Advanced Search
Article Contents

Comparison of Constant and Time-variant Optimal Forcing Approaches in El Niño Simulations by Using the Zebiak-Cane Model


doi: 10.1007/s00376-015-5174-8

  • Model errors offset by constant and time-variant optimal forcing vector approaches (termed COF and OFV, respectively) are analyzed within the framework of El Niño simulations. Applying the COF and OFV approaches to the well-known Zebiak-Cane model, we re-simulate the 1997 and 2004 El Niño events, both of which were poorly degraded by a certain amount of model error when the initial anomalies were generated by coupling the observed wind forcing to an ocean component. It is found that the Zebiak-Cane model with the COF approach roughly reproduced the 1997 El Niño, but the 2004 El Niño simulated by this approach defied an ENSO classification, i.e., it was hardly distinguishable as CP-El Niño or EP-El Niño. In both El Niño simulations, substituting the COF with the OFV improved the fit between the simulations and observations because the OFV better manages the time-variant errors in the model. Furthermore, the OFV approach effectively corrected the modeled El Niño events even when the observational data (and hence the computational time) were reduced. Such a cost-effective offset of model errors suggests a role for the OFV approach in complicated CGCMs.
  • 加载中
  • An S. I., F. F. Jin, 2001: Collective role of thermocline and zonal advective feedbacks in the ENSO mode. J.Climate, 14( 16), 3421- 3432.7142b668-36b0-40b9-a855-1be5c86e3c2f7cced414a9afbb584b1be474e9a03f7bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2001JCli...14.3421Arefpaperuri:(2987402d9d46b314f43cf81ab7d0c8e4)/s?wd=paperuri%3A%282987402d9d46b314f43cf81ab7d0c8e4%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2001JCli...14.3421A&ie=utf-8
    Ashok K., S. K. Behera, S. A. Rao, H. Y. Weng, and T. Yamagata, 2007: El Niño Modoki and its possible teleconnection. J. Geophys. Res., 112,C11007, doi: 10.1029/2006JC003798.10.1029/2006JC0037984c9ce4f6-43f6-4657-a703-db6ce800c67d9ed3e543c8e9bfe16c01f5a1395c1babhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2006JC003798%2Fabstract%3Bjsessionid%3DF2C5797F7539FC367A07E919851C7401.f03t02refpaperuri:(42a43d3a8b5a3852d00d87760b9d78f8)http://onlinelibrary.wiley.com/doi/10.1029/2006JC003798/abstract;jsessionid=F2C5797F7539FC367A07E919851C7401.f03t02[1] Using observed data sets mainly for the period 1979–2005, we find that anomalous warming events different from conventional El Ni09o events occur in the central equatorial Pacific. This unique warming in the central equatorial Pacific associated with a horseshoe pattern is flanked by a colder sea surface temperature anomaly (SSTA) on both sides along the equator. empirical orthogonal function (EOF) analysis of monthly tropical Pacific SSTA shows that these events are represented by the second mode that explains 12% of the variance. Since a majority of such events are not part of El Ni09o evolution, the phenomenon is named as El Ni09o Modoki (pseudo-El Ni09o) (“Modoki” is a classical Japanese word, which means “a similar but different thing”). The El Ni09o Modoki involves ocean-atmosphere coupled processes which include a unique tripolar sea level pressure pattern during the evolution, analogous to the Southern Oscillation in the case of El Ni09o. Hence the total entity is named as El Ni09o–Southern Oscillation (ENSO) Modoki. The ENSO Modoki events significantly influence the temperature and precipitation over many parts of the globe. Depending on the season, the impacts over regions such as the Far East including Japan, New Zealand, western coast of United States, etc., are opposite to those of the conventional ENSO. The difference maps between the two periods of 1979–2004 and 1958–1978 for various oceanic/atmospheric variables suggest that the recen
    Barkmeijer J., T. Iversen, and T. N. Palmer, 2003: Forcing singular vectors and other sensitive model structures. Quart. J. Roy. Meteor. Soc., 129( 592), 2401- 2423.10.1256/qj.02.126da4922be9ac454e4abf8c93b1c648205http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1256%2Fqj.02.126%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1256/qj.02.126/citedbyABSTRACT Model tendency perturbations can, like analysis perturbations, be an effective way to influence forecasts. In this paper, optimal model tendency perturbations, or forcing singular vectors, are computed with diabatic linear and adjoint T42L40 versions of the European Centre for Medium-Range Weather Forecasts' forecast model. During the forecast time, the spatial pattern of the tendency perturbation does not vary and the response at optimization time (48 hours) is measured in terms of total energy. Their properties are compared with those of initial singular vectors, and differences, such as larger horizontal scale and location, are discussed. Sensitivity calculations are also performed, whereby a cost function measuring the 2-day forecast error is minimized by only allowing tendency perturbations. For a given number of minimization steps, this approach yields larger cost-function reductions than the sensitivity calculation using only analysis perturbations. Nonlinear forecasts using only one type of perturbation confirm an improved performance in the case of tendency perturbations. For a summer experiment a substantial reduction of the systematic error is shown in the case of forcing sensitivity. Copyright 漏 2003 Royal Meteorological Society.
    Barreiro M., P. Chang, 2004: A linear tendency correction technique for improving seasonal prediction of SST. Geophys. Res. Lett., 31(23),L23209, doi: 10.1029/2004gl021148.10.1029/2004GL02114806ab0084843ca7b090edb009f9254252http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2004GL021148%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2004GL021148/fullAbstract Top of page Abstract 1.Introduction 2.Linear Tendency Correction Technique 3.Model and Present Application 4.Results 5.Summary Acknowledgments References [1] A methodology is presented to linearly correct the tendency of sea surface temperature (SST) anomalies in a coupled model. Using an atmospheric general circulation model (AGCM) coupled to a slab ocean as an example, we demonstrate the effectiveness of the linear correction methodology in improving the model's skill predicting SST in the tropical Atlantic Ocean during boreal spring. For this particular coupled model, the correction mainly takes into consideration the linear ocean dynamics absent in the slab ocean, thereby improving the skill in the tropical south and equatorial Atlantic. The corrected coupled model is further shown to produce a skillful rainfall forecast in the intertropical convergence zone (ITCZ) region during the boreal spring.
    Behringer D. W., M. Ji, and A. Leetmaa, 1998: An improved coupled model for ENSO prediction and implications for ocean initialization. Part I: The ocean data assimilation system. Mon. Wea. Rev., 126( 5), 1013- 1021.
    Blumenthal M. B., 1991: Predictability of a coupled ocean-atmosphere model. J.Climate, 4( 8), 766- 784.10.1175/1520-0442(1991)0042.0.CO;23f9d5714-9e26-47df-8e78-ec2b82eee9e4cc87c95915f17771852a1c52568374eahttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2F092479639190028Srefpaperuri:(70af9186f72d36879e1ffc80ec54a654)http://www.sciencedirect.com/science/article/pii/092479639190028SAbstract The predictability and variability of a coupled ocean-atmosphere model has been investigated by examining the growth of small initial perturbations during the evolution of the coupled system. The ocean model is first integrated in a forced mode for a duration of over 24 years beginning with January 1964 in which wind stress forcing for each month is prescribed from the observations. This provides surrogate analysis or control run with which predictions from the coupled model can be initiated and compared. Starting from January 1970 with each of the next 181 initial states from the control run, a prediction experiment was carried out for a duration of 36 months each using the fully coupled model. With this large ensemble of prediction experiments, a detailed analysis of growth of initial error and forecast errors was carried out. The SST forecasts are compared with observations as well as the control run. The root-mean-square difference between control and forecasts becomes larger than the standard deviation of the control as well as persistence error in about three months. As a result of differences between the simulated SST in the control run and observations, the forecasts are forced to have initial errors that are comparable to the standard deviation of the observations. Some significant systematic errors in the model are also noted. There is an indication that the forecasts may be improved to some extent by averaging a few of the most recent available forecasts and removing the known systematic error. Also carried out is a large ensemble of identical twin experiments, each for a duration of 15 years. In one of each pair of experiments a small random perturbation is introduced at the initial time in the surface winds. These experiments have shown that the growth of small initial errors in the coupled model is governed by processes with two well-separated time scales. The fast time scale process introduces errors that have a doubling time of about 5 months, while the slow time scale process introduces errors that have a typical doubling time of about 15 months. The existence of a slow time scale gives us optimism about long-range forecasts of ENSO-type events. However, the fast growth rate tends to saturate at a level that is comparable to the climatological standard deviation. Thus, a key to long-range forecasting of ENSO-type events may lie in the ability to identify those initial states that are not too sensitive to the processes associated with fast growth rate. The diagnostic analysis shows that the first three empirical orthogonal functions (EOF) of the observed wind stress together explain only about 36% of the total variance. Although the observed wind stress has considerable amplitude in the higher EOFs, it is shown that only the first three components are important for forcing the observed interannual variations using this model. The atmospheric component of the coupled model is not able to simulate these large-scale components of the observed wind stress accurately. This is partly because the atmospheric model is mainly driven by the underlying sea surface temperature anomalies (SSTA) and partly due to the structural differences between the SSTA simulated by the model and the observed SSTA. Thus, a combination of the atmospheric component's tight coupling to the ocean and the ocean model's inability to simulate the SST anomalies correctly seems to be responsible for the rather rapid growth of prediction errors.
    Bourassa, M. A., Rosario Romero, Shawn R. Smith, James J. O'Brien, 2005: A new FSU winds climatology. J. Climate,18, 3686-3698, doi: http://dx.doi.org/10.1175/JCLI3487.1.10.1175/JCLI3487.1096be2241d0bd4e7a17b8654bb5e2ee7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2005JCli...18.3686Bhttp://adsabs.harvard.edu/abs/2005JCli...18.3686BA new objective time series of in situased monthly surface winds has been developed as a replacement for the subjective tropical Pacific Florida State University (FSU) winds. The new time series begins in January 1978, and it is ongoing. The objective method distinguishes between observations from volunteer observing ships (VOSs) and buoys, allowing different weights for these different types of observations. An objective method is used to determine these weights and accounts for the differences in error characteristics and in spatial/temporal sampling. A comparison is made between the objective and subjective products, as well as scatterometer winds averaged monthly on the same grid. The scatterometer fields are a good proxy for truth. These three sets of fields have similar magnitudes, directions, and derivative fields. Both in situ wind products underestimate convergence about the intertropical convergence zone; however, the objective FSU product is a much better match to the scatterometer observations. Furthermore, the objective winds have smaller month-to-month variation than the subjective winds. Composites of ENSO phases are also examined and show minor differences between the subjective and objective wind products. The strengths and weaknesses of the objective and subjective winds are discussed.
    Chen D. K., S. E. Zebiak, A. J. Busalacchi, and M. A. Cane, 1995: An improved procedure for El Niño forecasting: Implications for predictability. Science, 269, 1699- 1702.f0c8e6c1-a7a5-42ca-a30d-b9e4b80bf729f10992904d7c885966104328d9c51383http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F245343887_An_improved_procedure_for_El_Nino_forecastingrefpaperuri:(20ab5bab11cc11dc54936082b810ee79)http://www.researchgate.net/publication/245343887_An_improved_procedure_for_El_Nino_forecasting
    Chen D. K., C. M. Cane, S. E. Zebiak, R. Ca\nizares, and A. Kaplan, 2000: Bias correction of an ocean-atmosphere coupled model. Geophys. Res. Lett.,27(16), 2585-2588, doi: 10.1029/ 1999gl011078.10.1029/1999GL0110784dcd5975-3950-4ba2-a85c-064ac6223a1f0e32511d08e58f56ce53c88292dcee18http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F1999GL011078%2Ffullrefpaperuri:(ced32067b187c6f7134ae41af4e87caa)http://onlinelibrary.wiley.com/doi/10.1029/1999GL011078/fullAbstract Top of page Abstract References A serious problem in the initialization of a climate forecast model is the model-data incompatibility caused by systematic model biases. Here we use the Lamont model to demonstrate that these biases can be effectively reduced with a simple statistical correction, and the bias-corrected model can have a more realistic internal variability as well as an improved forecast performance. The results reported here should be of practical use to other ocean-atmosphere coupled models for climate prediction.
    Chen D. K., M. A. Cane, A. Kaplan, S. E. Zebiak, and D. J. Huang, 2004: Predictability of El Niño over the past 148 years. Nature, 428( 6984), 733- 736.10.1038/nature02439150851270e566377-8b8f-4424-8c42-09e1127596eb8b9772bafd37304a03809c39c21a610bhttp%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM15085127refpaperuri:(996444e2daa89c031d9c25496a0d31f9)http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM15085127Forecasts of El Ni09o climate events are routinely provided and distributed, but the limits of El Ni09o predictability are still the subject of debate. Some recent studies suggest that the predictability is largely limited by the effects of high-frequency atmospheric `noise', whereas others emphasize limitations arising from the growth of initial errors in model simulations. Here we present retrospective forecasts of the interannual climate fluctuations in the tropical Pacific Ocean for the period 1857 to 2003, using a coupled ocean-atmosphere model. The model successfully predicts all prominent El Ni09o events within this period at lead times of up to two years. Our analysis suggests that the evolution of El Ni09o is controlled to a larger degree by self-sustaining internal dynamics than by stochastic forcing. Model-based prediction of El Ni09o therefore depends more on the initial conditions than on unpredictable atmospheric noise. We conclude that throughout the past century, El Ni09o has been more predictable than previously envisaged.
    D'andrea F., R. Vautard, 2000: Reducing systematic errors by empirically correcting model errors. Tellus A, 52( 2), 21- 41.10.1034/j.1600-0870.2000.520103.x37abd90094c0916ae5df873f9b1776b1http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1034%2Fj.1600-0870.2000.520103.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1034/j.1600-0870.2000.520103.x/citedbyA methodology for the correction of systematic errors in a simplified atmospheric general‐circulation model is proposed. First, a method for estimating initial tendency model errors is developed, based on a 4‐dimensional variational assimilation of a long‐analysed dataset of observations in a simple quasi‐geostrophic baroclinic model. Then, a time variable potential vorticity source term is added as a forcing to the same model, in order to parameterize subgrid‐scale processes and unrepresented physical phenomena. This forcing term consists in a (large‐scale) flow dependent parametrization of the initial tendency model error computed by the variational assimilation. The flow dependency is given by an analogues technique which relies on the analysis dataset. Such empirical driving causes a substantial improvement of the model climatology, reducing its systematic error and improving its high frequency variability. Low‐frequency variability is also more realistic and the model shows a better reproduction of Euro‐Atlantic weather regimes. A link between the large‐scale flow and the model error is found only in the Euro‐Atlantic sector, other mechanisms being probably the origin of model error in other areas of the globe.
    Duan W. S., B. Tian, and H. Xu, 2014: Simulations of two types of El Niño events by an optimal forcing vector approach. Climate Dyn.,43, 1677-1692, doi: 10.1007/s00382-013-1993-4.10.1007/s00382-013-1993-44b647156-3ec6-45dd-8bf0-f833b4c619d84ebc6b100c586b439019d3b0b03ea509http%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00382-013-1993-4refpaperuri:(796236032515a4c4ebd5299e25d13233)http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s00382-013-1993-4In this paper, an optimal forcing vector (OFV) approach is proposed. The OFV offsets tendency errors and optimizes the agreement of the model simulation with observation. We apply the OFV approach to the well-known Zebiak–Cane model and simulate several observed eastern Pacific (EP) El Ni09o and central Pacific (CP) El Ni09o events during 1980–2004. It is found that the Zebiak–Cane model with a proper initial condition often reproduces the EP-El Ni09o events; however, the Zebiak–Cane model fails to reproduce the CP-El Ni09o events. The model may be much more influenced by model errors when simulating the CP-El Nino events. As expected, when we use the OFV to correct the Zebiak–Cane model, the model reproduces the three CP-El Ni09o events well. Furthermore, the simulations of the corresponding winds and thermocline depths are also acceptable. In particular, the thermocline depth simulations for the three CP-El Ni09o events lead us to believe that the discharge process of the equatorial heat content associated with the CP-El Ni09o is not efficient and emphasizes the role of the zonal advection in the development of the CP-El Nino events. The OFVs associated with the three CP-El Ni09o events often exhibit a sea surface temperature anomaly (SSTA) tendency with positive anomalies in the equatorial eastern Pacific; therefore, the SST tendency errors occurring in the equatorial eastern Pacific may dominate the uncertainties of the Zebiak–Cane model while simulating CP-El Nino events. A further investigation demonstrates that one of the model errors offset by the OFVs is of a pattern similar to the SST cold-tongue cooling mode, which may then provide one of the climatological conditions for the frequent occurrence of CP-El Nino events. The OFV may therefore be a useful tool for correcting forecast models and then for helping improve the forecast skill of the models.
    Evensen G., 1994: Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99( C5), 10 143- 10 162.10.1029/94JC00572ef8053eecb8d37c88057c7928546f3a5http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F94JC00572%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/94JC00572/citedbySequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error statistics EVENSEN G. Journal of Geophysical Research 99(5), 10143-10162, 1994
    Feng F., W. S. Duan, 2013: The role of constant optimal forcing in correcting forecast models. Science China Earth Sciences, 56( 4), 434- 443.10.1007/s11430-012-4568-za30b82ade78da91c735a01d50cf294f5http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-JDXG201303010.htmhttp://www.cnki.com.cn/Article/CJFDTotal-JDXG201303010.htmIn this paper,the role of constant optimal forcing(COF) in correcting forecast models was numerically studied using the well-known Lorenz 63 model.The results show that when we only consider model error caused by parameter error,which also changes with the development of state variables in a numerical model,the impact of such model error on forecast uncertainties can be offset by superimposing COF on the tendency equations in the numerical model.The COF can also offset the impact of model error caused by stochastic processes.In reality,the forecast results of numerical models are simultaneously influenced by parameter uncertainty and stochastic process as well as their interactions.Our results indicate that COF is also able to significantly offset the impact of such hybrid model error on forecast results.In summary,although the variation in the model error due to physical process is time-dependent,the superimposition of COF on the numerical model is an effective approach to reducing the influence of model error on forecast results.Therefore,the COF method may be an effective approach to correcting numerical models and thus improving the forecast capability of models.
    Guilyardi E., H. Bellenger, M. Collins, S. Ferrett, W. J. Cai, and A. T. Wittenberg, 2012: A first look at ENSO in CMIP5. CLIVAR Exchange, 17, 29- 32.c88676fa67cf88170ce9b2ce482b172fhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F257979825_A_first_look_at_ENSO_in_CMIP5http://www.researchgate.net/publication/257979825_A_first_look_at_ENSO_in_CMIP5The El Ni09o–Southern Oscillation (ENSO) is a naturally occurring fluctuation that originates in the tropical Pacific region with severe weather and societal impacts worldwide (McPhaden et al. 2006). Despite considerable progress in our understanding of the impact
    Ham Y. G., J. S. Kug, 2012: How well do current climate models simulate two types of El Niño? Climate Dyn.,39(1-2), 383-398, doi: 10.1007/s00382-011-1157-3.
    Kalnay E., Coruthors, 1996: The NCEP/NCAR 40-year reanalysis project. Bull. Amer. Meteor. Soc., 77( 4), 437- 471.f539a4fb-a013-4942-ac7e-7f15017eedac23d674534321ec5c56bf181fd85f5561http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%3C0437%253ATNYRP%3E2.0.CO%253B2refpaperuri:(fe1c070047a030c900beb40441caee5a)/s?wd=paperuri%3A%28fe1c070047a030c900beb40441caee5a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%253C0437%253ATNYRP%253E2.0.CO%253B2&ie=utf-8
    Kao H. Y., J. Y. Yu, 2009: Contrasting eastern-Pacific and central-Pacific types of ENSO. J.Climate, 22( 4), 615- 632.10.1175/2008JCLI2309.145c44667cdecca214fcb46319cfa9a89http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093117308.htmlhttp://www.cabdirect.org/abstracts/20093117308.htmlAbstract Surface observations and subsurface ocean assimilation datasets are examined to contrast two distinct types of El Ni帽o鈥揝outhern Oscillation (ENSO) in the tropical Pacific: an eastern-Pacific (EP) type and a central-Pacific (CP) type. An analysis method combining empirical orthogonal function (EOF) analysis and linear regression is used to separate these two types. Correlation and composite analyses based on the principal components of the EOF were performed to examine the structure, evolution, and teleconnection of these two ENSO types. The EP type of ENSO is found to have its SST anomaly center located in the eastern equatorial Pacific attached to the coast of South America. This type of ENSO is associated with basinwide thermocline and surface wind variations and shows a strong teleconnection with the tropical Indian Ocean. In contrast, the CP type of ENSO has most of its surface wind, SST, and subsurface anomalies confined in the central Pacific and tends to onset, develop, and decay in situ. This type of ENSO appears less related to the thermocline variations and may be influenced more by atmospheric forcing. It has a stronger teleconnection with the southern Indian Ocean. Phase-reversal signatures can be identified in the anomaly evolutions of the EP-ENSO but not for the CP-ENSO. This implies that the CP-ENSO may occur more as events or epochs than as a cycle. The EP-ENSO has experienced a stronger interdecadal change with the dominant period of its SST anomalies shifted from 2 to 4 yr near 1976/77, while the dominant period for the CP-ENSO stayed near the 2-yr band. The different onset times of these two types of ENSO imply that the difference between the EP and CP types of ENSO could be caused by the timing of the mechanisms that trigger the ENSO events.
    Kim S. T., J. Y. Yu, A. Kumar, and H. Wang, 2012: Examination of the two types of ENSO in the NCEP CFS model and its extratropical associations. Mon. Wea. Rev., 140( 6), 1908- 1923.10.1175/MWR-D-11-00300.1bd0de78bd16348431738d80e7470980ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012MWRv..140.1908Khttp://adsabs.harvard.edu/abs/2012MWRv..140.1908KAbstract Two types of El Nioouthern Oscillation (ENSO) simulated by the National Centers for Environmental Prediction (NCEP) Climate Forecast System (CFS) model are examined. The model is found to produce both the eastern Pacific (EP) and central Pacific (CP) types of ENSO with spatial patterns and temporal evolutions similar to the observed. The simulated ENSO intensity is comparable to the observed for the EP type, but weaker than the observed for the CP type. Further analyses reveal that the generation of the simulated CP ENSO is linked to extratropical forcing associated with the North Pacific Oscillation (NPO) and that the model is capable of simulating the coupled airea processes in the subtropical Pacific that slowly spreads the NPO-induced SST variability into the tropics, as shown in the observations. The simulated NPO, however, does not extend as far into the deep tropics as it does in the observations and the coupling in the model is not sustained as long as it is in the observations. As a result, the extratropical forcing of tropical central Pacific SST variability in the CFS model is weaker than in the observations. An additional analysis with the Bjerknes stability index indicates that the weaker CP ENSO in the CFS model is also partially due to unrealistically weak zonal advective feedback in the equatorial Pacific. These model deficiencies appear to be related to an underestimation in the amount of the marine stratus clouds off the North American coasts inducing an ocean surface warm bias in the eastern Pacific. This study suggests that a realistic simulation of these marine stratus clouds can be important for the CP ENSO simulation.
    Kug J. S., F. F. Jin, and S. I. An, 2009: Two types of El Niño events: Cold tongue El Niño and warm pool El Niño. J.Climate, 22( 6), 1499- 1515.
    Kug J. S., J. Choi, S. I. An, F. F. Jin, and A. T. Wittenberg, 2010: Warm pool and cold tongue El Niño events as simulated by the GFDL 2.1 coupled GCM. J.Climate, 23( 6), 1226- 1239.10.1175/2009JCLI3293.156c55c26-c682-46bb-a106-284f1d8732b91a0bad1ebddf618704afd9c391af3091http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20103118385.htmlrefpaperuri:(205b7ff4e18f55f36d3fd66f1c2cc0ae)http://www.cabdirect.org/abstracts/20103118385.htmlAbstract Recent studies report that two types of El Ni09o events have been observed. One is the cold tongue (CT) El Ni09o, which is characterized by relatively large sea surface temperature (SST) anomalies in the eastern Pacific, and the other is the warm pool (WP) El Ni09o, in which SST anomalies are confined to the central Pacific. Here, both types of El Ni09o events are analyzed in a long-term coupled GCM simulation. The present model simulates the major observed features of both types of El Ni09o, incorporating the distinctive patterns of each oceanic and atmospheric variable. It is also demonstrated that each type of El Ni09o has quite distinct dynamic processes, which control their evolutions. The CT El Ni09o exhibits strong equatorial heat discharge poleward and thus the dynamical feedbacks control the phase transition from a warm event to a cold event. On the other hand, the discharge process in the WP El Ni09o is weak because of its spatial distribution of ocean dynamic field. The positive SST anomaly of WP El Ni09o is thermally damped through the intensified evaporative cooling.
    Leith C. E., 1978: Objective methods for weather prediction. Annual Review of Fluid Mechanics,10(2), 107-128, doi: 10.1146/annurev.fl.10.010178.000543.10.1146/annurev.fl.10.010178.00054358123377f897797fea8533cceaee06f6http%3A%2F%2Fwww.annualreviews.org%2Fdoi%2Fabs%2F10.1146%2Fannurev.fl.10.010178.000543http://www.annualreviews.org/doi/abs/10.1146/annurev.fl.10.010178.000543ABSTRACT This review deals with the objective methods for weather prediction that have been developed during this century and now supplement the subjective methods that have been used for millenia. From a technical point of view, the problem of weather prediction is especially intriguing in that the atmospheric system is essentially nonlinear and cannot be decomposed into independently acting modes. The inevitable errors in observing the smallest scales of motion must then contaminate larger scales and, in fact, they eventually destroy the accuracy of any prediction. Thus, weather prediction is seen as an unstable problem in the sense that small initial differences have large final consequences. Furthermore, uncertainties both at the beginning and later on in the prediction process require that prediction be treated as a statisti
    Lorenz E. N., 1965: A study of the predictability of a 28-variable atmospheric model. Tellus, 17( 4), 321- 333.9327e0fe-8035-465a-a20c-2187e271a72a5290b716f057ec79a3c6ace9c6fd2bd7http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1965.tb01424.x%2Fpdfrefpaperuri:(3c5589799a6ba39ebdc5944cf81c3457)/s?wd=paperuri%3A%283c5589799a6ba39ebdc5944cf81c3457%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1965.tb01424.x%2Fpdf&ie=utf-8
    Morss R. E., K. A. Emanuel, and C. Snyder, 2001: Idealized adaptive observation strategies for improving numerical weather prediction. J. Atmos. Sci., 58( 3), 210- 232.10.1175/1520-0469(2001)058<0210:IAOSFI>2.0.CO;28c38f303527a1ca601a3a4511a2456c7http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D1812005http://www.ams.org/mathscinet-getitem?mr=1812005Presents a study which examined two simplified adaptive strategies in a simulated idealized system used for improving numerical weather prediction. Description of the three-dimensional quasigeostrophic model and the data assimilation system; Comparison between global adaptive observations and global nonadaptive observations; Capability of the simplified adaptive observations to reduce analysis and weather forecast errors.
    Mu M., W. S. Duan, and B. Wang, 2003: Conditional nonlinear optimal perturbation and its applications. Nonlinear Processes in Geophysics, 10( 6), 493- 501.e25c177dc6379606b920ad8427a3ddedhttp%3A%2F%2Fnsr.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.5194%2Fnpg-10-493-2003%26link_type%3DDOIhttp://nsr.oxfordjournals.org/external-ref?access_num=10.5194/npg-10-493-2003&amp;link_type=DOI
    Rayner N. A., D. E. Parker, E. B. Horton, C. K. Folland , L. V. Alexand er, D. P. Rowell, E. C. Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res., 108(D14),4407, doi: 10.1029/2002JD 002670.10.1029/2002JD0026706c37a263-82ee-4b7e-b9ac-f70e68edd9490831f099871c89699f00bb6e2586346bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2002JD002670%2Ffullrefpaperuri:(bb68c298dcb75e337b3531e27068a3b0)http://onlinelibrary.wiley.com/doi/10.1029/2002JD002670/full[1] We present the Met Office Hadley Centre's sea ice and sea surface temperature (SST) data set, HadISST1, and the nighttime marine air temperature (NMAT) data set, HadMAT1. HadISST1 replaces the global sea ice and sea surface temperature (GISST) data sets and is a unique combination of monthly globally complete fields of SST and sea ice concentration on a 1 latitude-longitude grid from 1871. The companion HadMAT1 runs monthly from 1856 on a 5掳 latitude-longitude grid and incorporates new corrections for the effect on NMAT of increasing deck (and hence measurement) heights. HadISST1 and HadMAT1 temperatures are reconstructed using a two-stage reduced-space optimal interpolation procedure, followed by superposition of quality-improved gridded observations onto the reconstructions to restore local detail. The
    Roads J. O., 1987: Predictability in the extended range. J. Atmos. Sci., 44( 23), 3495- 3527.10.1175/1520-0469(1987)044<3495:PITER>2.0.CO;23b32e88bf2e238abcdb087318316ae2bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1987JAtS...44.3495Rhttp://adsabs.harvard.edu/abs/1987JAtS...44.3495RAbstract A two-level spherical quasi-geostrophic model is formulated for predictability experiments. The stationary external forcing for this model is calculated from observations. Both barotropic and baroclinic forcings are required in order to achieve a realistic model climatology. Realistic transient behavior is also present in the model. The most notable difference is that the observed transient kinetic energy has more energy in the smallestscales. Predictability experiments have an initial rms doubling time of approximately two days. This growth rate along with an initial error of about l/2 the initial error of present operational models produces an rms error equal to the climatological rms error and a correlation of 0.5 on about day 12 of the forecast. At the largest scales, this limiting point is reached shortly thereafter. The error continues to grow at a decreasing rate until at about 30 days the forecast skill is extremely small and comparable to the skill of a persistence forecast. Various time averages at various lags were examined for skill in the extended range. Filters that weighted most strongly the initial forecast days.were shown to provide increased skill. At the furthest limits (60-day time averages), filters improve the skill of prediction by an amount comparable to that which a numerical forecast is an improvement over a persistence forecast. A window filter improves forecasts of time averages by simply eliminating forecast days beyond about day 15. Besides the overall limit, no stable geographical or spectralvariations in the cutoff time could be determined from the limited sample of forecasts described in this paper.
    Talagrand O., 1997: Assimilation of observations, an introduction. J. Meteor. Soc.Japan, 75, 191- 209.10.1175/1520-0469(1997)054<0679:OTRBTS>2.0.CO;2673f22ba6ce06f270d540390d2d8b43dhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F229086740_Assimilation_of_observations_an_introductionhttp://www.researchgate.net/publication/229086740_Assimilation_of_observations_an_introductionABSTRACT Assimilation of meteorological or oceanographical observations can be described as the process through which all the available information is used in order to estimate as accurately as possible the state of the atmospheric or oceanic flow. The available information essentially consists of the observations proper, and of the physical laws which govern the evolution of the flow. The latter are available in practice under the form of a numerical model. The existing assimilation algorithms can be described as either sequential or variational. The links between these algorithms and the theory of statistical estimation are discussed. The performances of present algorithms, and the perspectives for future development, are also briefly discussed.
    Tang Y. M., Z. W. Deng, X. B. Zhou, Y. J. Cheng, and D. K. Chen, 2008: Interdecadal variation of ENSO predictability in multiple models. J.Climate, 21( 18), 4811- 4833.10.1175/2008JCLI2193.108932cd67ceed5ae50a514a1a98261b2http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2008JCli...21.4811Thttp://adsabs.harvard.edu/abs/2008JCli...21.4811TIn this study, El Ni01±o09outhern Oscillation (ENSO) retrospective forecasts were performed for the 120 yr from 1881 to 2000 using three realistic models that assimilate the historic dataset of sea surface temperature (SST). By examining these retrospective forecasts and corresponding observations, as well as the oceanic analyses from which forecasts were initialized, several important issues related to ENSO predictability have been explored, including its interdecadal variability and the dominant factors that control the interdecadal variability. The prediction skill of the three models showed a very consistent interdecadal variation, with high skill in the late nineteenth century and in the middle09恪發ate twentieth century, and low skill during the period from 1900 to 1960. The interdecadal variation in ENSO predictability is in good agreement with that in the signal of interannual variability and in the degree of asymmetry of ENSO system. A good relationship was also identified between the degree of asymmetry and the signal of interannual variability, and the former is highly related to the latter. Generally, the high predictability is attained when ENSO signal strength and the degree of asymmetry are enhanced, and vice versa. The atmospheric noise generally degrades overall prediction skill, especially for the skill of mean square error, but is able to favor some individual prediction cases. The possible reasons why these factors control ENSO predictability were also discussed.
    Toth Z., E. Kalnay, 1997: Ensemble forecasting at NCEP and the breeding method. Mon. Wea. Rev., 125( 12), 3297- 3319.
    Vannitsem S., Z. Toth, 2002: Short-term dynamics of model errors. J. Atmos. Sci., 59( 17), 2594- 2604.9df6e05668950f71c684e94ed3ac3696http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002JAtS...59.2594V/s?wd=paperuri%3A%28fed28bd18ff3c07d1a5f95ebeadb593f%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002JAtS...59.2594V&ie=utf-8
    Xue Y., M. A. Cane, S. E. Zebiak, and M. B. Blumenthal, 1994: On the prediction of ENSO: A study with a low-order Markov model. Tellus A, 46( 5), 512- 528.10.1034/j.1600-0870.1994.00013.xe0451a3c-50b4-4549-ad09-aafa9d753725bd3cb705d57cebad5144a89ad06352bchttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1034%2Fj.1600-0870.1994.00013.x%2Fcitedbyrefpaperuri:(cd7a6e8694ac35228fcc3b5a74c4b6a2)http://onlinelibrary.wiley.com/doi/10.1034/j.1600-0870.1994.00013.x/citedbyA linear model best fit to the Zebiak and Cane (1987) ENSO forecast model (ZC) is used to study the model's prediction skill. Multivariate empirical orthogonal functions (MEOFs) obtained from the sea surface temperature anomaly, sea level and wind stress anomaly fields in a suite of 3-year forecast runs of ZC starting from the monthly initial conditions in the period January 1970 to December 1991, are used to construct a series of seasonally varying linear Markov models. It is found that the model with 18 MEOFs fits the original nonlinear model reasonably well and has comparable or better forecast skill. Assimilating the observed SST into the initial conditions further improves forecast skill at short lead times (&lt;&nbsp;9&nbsp;months). The transient initial error growth in the model's prediction is attributed to the non-self-adjoint property as in Farrell and Blumenthal. Initial error grows fastest starting from spring and slowest starting from late summer and is sensitive to the initial error structures. Two singular vectors (SVs) of the linear evolution operator have significant transient growth dominating the total error growth. Since the optimal perturbation (fastest SV) has mostly high MEOF components, the error growth tends to be larger when there are more high mode components in the initial error fields. This result suggests a way to filter the initial condition fields: the MEOFs higher than the 18th in the initial fields are mostly noise and removing them improves prediction skill. The forecasts starting from late summer have the best predictability because the fastest growth season (summer) is just avoided. The well known, very rapid decline in forecast skill in the boreal spring (the &ldquo;spring barrier&rdquo;) is here attributed to the smallness of the signal to be forecast: the standard deviation of the NIN03 SST anomaly is smallest in spring.
    Yu J. Y., S. T. Kim, 2010: Three evolution patterns of central-Pacific El Niño. Geophys. Res. Lett., 37,L08706, doi: 10.1029/2010GL042810.
    Zebiak S. E., M. A. Cane, 1987: A model El Niño-southern oscillation. Mon. Wea. Rev., 115( 10), 2262- 2278.10.1038/302295a0fbff0f50-8c26-4685-a6ef-b4f44b04cc05c65dd6c79eaaa702eee2a149cb0bddbbhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F244948692_1987_A_model_El-Nino-Southern_Oscillationrefpaperuri:(02aa79a86a12d930aec389c613c4f943)http://www.researchgate.net/publication/244948692_1987_A_model_El-Nino-Southern_OscillationAt intervals that vary from 2 to 10 yr sea-surface temperatures and rainfall are unusually high and the tradewinds are unusually weak over the tropical Pacific Ocean. These Southern Oscillation El Ni帽o events which devastate the ecology of the coastal zones of Ecuador and Peru, which affect the global atmospheric circulation and which can contribute to severe winters over northern America, often develop in a remarkably predictable manner. But the event which began in 1982 has not followed this pattern.
  • [1] Junya HU, Rong-Hua ZHANG, Chuan GAO, 2019: A Hybrid Coupled Ocean-Atmosphere Model and Its Simulation of ENSO and Atmospheric Responses, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 643-657.  doi: 10.1007/s00376-019-8197-8
    [2] XUE Hai-Le, SHEN Xue-Shun, CHOU Ji-Fan, 2013: A Forecast Error Correction Method in Numerical Weather Prediction by Using Recent Multiple-time Evolution Data, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1249-1259.  doi: 10.1007/s00376-013-2274-1
    [3] PENG Yuehua, DUAN Wansuo, XIANG Jie, 2011: Effect of Stochastic MJO Forcing on ENSO Predictability, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 1279-1290.  doi: 10.1007/s00376-011-0126-4
    [4] XUE Haile, SHEN Xueshun, CHOU Jifan, 2015: An Online Model Correction Method Based on an Inverse Problem: Part I——Model Error Estimation by Iteration, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1329-1340.  doi: 10.1007/s00376-015-4261-1
    [5] XUE Haile, SHEN Xueshun, CHOU Jifan, 2015: An Online Model Correction Method Based on an Inverse Problem: Part II——Systematic Model Error Correction, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1493-1503.  doi: 10.1007/s00376-015-4262-0
    [6] ZHU Jieshun, SUN Zhaobo, ZHOU Guangqing, 2007: A Note on the Role of Meridional Wind Stress Anomalies and Heat Flux in ENSO Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 729-738.  doi: 10.1007/s00376-007-0729-y
    [7] Ling-Jiang TAO, Rong-Hua ZHANG, Chuan GAO, 2017: Initial Error-induced Optimal Perturbations in ENSO Predictions, as Derived from an Intermediate Coupled Model, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 791-803.  doi: 10.1007/s00376-017-6266-4
    [8] Xia LIU, Qiang WANG, Mu MU, 2018: Optimal Initial Error Growth in the Prediction of the Kuroshio Large Meander Based on a High-resolution Regional Ocean Model, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1362-1371.  doi: 10.1007/s00376-018-8003-z
    [9] Zhou Tianjun, Yu Rucong, Li Zhaoxin, 2002: ENSO-Dependent and ENSO-Independent Variability over the Mid-Latitude North Pacific: Observation and Air-Sea Coupled Model Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 1127-1147.  doi: 10.1007/s00376-002-0070-4
    [10] Ni Yunqi, Zou Li, Wu Aiming, 1996: Response Process of Ocean to Atmospheric Forcing and Optimal Response Frequency in the CZ Ocean Model, ADVANCES IN ATMOSPHERIC SCIENCES, 13, 43-58.  doi: 10.1007/BF02657027
    [11] Bin MU, Juhui REN, Shijin YUAN, Rong-Hua ZHANG, Lei CHEN, Chuan GAO, 2019: The Optimal Precursors for ENSO Events Depicted Using the Gradient-definition-based Method in an Intermediate Coupled Model, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1381-1392.  doi: 10.1007/s00376-019-9040-y
    [12] Runhua YANG, Jing GUO, Lars Peter RIISH?JGAARD, 2006: Application of an Error Statistics Estimation Method to the PSAS Forecast Error Covariance Model, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 33-44.  doi: 10.1007/s00376-006-0004-7
    [13] Jianguo LIU, Binghao JIA, Zhenghui XIE, Chunxiang SHI, 2016: Ensemble Simulation of Land Evapotranspiration in China Based on a Multi-Forcing and Multi-Model Approach, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 673-684.  doi: 10.1007/s00376-016-5213-0
    [14] Wang Yuan, Wu Rongsheng, 2002: An Optimal Spatial Finite-Difference Operator which Reduces Truncation Error to a Minimum, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 468-486.  doi: 10.1007/s00376-002-0080-2
    [15] Ben TIAN, Hong-Li REN, 2022: Diagnosing SST Error Growth during ENSO Developing Phase in the BCC_CSM1.1(m) Prediction System, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 427-442.  doi: 10.1007/s00376-021-1189-5
    [16] ZHI Hai, ZHANG Rong-Hua, LIN Pengfei, WANG Lanning, 2015: Simulation of Salinity Variability and the Related Freshwater Flux Forcing in the Tropical Pacific: An Evaluation Using the Beijing Normal University Earth System Model (BNU-ESM), ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1551-1564.  doi: 10.1007/s00376-015-4240-6
    [17] Bohua Huang, James L. Kinter III, Paul S. Schopf, 2002: Ocean Data Assimilation Using Intermittent Analyses and Continuous Model Error Correction, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 965-992.  doi: 10.1007/s00376-002-0059-z
    [18] Chen Jiabin, Ji Liren, Wu Wanli, 1987: DESIGN AND TEST OF AN IMPROVED SCHEME FOR GLOBAL SPECTRAL MODEL WITH REDUCED TRUNCATION ERROR, ADVANCES IN ATMOSPHERIC SCIENCES, 4, 156-168.  doi: 10.1007/BF02677062
    [19] Jiawei YAO, Wansuo DUAN, Xiaohao QIN, 2021: Which Features of the SST Forcing Error Most Likely Disturb the Simulated Intensity of Tropical Cyclones?, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 581-602.  doi: 10.1007/s00376-020-0073-z
    [20] DUAN Wansuo, ZHANG Rui, 2010: Is Model Parameter Error Related to a Significant Spring Predictability Barrier for El Nino events? Results from a Theoretical Model, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1003-1013.  doi: 10.1007/s00376-009-9166-4

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 29 July 2015
Manuscript revised: 15 November 2015
Manuscript accepted: 21 December 2015
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 Constant and Time-variant Optimal Forcing Approaches in El Niño Simulations by Using the Zebiak-Cane Model

  • 1. Laboratory for Climate Studies, National Climate Center, China Meteorological Administration, Beijing 100081
  • 2. State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029

Abstract: Model errors offset by constant and time-variant optimal forcing vector approaches (termed COF and OFV, respectively) are analyzed within the framework of El Niño simulations. Applying the COF and OFV approaches to the well-known Zebiak-Cane model, we re-simulate the 1997 and 2004 El Niño events, both of which were poorly degraded by a certain amount of model error when the initial anomalies were generated by coupling the observed wind forcing to an ocean component. It is found that the Zebiak-Cane model with the COF approach roughly reproduced the 1997 El Niño, but the 2004 El Niño simulated by this approach defied an ENSO classification, i.e., it was hardly distinguishable as CP-El Niño or EP-El Niño. In both El Niño simulations, substituting the COF with the OFV improved the fit between the simulations and observations because the OFV better manages the time-variant errors in the model. Furthermore, the OFV approach effectively corrected the modeled El Niño events even when the observational data (and hence the computational time) were reduced. Such a cost-effective offset of model errors suggests a role for the OFV approach in complicated CGCMs.

1. Introduction
  • The uncertainty in forecast results is an important factor in numerical weather forecasting and climate prediction. This uncertainty usually comprises both initial and model errors. Conventionally, much work has explored the initial errors in predictions (Lorenz, 1965; Evensen, 1994; Talagrand, 1997; Toth and Kalnay, 1997; Morss et al., 2001; Mu et al., 2003). Meanwhile, meteorologists seek to improve model predictability by minimizing model errors. Statistical methods dealing with time-variant model errors have been proposed and applied to numerical predictions with varying success (Leith, 1978; Chen et al., 2000; Barreiro and Chang, 2004). Alternatively, (D'andrea and Vautard, 2000) improved the forecast results by a perturbation approach, where an appropriate constant term is added to the tendency equations (also see Roads, 1987; Vannitsem and Toth, 2002). The constant forcing sensitivity vector (termed COF for convenience) proposed by (Barkmeijer et al., 2003) potentially achieves the greatest constant error reduction in the model. Meanwhile, (Feng and Duan, 2013) argued that the COF could be an effective approach to reducing model error even for the time-dependent model error by using a conceptual model. However, the performance of COF in correcting the most realistic models has yet not been shown. On the other side, (Duan et al., 2014) developed the optimal forcing vector (OFV) approach, which offsets the time-variant tendency errors and was shown to be effective in correcting a realistic ENSO model. Subsequently, we naturally ask: what is the difference between the roles of the two approaches in reducing the effects of the model error, especially when applied to the simulations of two types of El Niño [EP-El Niño and CP-El Niño (Ashok et al., 2007; Kao and Yu, 2009; Kug et al., 2009)] by using the most realistic Zebiak-Cane ENSO model (Zebiak and Cane, 1987)?

    According to recent studies, CP-El Niño events are more challenging to simulate than their counterparts, and model errors seem to be an important factor (Kug et al., 2010; Kim et al., 2012; Ham and Kug, 2012; Duan et al., 2014). (Duan et al., 2014) noted that the model errors may exert greater influence in CP-El Niño simulations than in EP-El Niño simulations, and the Zebiak-Cane model has the potential for reproducing EP-El Niño events. Nevertheless, the 1997/98 EP-El Niño event involved a poor simulation. For this EP-El Niño event, although it could be reproduced to some extent in (Duan et al., 2014) with the data assimilation initialization scheme, the simulated development and intensity of this event were still considerably different from the observation and reflected some degree of the model uncertainty.

    In this study, to improve El Niño simulations by the Zebiak-Cane model, we correct the model errors. We investigate and compare error offset by COF and OFV while simulating two types of El Niño events with the Zebiak-Cane model. The uncorrected model could not replicate all features of the 1997 EP-El Niño and 2004 CP-El Niño, and our depictions will focus on these two events. Section 2 introduces the basic concepts of COF and OFV, and presents the related calculations. Section 3 applies both approaches in El Niño simulations and then investigates the time interval of each OFV component in the OFV-based simulations. Specifically, section 3.3 discusses the model uncertainties that influence the El Niño simulations by comparing the spatial structures of the COF and OFV results. The paper concludes with a summary and discussion in section 4.

2. The optimal forcing vector approach
  • The motion of the atmosphere or oceans can be predicted from the following nonlinear partial differential equation: \begin{equation} \label{eq1} \left\{ \begin{array}{l} \dfrac{\partial\textbf{u}}{\partial t}=F(\textbf{u},t) ,\\ \textbf{u}|_{t=0}=\textbf{u}_0 , (1)\end{array} \right. \end{equation} where u(x,t)=[u1(x,t),u2(x,t),…,un(x,t)] is denoted as the state vector, F represents a nonlinear operator, u0 is the initial state, \((x,t)\in\Omega\times[0,\tau]\), Ω is a domain in Rn, \(\tau<+\infty\), x=(x1,x2,…,xn) and t indicates the time. Several errors are associated with this model. Given an initial field u0, Eq. (2) gives the following solution for the state vector u at time τ: \begin{equation} \label{eq2} \textbf{u}(x,\tau)=M_\tau(\textbf{u}_0) . (2)\end{equation} Here, Mτ is the propagator. Let the observations at time 0 and τ be u obs,t0 and u obs,tτ, respectively; the approximate prediction error introduced by the model is then written as \begin{equation} \label{eq3} E_\tau=\|M_\tau(\textbf{u}_{{obs},t_0})-\textbf{u}_{{obs},t_\tau}\| , (3)\end{equation} Here, ||•|| denotes the norm that measures the magnitudes of the prediction errors.

    When f(x) is taken so as to make the simulation generated by Eq. (5) closest to the observation at the terminal time, it can then be referred to as constant optimal external forcing [also known as COF; see (Feng and Duan, 2013)]. (Feng and Duan, 2013) showed that the COF can also reduce time-dependent model errors to a certain extent by using a conceptual model. However, the performance of COF in correcting the most realistic models has not been shown. \begin{equation} \label{eq4} \left\{ \begin{array}{l} \dfrac{\partial \textbf{u}}{\partial t}=F(\textbf{u},t)+f(\textbf{x}) ,\\ \textbf{u}|_{t=0}=\textbf{u}_0 . (4)\end{array} \right. \end{equation} (Duan et al., 2014) reduced the effects of the model errors by superimposing a time-variant external forcing [f(x,t) rather than f(x) in Eq. (5)] that drives the model results toward the observations. The choices of f(x,t) that minimize the differences between the model simulation and the observations constitute an optimization problem. That is, \begin{equation} \label{eq5} J(f_{\min t_i})=\min\|M_{t_{i+1}-t_i}(f_{t_i})(u_{t_i})-u_{{obs},t_{i+1}}\| , (5)\end{equation} where \(t_i,t_i+1\in[t_0,t_k]\). Here, the time window [t0,tk] is similar to the aforementioned [0,τ] and the time interval [ti,ti+1] is not necessary to be a time step of numerical integration, but could represent several days, a month or others. Mti+1-ti(fti) propagates in Eq. (6) from time ti to ti+1, and $u_{t_i}$=$M_{{t_i}-t_{i-1}} $ $(f_{\min,t_{i-1}})$ $(u_{t_{i-1}})$. The OFV (i.e., the optimal f(x,t)), denoted \(f_\min,t_k-t_0=(f_\min,t_0,f_\min,t_1,f_\min,t_2,\ldots,f_\min,t_k-1)\) is then obtained from Eq. (6) as the model simulation that best reproduces the observation during the time window [t0,tk]. Relevant details of OFV calculations are reported in (Duan et al., 2014).

3. Comparison of the constant and time-variant optimal forcing approaches in El Niño simulations
  • As mentioned in the introduction, the Zebiak-Cane model could not replicate all features of the 1997 and 2004 El Niño events. These kinds of difficulties are manifested from the effects of model errors. Following (Duan et al., 2014), here we investigate the COFs of the Zebiak-Cane model and explore the differences between simulations with COFs and OFVs for the observed El Niño events.

  • The Zebiak-Cane model is composed of a Gill-type steady-state linear atmospheric model and a reduced-gravity oceanic model, which depict the thermodynamics and dynamics of the tropical Pacific with oceanic and atmospheric anomalies near the mean climatological state specified from observations. The model has been extensively applied in dynamics and predictability studies of EP-El Niño events (Blumenthal, 1991; Xue et al., 1994; Chen et al., 2004; Tang et al., 2008). However, few studies have simulated CP-El Niño events using this model, largely because the model errors preclude an accurate reproduction of such events. The effects of model errors on the 1997 and 2004 El Niño simulations were highlighted in (Duan et al., 2014), who modeled several El Niño events with a corrected Zebiak-Cane model and emphasized the importance of model error cancellation for ENSO simulation, especially for CP-El Niño reproduction.

    Here, we require the observational data used by (Duan et al., 2014). Specifically, we adopt the SST data from the HadISST analyses datasets (Rayner et al., 2003) and the wind data from the NCEP-NCAR reanalysis products (Kalnay and Coauthors, 1996). The Zebiak-Cane model was initiated using the monthly wind stress anomalies derived from Florida State University analyses (Bourassa et al., 2005). We also borrowed the data of 20°C depth from the NOAA NCEP EMC (Environmental Modeling Center) CMB (Climate Modeling Branch) Pacific (hereafter referred to as the EMC/CMB data) (Behringer et al., 1998).

  • This section analyzes simulations of the El Niño years 1997 and 2004, which were poorly degraded by a certain degree of error in the Zebiak-Cane model. For each El Niño event, we compute the corresponding COF and OFV, respectively. Prior to simulating the El Niño events, the initial anomalies in the Zebiak-Cane model were obtained using the initialization procedure of (Chen et al., 1995), where the model was initialized in a coupled manner by nudging the modeled wind to the observations to some extent. Given the initial anomalous fields, we further calculate the OFV and OFV in relation to the tendency equation (namely, the SST equation) of the model. Subsequently, we superimpose the COF and OFV onto the model and attempt to fit the simulated warm events to their corresponding observations.

    As the time window [t0,tk] associated with the optimal terms (including both the OFV and COF) (see section 2), twelve months prior to the peak phase was selected for each El Niño event and specified just as the simulation period. For instance, the 1997 El Niño event peaked in December 1997; thus, the simulation time window for this event was January-December of 1997. Within the simulation time window, we compute one OFV component per month using the predetermined initial anomaly fields, yielding an OFV with 11 components. In contrast, the COF has a single component because it is time-invariant throughout the time window (see section 2). Considering the COF approach tries to fit the mature phase of the El Niño event to its corresponding observation, it is understandable that the closer to the end of the simulation, the better this approach behaves.

    To acquire the initial fields at the start time t0 of the optimization time window [t0,tk], here we adopt the same initialization procedure as (Chen et al., 1995). And then, based on the initial anomalies, the model integrations with corresponding COF and OFV are obtained for one year and compared directly to the observed El Niño events.

    Figure 1 plots the observed and simulated SST anomaly patterns during the 1997 El Niño year. Both the OFV- and COF-based simulations yield an EP-El Niño event but differ in their similarities to the observations. In the COF evolution, the reproduced SSTA of the first half-year is quite rough, and includes inconspicuous zonal warming along the equator; moreover, the intensity of this event in the mature phase is much weaker than in the unique 1997 EP-El Niño event. The simulations and observations are further compared in Table 1. The SSTA fields modeled by OFV and COF are directly contrasted. In particular, the fields simulated by OFV are strongly correlated with the observed spatial patterns and fit the data very well.

    Figure 1.  The SST anomaly (units: K) component of the seasonal evolutions of the observed 1997 El Niño event (a) and its simulations by the Zebiak-Cane model with OFV (b) and with COF (c). Here, a year is divided into four seasons: January to March (JFM), April to June (AMJ), July to September (JAS), and October to December (OND).

    Figure 2.  The zonal wind anomalies (units: m s$^-1$) in the 1997 El Niño event. The anomalies are averaged from March to July and from August to December. (a) Observed zonal wind, derived from the NCEP-NCAR reanalysis data; (b) zonal wind simulated with OFV; and (c) zonal wind simulated with COF. Panels (d-f) show the corresponding results for the 2004 El Niño event.

    Figure 3.  As in Fig. 2 but showing the thermocline depth anomalies (units: m). The observations are obtained from the EMC/CMB data.

    Figure 4.  (a) COF and (b) OFV simulations (units: K month$^-1$) of the 2004 El Niño event, and (c) their seasonal differences (OFV minus COF). The OFV includes 11 components in the El Niño year, corresponding to 11 time intervals (where one interval ranges from one month to the next). The COF includes a single component that remains constant throughout the year.

    Figure 5.  The SST anomaly (units: K) component of the seasonal evolutions of the observed 2004 El Niño event (a) and its simulations by the Zebiak-Cane model with OFV (b) and with COF (c).

    Figure 6.  The SST anomaly (units: K) components of the seasonal evolutions of the 1997 El Niño event simulated by the Zebiak-Cane model with (a) OFV, (b) S-OFV and (c) T-OFV, respectively.

    Figure 7.  As in Fig. 6 but simulating the 2004 El Niño.

    The other physical variables in the OFV-based simulation, such as the thermocline depth anomaly and zonal wind field, are also shown in accordance with the related observations (Figs. 2 and 3). In practice, the zonal westerly anomalies over the western Pacific present an eastward expansion (see Fig. 2), and ultimately cover a large part of the tropical Pacific. The subsurface water warms and cools in the eastern and western regions of the equatorial Pacific, respectively (see Fig. 3). As the thermocline deepens over the equatorial eastern Pacific, it is conducive to surface warming through warm vertical advection via in-situ mean upwelling (An and Jin, 2001). All of the aforementioned conditions in the OFV-based simulation were observed in the real 1997 EP-El Niño event. In fact, a strong westerly wind burst (WWB) event played an important role in this El Niño event, and whether or not this WWB event in spring can be simulated successfully determines the ultimate performance in simulating the 1997 El Niño (also see Duan et al., 2014). Both the COF and OFV methods help establish the WWB event to varying degrees (Figs. 3b and c), successfully reproducing the sudden occurrence of this event after January-March. In the COF simulation, the relationships between the variables are much weaker. The weakened seesaw pattern is accompanied by weaker basin-scale zonal wind, implying that COF cannot eliminate so much of the time-variant errors in the model.

    Next, the OFV and COF (Fig. 4) approaches in the simulations of the 2004 CP-El Niño event are compared (Fig. 5). As pointed out by (Duan et al., 2014), this El Niño event could hardly be reproduced without the introduction of OFV. Table 1 lists the spatial correlations of the observed versus simulated SST anomaly components during this El Niño year. The SST anomalies in the spatial patterns simulated with OFV are very highly correlated with the observed patterns, whereas those simulated with COF are less correlated with the observations. Moreover, from its appearance, the 2004 event simulated by COF is hardly distinguishable as CP-El Niño or EP-El Niño, which is a critical factor to justify the simulation appearance. Considering that subsurface feedback may rarely affect the evolution of CP-El Niño (Kao and Yu, 2009; Kug et al., 2009; Yu and Kim, 2010; Duan et al., 2014), the OFV-corrected Zebiak-Cane model may not describe the observed thermocline depth variation, although it reproduces the SSTA component of CP-El Niño events well (Duan et al., 2014). This result can also be explained by the accompanying air-sea variables in the 2004 CP-El Niño simulated with COF. The seesaw pattern in the thermocline depth fields (Fig. 3) and the slight eastward propagation of the zonal wind anomalies (Fig. 2) differ from the observations, muddling the classification of the event (which evolves similarly to an EP-El Niño).

    All of these results show that the OFV approach effectively offsets the time-variant model errors, which are better than those handled by COF. Corrected by the OFV method, the Zebiak-Cane model properly reproduces the El Niño events and improves the fit between the simulations and observations. In the COF approach, the model better simulated the 1997 El Niño than the 2004 El Niño. We also tested other El Niño events with the two approaches, as in (Duan et al., 2014), and the comparison of COF and OFV changed little. That is, the model correction always performed better in EP-El Niño simulations than in CP-El Niño simulations, regardless of whether OFV or COF was adopted, and both the COF and OFV approach could further improve the simulation of EP-El Niño. In contrast, the time-dependent errors in the Zebiak-Cane model are particularly severe in CP-El Niño simulations, and the OFV rather than the COF approach could help reproduce a clear classification of the warm events.

  • So why do both approaches help to reproduce the 1997 El Niño event, while for the 2004 El Niño event the simulation benefits greatly from the OFV approach only? The differences between COF and OFV for the 2004 CP-El Niño may help to explain.

    Specifically, in the 2004 El Niño, both the COF and OFV corrections yielded large positive values in the eastern tropical Pacific, suggesting that the uncertainties in the Zebiak-Cane model are dominated by SST tendency errors in this zone (Fig. 4). On the other hand, seasonal differences between OFV and COF are apparent over the one-year evolution period. Throughout the first season, the uncertainties in the SSTA tendency traverse the entire tropical Pacific (Fig. 4b) in the OFV simulations but are lost in the western Pacific in the COF simulations. During spring (April-June), the differential plots (OFV-COF plots; see Fig. 4c) exhibit limited regions of negative values in the eastern Pacific. In this season, it is apparent that the intensity of OFV is much weaker than at other times, while it helps to maintain the zonal gradient of SSTA fields (Fig. 5b). In the latter half of the year, the positive values of COF are much smaller and the differences between OFV and COF cluster around the equator, particularly in the central-eastern Pacific (Figs. 4b and c). Considering the SSTA growth rates related to OFV are nearly the same in the central and eastern Pacific in the latter half of the year (Fig. 4b), the zonal gradient of SSTA evolution (Fig. 5b) remains and helps concentrate warm water westward. Consequently, the SSTA increase in the equatorial central-eastern Pacific follows the zonal gradient of the SSTA field in the same area, favoring the westward march of the warm center from the Pacific east coast (Figs. 5b and c).

  • As described in section 2, the model error effects are offset by a proper time-variant external forcing f(x,t) that matches the simulations with the observations. To cope with the time-variant errors, the OFV method consists of several components within the time window [t0,tk] of the El Niño simulation. Selecting the interval between each of these components is largely subjective but is important for reducing the computational time of the simulations. In the simple Zebiak-Cane model, the time interval can be set to one or several months without affecting the computation. However, when the model tendency errors are offset by the OFV approach concerning a much more complicated model, the computational time may become unacceptable under the calculation conditions. On the one hand, extending the time interval of each OFV component reduces the cost of obtaining the OFV. On the other hand, adopting different time intervals for the OFV components and the different amount of observational data required consequently obscures the simulation performance.

    Within the time window [t0,tk] of the OFV (see section 2), we selected time intervals of 1, 2 and 3 months prior to the peak phase of the El Niño events for the OFV components (the 1-month interval was adopted in section 3.2). Given the predetermined initial anomaly fields and these simulation time windows, we computed one OFV component every month, every second month (January, March, May, July, September, November, December), and every third month (January, April, July, October, December), obtaining OFVs with 11 components, 6 components and 4 components, respectively. For simplicity, we refer to these El Niño simulations as OFV, S-OFV and T-OFV, respectively.

    To compare the SSTA evolutions simulated by OFV, we ran the Zebiak-Cane model forced by the corresponding S-OFV and T-OFV for one year, obtained the simulations of both El Niño events, and checked their differences. The overall evolution features of the 1997 EP-El Niño event (Fig. 6) are reproduced in both S-OFV and T-OFV, although the simulations differ slightly from observations in the first season of the year. That is, the overall performances of S-OFV and T-OFV are comparable to that of OFV. Similar conclusions can be drawn for the 2004 CP-El Niño (Fig. 7). Central warming is observed even in the T-OFV simulation (three-month interval), and the SSTA pattern improves during the latter half of the year (see Table 2), indicating that the OFV-corrected Zebiak-Cane model can reasonably simulate the 1997 and 2004 El Niño events at reduced computational cost.

    Typically, as the model becomes more complicated, the integrations consume increasing amounts of time, regardless of finding the optimal forcing (OFV) for the tendency equation. We found that reducing the input of observational data (and hence the computational time) improved the cost-effectiveness of the Zebiak-Cane simulations. Such cost reductions would benefit the OFV approach in more complicated CGCMs.

4. Conclusion and discussion
  • Model errors are generally time-dynamic; therefore, an error-cancelation method that treats time-variant errors can potentially improve model simulations and predictions. In this study, ENSO simulations by the Zebiak-Cane model were corrected by COF and OFV, and the performances of the corrections were compared. The 1997 and 2004 El Niño events, which were badly reproduced by the uncorrected Zebiak-Cane model to some extent, were better simulated by OFV and COF. But, nonetheless, it was found that the COF correction yields worse fits to El Niño observations compared with the OFV results, mainly because the latter handles time-variant model errors better. Specifically, when the initial anomalies were generated by coupling the observed wind forcing to the ocean component, the Zebiak-Cane model with COF roughly reproduced the 1997 El Niño event; however, the 2004 El Niño simulated by COF defied an ENSO classification and the simulated evolution (warm SST concentrating in the central-eastern Pacific) typified an EP-El Niño event. Furthermore, the OFV approach effectively corrects the El Niño simulation model even when the observational data input (and thus the computational time) is reduced. This cost-effective offset of model errors will favor the OFV approach in more complicated CGCMs.

    Regarding the 2004 El Niño event, the differences in the SSTA patterns obtained by OFV and COF revealed many seasonal variations. Signals of positive SSTA tendency prevailed in the western Pacific at the beginning of the year; in the latter half, positive anomalies were concentrated in the equatorial central-eastern Pacific (see Fig. 4). Therefore, in this event, the SST tendency errors (which dominate the uncertainties in the Zebiak-Cane model) reflect the time-varying characteristics of the model errors. Furthermore, although the same SSTA growth rate introduced by OFV occurs in the central and eastern Pacific during the latter half-year, the warmer background to the west helps to generate stronger feedbacks between SSTAs and wind fields, and favors westward warming from the east coast of the Pacific. Consequently, the 2004 event appears as a CP-El Niño rather than an EP-El Niño.

    In ENSO simulations, the multi-model mean state does not significantly change from CMIP3 to CMIP5, highlighting the potential for model error cancelation (Guilyardi et al., 2012). CP-El Niño simulations are particularly challenged by new feedback (which amplifies biases in the model), uncertain model parameters, and subjective contact with observational data. The present paper proposes an efficient OFV application for model error treatment. Meanwhile, if we consider additional tendency equations of different physical variables including the SSTA equation, the performance of the approach might be further improved. Furthermore, it is necessary to introduce complicated models concerning ENSO simulations, which will properly benefit from the OFV approach.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return