTH, 2016: Ensemble Transform Sensitivity Method for Adaptive Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 10-20.  doi: 10.1007/s00376-015-5031-9
    [20] Zhou Jiabin, Wang Yunkuan, Yang Guiying, Wu Jinsheng, 1994: A Forecasting Model of Vector Similarity in Phase Space for Flood and Drought over the Huanghe-Huaihe-Haihe Plain in China, ADVANCES IN ATMOSPHERIC SCIENCES, 11, 224-229.  doi: 10.1007/BF02666548

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 04 January 2016
Manuscript revised: 13 April 2016
Manuscript accepted: 04 May 2016
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

Comparison of Nonlinear Local Lyapunov Vectors with Bred Vectors, Random Perturbations and Ensemble Transform Kalman Filter Strategies in a Barotropic Model

  • 1. State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029
  • 2. College of Earth Science, University of Chinese Academy of Sciences, Beijing 100049
  • 3. Global Systems Division, Earth System Research Laboratory/Oceanic and Atmospheric Research/National Oceanic and Atmospheric Administration, Boulder, CO, 80305, USA
  • 4. Plateau Atmosphere and Environment Key Laboratory of Sichuan Province, Chengdu University of Information Technology, Chengdu 610225
  • 5. College of Global Change and Earth System Sciences, Beijing Normal University, Beijing 100875
  • 6. Joint Center for Global Change Studies, Beijing 100875
  • 7. Fujian Meteorological Observatory, Fuzhou 350001

Abstract: The breeding method has been widely used to generate ensemble perturbations in ensemble forecasting due to its simple concept and low computational cost. This method produces the fastest growing perturbation modes to catch the growing components in analysis errors. However, the bred vectors (BVs) are evolved on the same dynamical flow, which may increase the dependence of perturbations. In contrast, the nonlinear local Lyapunov vector (NLLV) scheme generates flow-dependent perturbations as in the breeding method, but regularly conducts the Gram-Schmidt reorthonormalization processes on the perturbations. The resulting NLLVs span the fast-growing perturbation subspace efficiently, and thus may grasp more components in analysis errors than the BVs. In this paper, the NLLVs are employed to generate initial ensemble perturbations in a barotropic quasi-geostrophic model. The performances of the ensemble forecasts of the NLLV method are systematically compared to those of the random perturbation (RP) technique, and the BV method, as well as its improved version——the ensemble transform Kalman filter (ETKF) method. The results demonstrate that the RP technique has the worst performance in ensemble forecasts, which indicates the importance of a flow-dependent initialization scheme. The ensemble perturbation subspaces of the NLLV and ETKF methods are preliminarily shown to catch similar components of analysis errors, which exceed that of the BVs. However, the NLLV scheme demonstrates slightly higher ensemble forecast skill than the ETKF scheme. In addition, the NLLV scheme involves a significantly simpler algorithm and less computation time than the ETKF method, and both demonstrate better ensemble forecast skill than the BV scheme.

1. Introduction
  • In the past two decades, ensemble forecasting has been substantially developed to become a powerful approach that improves numerical weather prediction (NWP). The basic principle of the generation of initial ensemble members is to sample the uncertainties related to the initial analysis (Epstein, 1969; Leith, 1974). Various ensemble generation schemes based on dynamical error growth theory have been tested and used in weather prediction centers; for example, the bred vector (BV) method (Toth and Kalnay, 1993, 1997) used at NCEP, and the singular vector (SV) method (Lorenz, 1965; Molteni et al., 1996) at ECMWF. Afterwards, data assimilation (DA) schemes were further combined with the dynamical methods to better sample the analysis uncertainties, such as in the ensemble transform Kalman filter (ETKF) scheme (Bishop et al., 2001; Wang and Bishop, 2003; Wei et al., 2006, 2008) which is an improved version of the BV method, and in the introduction of advanced DA schemes in the SV method (Lawrence et al., 2009).

    Numerous studies have been carried out to compare the BV and ETKF approaches (Wang and Bishop, 2003; Bowler, 2006; Wei et al., 2006, 2008; Descamps and Talagrand, 2007), from which is has been found that the better forecast skill of the ETKF scheme than the breeding scheme is mainly associated with two specific features. One is that the ensemble perturbations of the ETKF scheme can better reflect the geographical variations of analysis error variance than those of the breeding scheme. This may be because the BV perturbations are tuned by a time-invariant estimate of analysis error variance ("mask"; Toth and Kalnay, 1997), while the ETKF perturbations at each time are updated through the DA algorithm. The other is that the ETKF perturbations are more independent and uncorrelated than the BVs. The BVs are naturally evolved perturbations that represent the unstable modes associated with the day-to-day dynamical flow. However, dynamical flow would drive the BVs to project on the growing type of directions and thus increase the similarity of BVs, especially in local regions. Such similarities are reduced by the nonlinear error growth in NWP models, but can also be seen by the eigenvalue spectrum for BVs (Wang and Bishop, 2003; Bowler, 2006). In contrast, the ETKF ensemble perturbations are orthogonal in observation space (Wang and Bishop, 2003; Wei et al., 2006).

    In view of the interdependence among BVs, (Feng et al., 2014) developed the nonlinear local Lyapunov vectors (NLLVs), which are the nonlinear extension of the linear Lyapunov vectors (LVs). The traditional LVs define a set of time-dependent orthogonal perturbation structures (Kalnay, 2002; Trevisan and Palatella, 2011). These perturbations point to the directions with different growth rates in phase space specified by positive, null and negative Lyapunov exponents (Benettin et al., 1980; Wolf et al., 1985; Fraedrich, 1987; Trevisan and Legnani, 1995; Vannitsem and Nicolis, 1997). Therefore, the LVs can be used to identify the unstable, neutral and stable subspaces. In reality, on the one hand, the unstable perturbation subspace has a significantly smaller dimension than the stable subspace, and thus is much easier to be sampled (Toth and Kalnay, 1997); while on the other hand, any perturbation, through model integration, will eliminate the non-growing components in it, and project onto the fast-growing subspace with relatively small dimension. Therefore, the identification of unstable perturbations is critical to the improvement of NWP (Zhang et al., 2015), like the accurate estimation of error variances in the background forecasts of DA (Trevisan and Uboldi, 2004; Carrassi et al., 2007), and the generation of ensemble perturbations in ensemble forecasting (Toth and Kalnay, 1997), etc. This is also why the LVs are gradually developed and applied in DA, ensemble generation and other areas in numerical forecasting. The NLLVs bear some similarities to the LVs; for example, they are orthogonal to each other and have different error growth rates from the largest to the smallest. However, they also have significant differences. The LVs are generated through a long-term integration using the tangent linear model and the direction of the leading LV is independent of the initial perturbation (Kalnay, 2002). In comparison, the NLLVs are obtained by integrating the nonlinear model from both the perturbed and unperturbed initial conditions, and represent different types of instabilities with different rescaling amplitudes of perturbations.

    Both the BVs and the NLLVs are nonlinear extensions of LLVs, and can be dynamically produced by a sequence of two nonlinear model integrations with a periodic rescaling of their difference. BVs are random samples in the small fastest growing space, and have consistently large projection on the leading LLV (Magnusson et al., 2008). However, NWP models usually have a multidimensional growing subspace. Various independent fast growing directions should be comprehensively considered to span the unstable subspace (Duan and Huo, 2016). The fastest growing directions captured by the BVs have limited diversity and may not contain sufficient information to represent the development of analysis errors. In contrast, apart from the leading NLLV (LNLLV) being similar to the BVs, which tend to the fastest growing direction, other successively orthogonalized NLLVs point to additional fast growing directions in phase space. The NLLVs have larger diversity than the BVs and may capture a larger amount of components in analysis errors (Feng et al., 2014).

    The NLLV method is related to the ETKF method, both of which are designed to orthogonalize perturbations. In the ETKF approach, the ensemble perturbation matrix is transformed by the matrix T associated with the covariance matrix of forecast members and observations (Wang and Bishop, 2003). In comparison, the NLLVs are orthogonalized using the Gram-Schmidt reorthonormalization (GSR) technique. Given that the ETKF scheme has been successfully implemented in operational forecast centers, it is worthwhile comparing the NLLV approach with the ETKF approach. The results of the BV and the random perturbation (RP) approaches are also given as reference, and all the comparisons are presented in a barotropic model. To clarify the differences between these four initialization methods, the experiments are implemented in a perfect model environment, and by using the same analysis and the same forecast model. An assimilation scheme independent of each ensemble initialization scheme —— namely, the ensemble Kalman filter (EnKF; Houtekamer and Mitchell, 1998; Evensen, 2003, 2004) —— is used to produce the initial analysis states.

    This paper is organized as follows: Section 2 describes the BV, ETKF and NLLV ensemble generation schemes. An introduction to the barotropic model and the basic setup of the numerical experiments is provided in section 3. Section 4 presents the results of the quantitative comparison of the four ensemble methods. Section 5 provides a discussion and conclusions.

2. Initial perturbation schemes
  • The breeding method proposed by Toth and Kalnay (1993, 1997) is used to simulate the error evolution in each analysis cycle. RPs are superposed on the analysis state to grow as a perturbation between two model integrations. At the end of each breeding cycle (12 hours in our experiments), the perturbations are rescaled by an appropriate factor. This, without the insertion of further random noise, defines a new set of perturbations that are used to generate the next perturbed initial states. This process is repeated and, as the effect of the particular initial RP diminishes, the resulting perturbations become the BVs. Initially, (Toth and Kalnay, 1993) used a global rescaling factor that is close to the empirically estimated global analysis RMSE. However, to ensure that the ensemble perturbation magnitude reflects the regionally varying uncertainties in the analysis, the breeding method used at NCEP includes an estimated mask as a constraint of BVs (Toth and Kalnay, 1997). Owing to our experiments having observations on every grid over the whole field and having the same precision, the distribution of analysis error variances is mainly determined by the forecast error variances. Therefore, only a simple global rescaling factor is applied in this study. The BVs are calculated using a Euclidean norm. Through a five-day breeding process, the generated BVs are tuned to the amplitude of the global analysis RMSE to use as the initial ensemble perturbations.

  • The ETKF perturbations are generated based on an ensemble DA algorithm (Wang and Bishop, 2003). We denote X' f and X' a as N forecast and analysis perturbations superposed on the control analysis. Forecast perturbations are listed as columns in the matrix X' f and analysis perturbations are listed as columns in the matrix X' a. The scheme is created to calculate a transformation matrix, T, which transforms the forecast perturbations according to \begin{equation} \label{eq1} {X}'_{\rm a}={X}'_{\rm f}{T} . (1)\end{equation} In this expression, T is an N× N matrix, and can be calculated as \begin{equation} \label{eq2} {T}={C}({\pmb{\it\Gamma}}+{I}_N)^{-1/2} ,(2) \end{equation} where columns of the matrix C contain the eigenvectors of (X' f) TH TR-1HX' f/(N-1), and the diagonal matrix \(\pmb\it\Gamma\) contains the corresponding eigenvalues. H denotes the mapping from the model space to the observation space. R is the observation error covariance matrix. The ETKF ensemble perturbations can be proved orthogonal to each other in the space of the observations (Wang and Bishop, 2003).

    The forecast perturbations in each cycle are multiplied by an inflation factor of 7% to maintain sufficient spread. The updated ensemble perturbations will be added to the control analysis to generate the initial conditions for the next filter step. The length of a cycle is 12 hours and the process is continued over 5 days to the initial analysis for forecasting. Each model grid has the same observation error variance of 0.02, i.e. the observation error covariance matrix R is a diagonal matrix with 0.02 entries.

  • Let us consider an n-dimensional nonlinear dynamical system: \begin{equation} \label{eq3} \frac{d}{dt}{x}(t)={F}[{x}(t)] , (3)\end{equation} where x(t)=(x1(t),x2(t),……,xn(t)) T represents the state vector at time t governed by the dynamics F. Let δ(t0) be a small error superposed on x(t0) at an initial time t=t0. The differential equation of the small error δ(t) is written as follows: \begin{equation} \label{eq4} \frac{d}{dt}{\delta}(t)={J}({x}(t)){\delta}(t)+{G}({x}(t),{\delta}(t)) , (4)\end{equation} where J(x(t))δ(t) and G(x(t),δ(t)) are the tangent linear terms and the high order nonlinear terms of the errors δ(t), respectively. By integrating Eq. (5) from t=t0 to t0, the nonlinear error evolution can be obtained: \begin{equation} \label{eq5} {\delta}(t_0+\tau)={\eta}({x}(t_0),{\delta}(t_0),\tau){\delta}(t_0) ,(5) \end{equation} where η(x(t0),δ(t0),τ) is the nonlinear propagator (Ding and Li, 2007). In a chaotic system, each initial error vector tends to fall along the fastest growing direction. Therefore, for any initial error vector δ'(t0), after a short time τ, δ'(t0+τ) will capture the fastest growing direction. If the short time τ is not taken into consideration, taking δ'(t0+τ) as the initial error δ1(t0), δ1(t0) is defined as the LNLLV here, which is similar to the BV. The first nonlinear local Lyapunov exponent (NLLE) along the fastest growing direction LNLLV can be approximately defined as \begin{equation} \label{eq6} \lambda_1({x}(t_0),{\delta}_1(t_0),\tau)=\frac{1}{\tau}\ln\frac{\|{\delta}_1(t_0+\tau)\|}{\|{\delta}_1(t_0)\|} , (6)\end{equation} where Λ1(x(t0),δ1(t0),τ) is the function of the initial state x(t0) in phase space, the initial error δ1(t0), and evolution time τ. In contrast, the linear Lyapunov exponent relies on the assumptions of infinite evolution time and infinitesimal initial error (Lorenz, 1965; Yoden and Nomura, 1993; Ziehmann et al., 2000), and thus has limitations when applied to nonlinear error growth. The NLLE has been widely used to investigate the predictability of weather and climate (Li et al., 2006; Ding and Li, 2007; Ding et al., 2008, 2010, 2011; Li and Ding, 2011, 2013).

    While the LNLLV tends to the fastest growing direction, the rest of the NLLVs can be derived via a periodic reorthogonalization by the GSR process (Wolf et al., 1985; Li and Wang, 2008, Feng et al., 2014). The orthogonalized perturbations are then globally rescaled to the initial size and again superposed on the new analysis to integrate (see Fig. 1). The process is performed for five days with a 12-hour cycle to acquire the NLLVs denoted by δ1(t0),δ2(t0),…,δN(t0). This 5-day period is basically close to that used in (Vannitsem and Nicolis, 1997) to generate LVs in a quasigeostrophic model. The NLLVs are orthogonal to each other, representing the vectors along the directions from the fastest growing direction to the fastest shrinking direction in local phase space. The resulting NLLV perturbations are tuned to the amplitude of the global analysis RMSE and superposed on the analysis to generate ensemble members.

    Once the NLLVs are obtained, the growth rate of perturbations along each direction can be estimated by the NLLEs, which are defined by \begin{equation} \label{eq7} \lambda_i({x}(t_0),{\delta}_i(t_0),\tau)=\frac{1}{\tau}\ln\frac{\|{\delta}_i(t_0+\tau)\|}{\|{\delta}_i(t_0)\|} , (7)\end{equation} where δi(t0) is the i-th NLLV at time t0. If we want to study the dynamic characteristics of the whole system, the ensemble average of the i-th NLLE, \(\overline \lambda_i(\delta_i(t_0),\tau)=\langle\lambda_i(x(t_0),\delta_i(t_0),\tau)\rangle\) (the symbol \(\langle\ \ \rangle\) denotes the ensemble average over a great number of different initial states) should be introduced.

    The background forecast errors inserted in the analysis errors are the combined effects of the multidimensional growing subspace (Palatella and Trevisan, 2015) spanning the NLLVs. Therefore, to better sample the analysis errors, the initial ensemble perturbations are constructed by a simple random linear combination of the first N NLLVs (N=10). The NLLVs are denoted by \(\pmb\it\Delta=\delta_1,\delta_2,\ldots,\delta_N\), then the orthogonal ensemble perturbations \(\pmb\it\Delta\) can be calculated by \begin{equation} \label{eq8} {\pmb{\it\Delta}}_{\rm e}={\pmb{\it\Delta}}{\beta} , (8)\end{equation} where β is an N× N matrix. The column vectors of β are constructed by random numbers of a standard normal distribution and are independent from each other. The resulting ensemble perturbations are rescaled to the amplitude of the global analysis error and superposed on the analysis to generate ensemble members.

    Figure 1.  Creation of NLLVs. A group of RPs is initially introduced on the analysis state, and integrated to the end of a breeding cycle. The orthogonalization processes are conducted regularly. The red solid curve represents the leading perturbed forecast during each cycle, and the black solid curves represent other perturbed forecasts from the second to the $n$th fastest. The direction of the fastest growing mode (LNLLV, red dashed line) is kept, and the other directions (black dashed lines) are orthogonalized successively to acquire NLLV$_{2}$ to NLLV$_{n}$ (blue dashed line). These perturbations are then rescaled to the same size as the initial perturbations, and added to the new analysis. After several days of cycling, the perturbations are dominated by growing components [modified from Feng et al. (2014)]

3. Experimental setup
  • The two-dimensional quasi-geostrophic model is governed by the following system of dimensionless differential equations: \begin{equation} \label{eq9} \left\{ \begin{array}{l} \dfrac{\partial q}{\partial t}+J(\psi,q)=0\\[3mm] q=\nabla^2\psi-F\psi+f_0+f_0h_s,\quad {\rm in}\quad \Omega\times[0,T]\\[1mm] \psi|_{t=0}=\psi_0 \end{array} \right. ,(9) \end{equation} where the variables q and ψ are the potential vorticity and the stream function, respectively; 2=∂2/∂ k2+∂2/∂ l2 and J(ψ,q)=ψkqlkql are the Laplacian operator and the horizontal Jacobian operator, respectively; k and l represent the zonal and meridional coordinates; and t denotes time. The planetary Froude number F=0.102, and the Coriolis parameter f0=10.0. A double periodical boundary condition is used for \(\Omega=[0,K]\times[0,L]\). The barotropic model has been used in various studies of predictability and ensemble forecasts (Barkmeijer, 1992; Cheung and Chan, 1999; Mu and Zhang, 2006; Mu and Jiang, 2008; Durran and Gingrich, 2014).

    All the experiments below are performed on a 32× 16 latitude-longitude horizontal grid. The grid spacing d=0.2 corresponds to a dimensional length of 200 km, and the time step dt=0.006 corresponds to 10 min. The basic flow is obtained by integrating (10) with an initial state \(\psi_0=0.5\sin(\pi k/6.4)+\sin(\pi l/3.2)+0.5\), as used by (Mu and Jiang, 2008). For simplicity, the topography is only a function of the l-direction \(h_s=h_0\times[\sin(2\pi l/3.2)+1]\).

    The quasi-geostrophic model is integrated for 800 days as a "truth run", after a first "spin-up" run of 100 days. The quality of the forecasts are assessed by comparing with the truth run. For a fair comparison, the EnKF assimilation method, which is irrelevant to the ensemble generation schemes, is used to generate the control analysis. The observations are distributed at each grid point, and are produced by adding a random noise to the true state. The assimilation process is implemented every 12 hours on the stream function. The details of the EnKF assimilation steps are described in Appendix A.

    Once the analysis states have been obtained, 10-day ensemble forecasts initialized from the four ensemble generation methods, respectively, are implemented. All the initial ensemble perturbations are produced using the Euclidean norm. The RP perturbations are simply generated by random numbers with a standard normal distribution, and then rescaled to the amplitude of initial analysis errors. The ensemble perturbations of the RP scheme are flow-independent and thus can provide a proper reference for the other theory-based schemes (Magnusson et al., 2009). To clarify the differences of the four initialization schemes in a simple way, in this study, the N=10 ensemble perturbations for the four schemes are directly superposed on the analysis state, without being centered, to generate ensemble members. The statistical quality of the forecasts is assessed by averaging a set of 100 different experiments, each corresponding to a different initial state randomly chosen from the 800-day truth run. All experiments are "perfect model" experiments in which the assimilation and forecasts are performed with the same model used for producing the truth run.

4. Results
  • Before using the barotropic model to compute the NLLEs and NLLVs, it should be ensured that this model is a nonlinear dynamical system and has a chaotic attractor (Li and Chou, 1997). Figure 2 shows the temporal evolution of the stream function at one grid point for 800 days. It can be seen that the trajectory presents a chaotic behavior, which indicates the model is suitable for the research on nonlinear error growth.

    To see if different growing directions can be effectively captured by the NLLVs, the transient error growths of NLLVs are measured by the NLLEs. Figure 3 shows the first 50 NLLEs averaged over 800 initial conditions with a one-day interval. The NLLEs are calculated for the error growth during the initial τ=12 h and the initial amplitude of perturbations is 0.01. It can be seen that the first 50 NLLEs are consistently positive and continuously arranged in a decreasing order. The results demonstrate that the barotropic model has a high-dimensional growing subspace. The development of the analysis error is a combined effect of various independent growing directions, and the NLLVs may be capable of capturing the unstable subspace effectively.

    Figure 2.  The evolution of the stream function at one grid point of the barotropic model for 800 days.

    Figure 3.  The first 50 NLLEs for the barotropic model with the evolution time $\tau$ equal to 12 hours and a finite initial amplitude of 0.01

  • Since analysis errors are poorly known, the initial ensemble perturbations should be as independent as possible to sample the possible direction of analysis errors. To see the independence of the global perturbations, the average variances explained by the eigenvectors of the ensemble covariance matrix are calculated for the four schemes, respectively (Wang and Bishop, 2003; Bowler, 2006). The assessments for the ensemble perturbations superposed on the control forecasts at the different lead times are shown in Fig. 4. A flatter distribution of the explained variance indicates better perturbation independence. At day 0, the explained variances for the NLLV and ETKF perturbations are strictly evenly distributed, which can be induced by their theoretical concepts. The RP perturbations are very close to being orthogonal to each other. However, the first eigenvector of the BVs explains almost 30% of the error variance when 10 perturbations are used, indicating that there is certain similarity among the perturbations. That is to say, the effective number of ensemble perturbations for the BV method is lower than its actual number. With the increase in lead time, the similarities among the ensemble perturbations for the four schemes gradually increase due to the dynamical evolution on the same basic flows. The performances of the four initialization schemes tend to be similar after about day 6.

    Figure 4.  Mean explained variances of the eigenvectors of forecast covariance matrices measured from day 0 to 10. The BV, NLLV, RP and ETKF methods are denoted by triangles, open circles, squares and crosses, respectively.

    Figure 5.  Evolution of mean error variance of the control run explained by the ensemble perturbations as a function of lead time.

    Initial ensemble perturbations are created to sample the possible forecast errors. Therefore, the variance of forecast errors explained by the subspace of ensemble perturbations is an important assessment of the quality of perturbations. For a fair comparison of the four schemes, the ensemble forecast errors are defined as the difference between each ensemble forecast member and the control forecast, while the control forecast error is the difference between the true state and the control forecast. Let us assume the normalized ensemble forecast errors are denoted by E=e1,e2,…,eN, and the control forecast error is e0. The projection of the control forecast error on the subspace of ensemble forecast errors would be derived by p=EE Te0. Then, the error variance of the control forecast explained by the ensemble perturbation subspace can be calculated through the correlation between p and e0.

    Figure 5 illustrates the variance of analysis errors explained by the ensemble perturbations superposed on the control forecasts as a function of lead time for the four methods. At the initial time, the RP approach performs the worst in terms of capturing the analysis error direction, as the RPs are independent of the flow at initial time. The NLLV and ETKF perturbations have similar explained variance, which is larger than that of the BV perturbations. This may be due to the more independent perturbations of the former two approaches than the latter one (see Fig. 4). The initial explained variances of the four schemes may also cause their relative performance afterwards. The ETKF and NLLV techniques basically remain similar and the best from day 0 to 10. The BV approach always has a lower score compared to the ETKF and NLLV approaches during the 10 days. The explained variance for the RP, despite exhibiting the fastest increase among the four schemes, has the worst performance throughout.

  • The RMSE and the pattern anomaly correlation (PAC) of the ensemble mean are often used to measure the overall forecast performance of ensemble forecasts. Figure 6 illustrates the RMSE and PAC as a function of time for the four schemes. The ensemble spread, calculated as the standard deviation of ensemble members, is also given as a reference. The ensemble spread of a statistically reliable ensemble system is usually close to the RMSE of the ensemble mean for all lead times (Buizza et al., 2005). From Fig. 6a, it can be seen that for the first 4 days the NLLV, RP and ETKF ensembles have similar RMSE scores, which are slightly smaller than that of the BV scheme. This may because the BVs are constrained to a smaller perturbation subspace compared to the other three types of initial perturbations (see Fig. 4). Note that the ensemble spread of the RP approach is smaller than the other three for 0-4 days. This is due to the RPs pointing to non-growing directions and will shrink initially (Toth and Kalnay, 1997). From day 4 to 10, the NLLV has slightly smaller RMSE and similar ensemble spread compared with ETKF, while the BV scheme has somewhat larger RMSE and smaller spread than both the NLLV and ETKF schemes. This is because the NLLV perturbations, which consider the nonlinear interactions of growing directions, have a better sampling of the development of the analysis errors than the ETKF perturbations, and both of them span more orthogonal and uncorrelated subspaces than do the BVs. Although the RP approach has the largest ensemble spread, its forecast skill decreases rapidly and remains the lowest among the four schemes. This can be explained by the RPs having very limited ability to capture the evolution of initial analysis errors. The ensemble spread for both the NLLV and ETKF methods are closer to their respective ensemble mean errors than the BV method, which implies that the former two have better statistical reliability than the latter one. The PAC score in Fig. 6b presents similar relative performance of the ensemble mean skill as the RMSE.

    Figure 7 shows the ratio of the highest ensemble mean skill among the four initialization schemes for each of them as a function of lead time. It can be seen that the ETKF and NLLV approaches basically have similar ratios to reach the highest forecast skill, which is evidently higher than that of the BV and RP approaches. The RP approach performs better than the BV approach from day 1 to 4. Thereafter, the ratio of the RP scheme remains the lowest among the four schemes to the end of the lead time. The statistical results for each case basically correspond to the sample mean results.

    Figure 6.  Evolution of the (a) mean RMSE (solid lines) and ensemble spread (dashed lines) and (b) PAC (solid lines) for the BV, NLLV, RP and ETKF schemes as a function of lead time.

    Figure 7.  Evolution of the ratio of the (a) smallest RMSE and (b) highest PAC among the BV, NLLV, RP and ETKF initialization schemes for each of them as a function of lead time.

  • A commonly used measure in probabilistic forecasts is the Brier score (BS), which can be used to evaluate the reliability (BS rel) and resolution (BS res) of an ensemble forecast system (Murphy, 1973; Descamps and Talagrand, 2007; Stephenson et al., 2008). The score assesses the performance of the probability forecast of the occurrence of an event Φ. Further details are given in, for example, (Descamps and Talagrand, 2007) and (Stephenson et al., 2008). In this paper, Φ is defined as the event ψ>ψ c, where ψ c is the climatological mean at each grid point. A smaller value of BS and BS rel indicates a better forecast skill; while the BS res acts in an opposite way. Another event interval (ψ>ψ c, where σ is the standard deviation of ψ) is also introduced for diagnostics, but does not appear to affect the relative performance of the ensemble forecasts. Only the results of the event Φ are shown in this paper.

    Figure 8a shows the basic BS evaluated for the four methods as a function of lead time. The four methods have similar basic BS scores during days 0-4. The NLLV scheme shows slightly better performance than the ETKF scheme. The BV scheme has a lower basic BS than the RP scheme for days 4-10, and both of them perform worse than the ETKF and NLLV schemes. Their differences in BS performance mainly come from the BS res and the BS rel components, as shown in Fig. 8b. After 4 days, the higher BS res of the NLLV and the ETKF schemes than the BV scheme may be attributable to the worse sampling of the analysis errors by the BV approach. Meanwhile, the former two approaches have better independence and larger ensemble spread than the BV approach, which may result in a lower BS rel relative to the BV approach. The RP scheme performs worst among the four schemes generally. This is because the RP ensemble is randomly generated, and thus is less likely to describe the probability distribution of the truth relative to the other three schemes.

    Figure 8.  Evolution of (a) average basic BS as a function of lead time for the BV, NLLV, RP and ETKF schemes; and (b) resolution (dotted lines) and reliability (solid lines) contributions to the BS.

5. Discussion and conclusions
  • The NLLV ensemble generation scheme retains the advantage of the BV scheme, which uses the breeding cycles to capture the development of analysis errors in the assimilation cycles and is very simple and cheap to implement. Different from the BVs, the NLLV perturbations are periodically orthogonalized by the GSR process to identify various distinguished growing perturbations, and thus have more diversity and independence. In this paper, the application of the NLLV method introduced by (Feng et al., 2014) for the generation of ensemble perturbations is further extended to a moderately complex barotropic model. Except for the BV and RP approaches, the performance of the NLLV scheme for ensemble forecasting is further compared with that of the ETKF scheme, which is one of the most advanced ensemble initialization schemes. It is found that the NLLV scheme performs slightly better overall than the ETKF scheme for various scores when the observations to be assimilated are evenly distributed at each grid. This may because the development of analysis errors can be effectively sampled by the unstable perturbations acquired by the NLLVs. The BV approach performs worse than the ETKF and NLLV approaches due to the dependence among perturbations, but still dramatically exceeds the RP approach.

    This study simply uses a barotropic model without model error, and the model size is less than that of operational forecasts. However, the quasi-geostrophic model is sufficiently large to demonstrate the relative performance of the ensemble schemes accurately. Despite the similar forecast skills of the NLLV and ETKF schemes, the generation of the NLLVs is significantly time-saving and easy to implement, compared to the ETKF scheme. The computational expense of the former is only about one third of the latter in the experimental environment of this paper. Meanwhile, the quality of ETKF perturbations largely depends on the estimation of the background and observation covariance matrices, while an accurate estimate of the forecast error covariance is very challenging due to such problems as ensemble "collapse" (Anderson and Anderson, 1999) and spurious correlations at large distances (Houtekamer and Mitchell, 2001). In a real world situation where the observations are irregularly distributed, a prior knowledge of the regional rescaling factors could be used to modulate the NLLV perturbations to better sample the analysis error variances. Several current approaches can provide the estimated masks, like the error growth statistical method (Pe\ na and Toth, 2014). Introducing the estimation of analysis error variances in tuning the NLLV perturbations may further improve the ensemble forecast skill of the NLLV scheme.

  • The Ensemble Kalman Filter

    The true state is denoted by ${x}_{\rm true}$, which is derived from a long-term model integration. The simulated observations ${y}$ are generated based on the true state using \begin{equation} {y}={Hx}_{\rm true}+{\varepsilon}~,(A1)\end{equation} where ${H}$ provides a mapping from the model space to the observation space, and ${\varepsilon}$ represents independent realizations of the noise with a Gaussian distribution $N(0,0.02)$. The ensemble of forecast states is adopted as background states. The ensemble matrix is then defined as \begin{equation} {X}_{\rm f}=({x}_{\rm f,1},{x}_{\rm f,2},\ldots,{x}_{\rm f,{N}})~. (A2)\end{equation} where $\overline{{x}}_{\rm f}$ represents the mean of the ensemble. An ensemble perturbation matrix can then be written as \begin{equation} {X}'_{\rm f}=({x}_{\rm f,1}-\overline{{x}}_{\rm f},{x}_{\rm f,2}-\overline{{x}}_{\rm f},\ldots,{x}_{\rm f,{N}}-\overline{{x}}_{\rm f})~. (A3)\end{equation} The covariance matrix of the ensemble ${X}_{\rm f}$ is: \begin{equation} {P}_{\rm f}=\frac{1}{N-1}{X}'_{\rm f}{X}'^{\rm T}_{\rm f}~. (A4)\end{equation} The background forecast ensemble is to be updated by the observations. The set ${y}_i(i=1,2,\ldots,N)$ is a set of perturbed observations that are associated with each previous forecast of the ensemble. They are defined as: \begin{equation} {y}_i={y}+{\varepsilon}_i~. (A5)\end{equation} The RPs ${\varepsilon}_{i}$ follow the same distribution as ${\varepsilon}$. The observations are assimilated to produce a new analysis of the state: \begin{equation} {x}_{{\rm a},i}={x}_{{\rm f},i}+{K}({y}_i-{Hx}_{{\rm f},i}),\quad i=1,2,\ldots,N~. (A6)\end{equation} The Kalman gain ${K}$ is calculated by: \begin{equation} {K}={P}_{\rm f}{H}^{\rm T}({HP}_{\rm f}{H}^{\rm T}+{R})^{-1}~. (A7)\end{equation} This is actually a weight measuring the ratio of the forecast and observational error covariance. ${R}$ is the observational error covariance matrix.

    To overcome the problem of undersampling, a 7{\%} variance inflation factor is applied to ${X}'_{\rm f}$ in this barotropic model. Moreover, the number of the ensemble is 100, which is much smaller than the model dimension. Therefore, the localization technique is applied to the matrix ${P}_{\rm f}$ to prevent spurious correlations at large distances. This is realized by the fifth-order function of \cite{Gaspari1999} with a distance of zero correlation equal to 800 km (four grids). The assimilation cycles are repeated for 30 days in each case to generate the analysis ensemble. The mean $\overline{{x}}_{\rm a}$ of the analyzed ensemble ${x}_{{\rm a},i}$($i=1,2,\ldots,N$) is regarded as the initial analysis state when performing the forecasts.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return