Advanced Search
Article Contents

Evaluation of the Tropical Variability from the Beijing Climate Center's Real-Time Operational Global Ocean Data Assimilation System


doi: 10.1007/s00376-015-4282-9

  • The second-generation Global Ocean Data Assimilation System of the Beijing Climate Center (BCC_GODAS2.0) has been run daily in a pre-operational mode. It spans the period 1990 to the present day. The goal of this paper is to introduce the main components and to evaluate BCC_GODAS2.0 for the user community. BCC_GODAS2.0 consists of an observational data preprocess, ocean data quality control system, a three-dimensional variational (3DVAR) data assimilation, and global ocean circulation model [Modular Ocean Model 4 (MOM4)]. MOM4 is driven by six-hourly fluxes from the National Centers for Environmental Prediction. Satellite altimetry data, SST, and in-situ temperature and salinity data are assimilated in real time. The monthly results from the BCC_GODAS2.0 reanalysis are compared and assessed with observations for 1990-2011. The climatology of the mixed layer depth of BCC_GODAS2.0 is generally in agreement with that of World Ocean Atlas 2001. The modeled sea level variations in the tropical Pacific are consistent with observations from satellite altimetry on interannual to decadal time scales. Performances in predicting variations in the SST using BCC_GODAS2.0 are evaluated. The standard deviation of the SST in BCC_GODAS2.0 agrees well with observations in the tropical Pacific. BCC_GODAS2.0 is able to capture the main features of El Niño Modoki I and Modoki II, which have different impacts on rainfall in southern China. In addition, the relationships between the Indian Ocean and the two types of El Niño Modoki are also reproduced.
  • 加载中
  • Alves O., D. Hudson, M. Balmaseda, and L. Shi, 2011: Seasonal and decadal prediction. Operational Oceanography in the 21st Century, A. Schiller and G. B. Brassington, Eds., Springer, Netherlands, 513- 542.10.1007/978-94-007-0332-2_2098099585-dc25-4327-82be-4b71f38dc4a08ad5a03f24725ef51453148169dc668ahttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F229062403_Seasonal_and_decadal_prediction%3Fev%3Dauth_pubrefpaperuri:(998ef86b3579de47d79a9c8c6ebfa745)http://www.researchgate.net/publication/229062403_Seasonal_and_decadal_prediction?ev=auth_pubThe paper also briefly discusses decadal prediction, for which there is growing demand, particularly in the context of climate change adaptation. Although decadal prediction is still in its infancy, recent development shows promising results, highlighting the role of ocean initial conditions. The initialisation of the ocean for decadal predictions is a major challenge for the next decade.
    Ashok K., S. K. Behera, S. A. Rao, H. Y. Weng, and T. Yamagata, 2007: El Ni帽o 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
    Balmaseda M. A., K. Mogensen, and A. T. Weaver, 2013: Evaluation of the ECMWF ocean reanalysis system ORAS4. Quart. J. Roy. Meteor. Soc., 139, 1132- 1161.10.1002/qj.2063e13b6b23d5a8db75db04cc2154c053f2http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.2063%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/qj.2063/pdfAbstract A new operational ocean reanalysis system (ORAS4) has been implemented at ECMWF. It spans the period 1958 to the present. This article describes its main components and evaluates its quality. The adequacy of ORAS4 for the initialization of seasonal forecasts is discussed, along with the robustness of some prominent climate signals. ORAS4 has been evaluated using different metrics, including comparison with observed ocean currents, RAPID-derived transports, sea-level gauges, and GRACE-derived bottom pressure. Compared to a control ocean model simulation, ORAS4 improves the fit to observations, the interannual variability, and seasonal forecast skill. Some problems have been identified, such as the underestimation of meridional overturning at 26掳N, the magnitude of which is shown to be sensitive to the treatment of the coastal observations. ORAS4 shows a clear and robust shallowing trend of the Pacific Equatorial thermocline. It also shows a clear and robust nonlinear trend in the 0–700 m ocean heat content, consistent with other observational estimates. Some aspects of these climate signals are sensitive to the choice of sea-surface temperature product and the specification of the observation-error variances. The global sea-level trend is consistent with the altimeter estimate, but the partition into volume and mass variations is more debatable, as inferred by discrepancies in the trend between ORAS4- and GRACE-derived bottom pressure.
    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, 1013- 1021.10.1175/1520-0493(1998)126<1013:AICMFE>2.0.CO;213b402e455e96281c31f47f82b9363d9http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249621199_An_Improved_Coupled_Model_for_ENSO_Prediction_and_Implications_for_Ocean_Initialization._Part_I_The_Ocean_Data_Assimilation_Systemhttp://www.researchgate.net/publication/249621199_An_Improved_Coupled_Model_for_ENSO_Prediction_and_Implications_for_Ocean_Initialization._Part_I_The_Ocean_Data_Assimilation_SystemAn improved forecast system has been developed for El Nino--Southern Oscillation (ENSO) prediction at the National Centers for Environmental Prediction. Improvements have been made both to the ocean data assimilation system and to the coupled ocean--atmosphere forecast model. In Part I of a two-part paper the authors describe the new assimilation system. The important changes are 1) the incorporation of vertical variation in the first-guess error variance that concentrates temperature corrections in the thermocline and 2) the overall reduction in the magnitude of the estimated first-guess error. The new system was used to produce a set of retrospective ocean analyses for 1980--95. The new analyses are less noisy than their earlier counterparts and compare more favorably with independent measurements of temperature, currents, and sea surface height variability. Part II of this work presents the results of using these analyses to initialize the coupled forecast model for ENSO prediction.
    Counillon F., I. Bethke, N. Keenlyside, M. Bentsen, L. Bertino, and F. Zheng, 2014: Seasonal-to-decadal predictions with the ensemble Kalman filter and the Norwegian Earth System Model: A twin experiment. Tellus A, 66, 21074.10.3402/tellusa.v66.21074a38838ef211e166aa12a26b8f6fed3b0http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F273707203_Seasonal-to-decadal_predictions_with_the_ensemble_Kalman_filter_and_the_Norwegian_Earth_System_Model_a_twin_experimenthttp://www.researchgate.net/publication/273707203_Seasonal-to-decadal_predictions_with_the_ensemble_Kalman_filter_and_the_Norwegian_Earth_System_Model_a_twin_experimentHere, we firstly demonstrate the potential of an advanced flow dependent data assimilation method for performing seasonal-to-decadal prediction and secondly, reassess the use of sea surface temperature (SST) for initialisation of these forecasts. We use the Norwegian Climate Prediction Model (NorCPM), which is based on the Norwegian Earth System Model (NorESM) and uses the deterministic ensemble Kalman filter to assimilate observations. NorESM is a fully coupled system based on the Community Earth System Model version 1, which includes an ocean, an atmosphere, a sea ice and a land model. A numerically efficient coarse resolution version of NorESM is used. We employ a twin experiment methodology to provide an upper estimate of predictability in our model framework (i.e. without considering model bias) of NorCPM that assimilates synthetic monthly SST data (EnKF-SST). The accuracy of EnKF-SST is compared to an unconstrained ensemble run (FREE) and ensemble predictions made with near perfect (i.e. microscopic SST perturbation) initial conditions (PERFECT). We perform 10 cycles, each consisting of a 10-yr assimilation phase, followed by a 10-yr prediction. The results indicate that EnKF-SST improves sea level, ice concentration, 2 m atmospheric temperature, precipitation and 3-D hydrography compared to FREE. Improvements for the hydrography are largest near the surface and are retained for longer periods at depth. Benefits in salinity are retained for longer periods compared to temperature. Near-surface improvements are largest in the tropics, while improvements at intermediate depths are found in regions of large-scale currents, regions of deep convection, and at the Mediterranean Sea outflow. However, the benefits are often small compared to PERFECT, in particular, at depth suggesting that more observations should be assimilated in addition to SST. The EnKFSST system is also tested for standard ocean circulation indices and demonstrates decadal predictability for Atlantic overturning and sub-polar gyre circulations, and heat content in the Nordic Seas. The system beats persistence forecast and shows skill for heat content in the Nordic Seas that is close to PERFECT.
    de Boyer Mont茅gut., C., G. Madec, A. S. Fischer, A. Lazar, and D. Iudicone, 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology. J. Geophys. Res., 109,C12003, doi: 10.1029/2004JC002378.23c22191494c47848620f34ac37a4a43http%3A%2F%2Fwww.agu.org%2Fpubs%2Fcrossref%2F2004%2F2004JC002378.shtmlhttp://www.agu.org/pubs/crossref/2004/2004JC002378.shtml
    Dong S. F., S. T. Gille, and J. Sprintall, 2007: An assessment of the Southern Ocean mixed layer heat budget. J. Climate,20, 4425-4442, doi: 10.1175/JCLI4259.1.10.1175/JCLI4259.1e466f2c72d8ab0fa4fc00ac077c01420http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F253003137_An_Assessment_of_the_Southern_Ocean_Mixed-Layer_Heat_Budgethttp://www.researchgate.net/publication/253003137_An_Assessment_of_the_Southern_Ocean_Mixed-Layer_Heat_BudgetAbstract The mixed layer heat balance in the Southern Ocean is examined by combining remotely sensed measurements and in situ observations from 1 June 2002 to 31 May 2006, coinciding with the period during which Advanced Microwave Scanning Radiometer-Earth Observing System (EOS) (AMSR-E) sea surface temperature measurements are available. Temperature/salinity profiles from Argo floats are used to derive the mixed layer depth. All terms in the heat budget are estimated directly from available data. The domain-averaged terms of oceanic heat advection, entrainment, diffusion, and airea flux are largely consistent with the evolution of the mixed layer temperature. The mixed layer temperature undergoes a strong seasonal cycle, which is largely attributed to the air鈥搒ea heat fluxes. Entrainment plays a secondary role. Oceanic advection also experiences a seasonal cycle, although it is relatively weak. Most of the seasonal variations in the advection term come from the Ekman advection, in contrast with western boundary current regions where geostrophic advection controls the total advection. Substantial imbalances exist in the regional heat budgets, especially near the northern boundary of the Antarctic Circumpolar Current. The biggest contributor to the surface heat budget error is thought to be the airea heat fluxes, because only limited Southern Hemisphere data are available for the reanalysis products, and hence these fluxes have large uncertainties. In particular, the lack of in situ measurements during winter is of fundamental concern. Sensitivity tests suggest that a proper representation of the mixed layer depth is important to close the budget. Salinity influences the stratification in the Southern Ocean; temperature alone provides an imperfect estimate of mixed layer depth and, because of this, also an imperfect estimate of the temperature of water entrained into the mixed layer from below.
    Dong S. F., S. L. Garzoli, and M. Baringer, 2009: An assessment of the seasonal mixed layer salinity budget in the Southern Ocean. J. Geophys. Res., 114,C12001, doi: 10.1029/2008JC 005258.10.1029/2008JC005258b87c109a584c29cbeb5f9525b8419491http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008JC005258%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2008JC005258/full[1] The seasonal cycle of mixed layer salinity and its causes in the Southern Ocean are examined by combining remotely sensed and in situ observations. The domain-averaged terms of oceanic advection, diffusion, entrainment, and air-sea freshwater flux (evaporation minus precipitation) are largely consistent with the seasonal evolution of mixed layer salinity, which increases from March to October and decreases from November to February. This seasonal cycle is largely attributed to oceanic advection and entrainment; air-sea freshwater flux plays only a minimal role. Both oceanic advection-diffusion and the freshwater flux are negative throughout the year, i.e., reduce mixed layer salinity, while entrainment is positive year-round, reaching its maximum in May. The advection-diffusion term is dominated by Ekman advection. Although the spatial structure of the air-sea freshwater flux and oceanic processes are similar for the steady state, the magnitude of the freshwater flux is relatively small when compared to that of the oceanic processes. The spatial structure of the salinity tendency for each month is also well captured by the sum of the contributions from the air-sea freshwater flux, advection-diffusion, and entrainment processes. However, substantial imbalances in the salinity budget exist locally, particularly for regions with strong eddy kinetic energy and sparse in situ measurements. Sensitivity tests suggest that a proper representation of the mixed layer depth, a better freshwater flux product, and an improved surface salinity field are all important for closing the mixed layer salinity budget in the Southern Ocean.
    D'Ortenzio, F., D. Iudicone, C. de Boyer Montegut, P. Testor, D. Antoine, S. Marullo, R. Santoleri, G. Madec, 2005: Seasonal variability of the mixed layer depth in the Mediterranean Sea as derived from in situ profiles . Geophys. Res. Lett., 32,L12605, doi: 10.1029/2005GL022463.10.1029/2005GL02246314b36f88dde0488821f7c7222bc55560http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005GL022463%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2005GL022463/citedbyA new 0.5 resolution Mediterranean climatology of the mixed layer depth based on individual profiles of temperature and salinity has been constructed. The criterion selected is a threshold value of temperature from a near-surface value at 10 m depth, mainly derived by a method applied on the global (de Boyer Mont茅gut et al., 2004 dBM04). With respect to dBM04, the main differences reside in the absence of spatial interpolation of the final fields and in the improved spatial resolution. These changes to the method are necessary to reproduce the Mediterranean mixed layer's behavior. In the derived climatological maps, the most relevant features of the basin surface circulation are reproduced, as well as the areas prone of the deep water formation are clearly identified. Finally, the role of density in the definition of the mixed layer's differing behaviors between the oriental and the occidental regions of the basin is presented.
    Fu W. W., J. Zhu, and C. X. Yan, 2009a: A comparison between 3DVAR and EnOI techniques for satellite altimetry data assimilation. Ocean Modelling, 26, 206- 216.10.1016/j.ocemod.2008.10.002cc82bbf17a665582251ed9f88c991d09http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS1463500308001546http://www.sciencedirect.com/science/article/pii/S1463500308001546Two conceptually different assimilation schemes, three dimensional variational (3DVAR) assimilation and Ensemble Optimum Interpolation (EnOI) are compared in the context of satellite altimetric data assimilation. Similarities and differences of the two schemes are briefly discussed and their impacts on the model simulation are investigated. With a tropical Pacific ocean model, two assimilation experiments of sea level anomaly (SLA) data from TOPEX/Poseidon are performed for 5 years from 1997 to 2001. Annual mean states of temperature and salinity fields are compared with analysis data and some independent observations. It is found that EnOI generally produces moderate improvements on both temperature and salinity fields, while changes induced by 3DVAR assimilation are strong and vary remarkably in different areas. For instance, 3DVAR tends to excessively modify the temperature field along the thermocline depth and even deteriorate the simulation, but it is more effective than EnOI below the thermocline depth. However, for the salinity field 3DVAR outperforms EnOI nearly for almost the whole layer. As the difference relative to the WOA01 analysis is compared, it is apparently reduced to below 0.3 psu in most areas in the 3DVAR experiment. On the other hand, the pattern of difference in the EnOI experiment resembles that of the simulation and the magnitude is only diminished to some extent. One advantage of EnOI is that it yields more consistent improvements even in areas where there are large model errors. It is more reliable than 3DVAR in such a sense. It is also revealed that the T- S relation plays a very important role in altimetric data assimilation. Further, the distinct performance of the two schemes can be partly accounted for by their inherent assumptions and settings.
    Fu W. W., J. Zhu, C. X. Yan, and H. L. Liu, 2009b: Toward a global ocean data assimilation system based on ensemble optimum interpolation: Altimetry data assimilation experiment. Ocean Dynamics, 59, 587- 602.10.1007/s10236-009-0206-585ad1d9335c0812ee9a8b8b351bd6de8http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2Fs10236-009-0206-5http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s10236-009-0206-5A global ocean data assimilation system based on the ensemble optimum interpolation (EnOI) has been under development as the Chinese contribution to the Global Ocean Data Assimilation Experiment. The system uses a global ocean general circulation model, which is eddy permitting, developed by the Institute of Atmospheric Physics of the Chinese Academy of Sciences. In this paper, the implementation of the system is described in detail. We describe the sampling strategy to generate the stationary ensembles for EnOI. In addition, technical methods are introduced to deal with the requirement of massive memory space to hold the stationary ensembles of the global ocean. The system can assimilate observations such as satellite altimetry, sea surface temperature (SST), in situ temperature and salinity from Argo, XBT, Tropical Atmosphere Ocean (TAO), and other sources in a straightforward way. As a first step, an assimilation experiment from 1997 to 2001 is carried out by assimilating the sea level anomaly (SLA) data from TOPEX/Poseidon. We evaluate the performance of the system by comparing the results with various types of observations. We find that SLA assimilation shows very positive impact on the modeled fields. The SST and sea surface height fields are clearly improved in terms of both the standard deviation and the root mean square difference. In addition, the assimilation produces some improvements in regions where mesoscale processes cannot be resolved with the horizontal resolution of this model. Comparisons with TAO profiles in the Pacific show that the temperature and salinity fields have been improved to varying degrees in the upper ocean. The biases with respect to the independent TAO profiles are reduced with a maximum magnitude of about 0.25掳C and 0.1聽psu for the time-averaged temperature and salinity. The improvements on temperature and salinity also lead to positive impact on the subsurface currents. The equatorial under current is enhanced in the Pacific although it is still underestimated after the assimilation.
    Griffies S. M., M. J. Harrison, R. C. Pacanowski, and A. Rosati, 2003: A technical guide to MOM4. NOAA/Geophysical Fluid Dynamics Laboratory, GFDL Ocean Group Tech. Rep. No. 5, 371 pp.94ea74fdf82b7f26c2f9c7d8a96f1d86http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F245105251_A_Technical_Guide_to_MOM4http://www.researchgate.net/publication/245105251_A_Technical_Guide_to_MOM4Available online at www.gfdl.noaa.gov Information about how to download and run MOM4 can be found at the GFDL Flexible Modeling System (FMS) web site accessible from www.gfdl.noaa.gov. This document was prepared using L ATEX as described by Lamport (1994) and Goosens et al. (1994).CONTENTS I Basics of MOM4 11
    Griffies S.M., Coauthors, 2005: Formulation of an ocean model for global climate simulations. Ocean Science, 1, 45- 79.797feb720ddb4518819b7827baeee2f0http%3A%2F%2Ficesjms.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.5194%2Fos-1-45-2005%26link_type%3DDOIhttp://icesjms.oxfordjournals.org/external-ref?access_num=10.5194/os-1-45-2005&amp;link_type=DOI
    Han G. J., H. L. Fu, X. F. Zhang, W. Li, X. R. Wu, X. D. Wang, and L. X. Zhang, 2013: A global ocean reanalysis product in the China Ocean Reanalysis (CORA) project. Adv. Atmos. Sci.,30, 1621-1631, doi: 10.1007/s00376-013-2198-9.10.1007/s00376-013-2198-963654f8cb65b369990264c42d0e01dd7http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00376-013-2198-9http://d.wanfangdata.com.cn/Periodical_dqkxjz-e201306010.aspxEvaluation showed that compared to the model simulation, the annual mean heat content of the global reanalysis is significantly approaching that of World Ocean Atlas 2009 (WOA09) data. The quality of the global temperature climatology was found to be comparable with the product of Simple Ocean Data Assimilation (SODA), and the major ENSO events were reconstructed. The global and Atlantic meridional overturning circulations showed some similarity as SODA, although significant differences were found to exist. The analysis of temperature and salinity in the current version has relatively larger errors at high latitudes and improvements are ongoing in an updated version. CORA was found to provide a simulation of the subsurface current in the equatorial Pacific with a correlation coefficient beyond about 0.6 compared with the Tropical Atmosphere Ocean (TAO) mooring data. The mean difference of SLAs between altimetry data and CORA was less than 0.1 m in most years.
    Huang B. Y., Y. Xue, and D. W. Behringer, 2008: Impacts of Argo salinity in NCEP Global Ocean Data Assimilation System: The tropical Indian Ocean. J. Geophys. Res., 113,C08002, doi: 10.1029/2007JC004388.10.1029/2007JC0043889408fdf81730ff2b096cc371c4b87a11http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2007JC004388%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2007JC004388/fullAbstract Top of page Abstract 1.Introduction 2.Operational GODAS 3.Experimental GODAS 4.Validation Data 5.Impacts of Argo Salinity 6.Summary Acknowledgments References Supporting Information [1] Salinity profiles collected by the International Argo Project (International Argo Project data are available at TODO: clickthrough URL http://argo.jcommops.org ) since 2000 provide us an unprecedented opportunity to study impacts of salinity data on the quality of ocean analysis, which has been hampered by a lack of salinity observations historically. The operational Global Ocean Data Assimilation System (GODAS) developed at the National Centers for Environmental Prediction (NCEP) assimilates temperature and synthetic salinity profiles that were constructed from temperature and a local T-S climatology. In this study, we assess impacts of replacing synthetic salinity by Argo salinity on the quality of the GODAS ocean analysis with a focus on the tropical Indian Ocean. The study was based on two global ocean analyses for 2001–2006 with (NCEP_Argo) and without (NCEP_Std) inclusion of Argo salinity. The quality of the ocean analyses was estimated by comparing them with various independent observations such as the surface current data from drifters, the salinity data from the Triangle Trans-Ocean Buoy Network moorings, and the sea surface height (SSH) data from satellite altimeters. We found that by assimilating Argo salinity, the biases in the salinity analysis were reduced by 0.6 practical salinity units (psu) in the eastern tropical Indian Ocean and by 1 psu in the Bay of Bengal. Associated with these salinity changes, the zonal current increased by 30–40 cm s 611 toward the east in the central equatorial Indian Ocean during the winter seasons. When verified against drifter currents, the biases of the annually averaged zonal current in the tropical Indian Ocean were reduced by 5–10 cm s 611 , and the root-mean-square error of surface zonal current was reduced by 2–5 cm s 611 . The SSH biases were reduced by 3 cm in the tropical Indian Ocean, the Bay of Bengal, and the Arabian Sea. These results suggest that the Argo salinity plays a critical role in improving salinity analysis, which in turn contributed to improved surface current and sea surface height analyses.
    Hunt B. R., E. J. Kostelich, and I. Szunyogh, 2007: Efficient data assimilation for spatiotemporal chaos: A local ensemble transform Kalman filter. Physica D: Nonlinear Phenomena, 230, 112- 126.10.1016/j.physd.2006.11.008fc3a71e346f460604587cf9e4b122b55http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0167278906004647http://www.sciencedirect.com/science/article/pii/S0167278906004647Data assimilation is an iterative approach to the problem of estimating the state of a dynamical system using both current and past observations of the system together with a model for the system&rsquo;s time evolution. Rather than solving the problem from scratch each time new observations become available, one uses the model to &ldquo;forecast&rdquo; the current state, using a prior state estimate (which incorporates information from past data) as the initial condition, then uses current data to correct the prior forecast to a current state estimate. This Bayesian approach is most effective when the uncertainty in both the observations and in the state estimate, as it evolves over time, are accurately quantified. In this article, we describe a practical method for data assimilation in large, spatiotemporally chaotic systems. The method is a type of &ldquo;ensemble Kalman filter&rdquo;, in which the state estimate and its approximate uncertainty are represented at any given time by an ensemble of system states. We discuss both the mathematical basis of this approach and its implementation; our primary emphasis is on ease of use and computational speed rather than improving accuracy over previously published approaches to ensemble Kalman filtering. We include some numerical results demonstrating the efficiency and accuracy of our implementation for assimilating real atmospheric data with the global forecast model used by the US National Weather Service.
    Kao H. -Y., J. -Y. Yu, 2009: Contrasting Eastern-Pacific and Central-Pacific types of ENSO. J.Climate, 22, 615- 632.10.1175/2008JCLI2309.145c44667cdecca214fcb46319cfa9a89http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093117308.htmlhttp://www.cabdirect.org/abstracts/20093117308.htmlSurface observations and subsurface ocean assimilation datasets are examined to contrast two distinct types of El Nino-Southern 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.
    Kug J. -S., F. -F. Jin, and S. -I. An, 2009: Two types of El Ni帽o events: Cold tongue El Ni帽o and warm pool El Nino. J.Climate, 22, 1499- 1515.
    Larkin N. K., D. E. Harrison, 2005: Global seasonal temperature and precipitation anomalies during El Ni帽o autumn and winter. Geophys. Res. Lett., 32,L16705, doi: 10.1029/ 2005GL022860.10.1029/2005GL0228602bf7ad61-6dd5-403a-b3bf-9e6d4f5f279d756f64535a48342e9fde3d3d5cb1c9a4http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005GL022860%2Ffullrefpaperuri:(6037dd1981a602a2b9fb7adec2fd870d)http://onlinelibrary.wiley.com/doi/10.1029/2005GL022860/full[1] One of the consequences of the new NOAA definition of El Ni09o is the identification of a number of boreal autumns and winters as El Ni09o seasons that are not conventionally so identified. In these periods SST anomalies are concentrated significantly more toward the International Dateline than usual. We show here that the seasonal weather anomalies typically associated with these additional “Dateline El Ni09o” seasons are different in useful respects over much of the world, and suggest that it is useful to treat these as different types of “El Ni09o” for purposes of seasonal weather forecasting.
    Levitus S., 1982: Climatological atlas of the world ocean. NOAA/ ERL GFDL Professional Paper 13,Princeton, N. J., 173 pp.10.1029/EO064i049p00962-022a22d895dc8e0375396f6174f150597chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2FEO064i049p00962-02%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/EO064i049p00962-02/fullClimatological atlas of the world ocean LEVITUS S. NOAA Prof.Paper 13 173, 1982
    Liu Y. M., R. H. Zhang, Y. H. Yin, and T. Niu, 2005: The application of ARGO data to the global ocean data assimilation operational system of NCC. Acta Meteorologica Sinica, 19, 355- 365.70811829cc9bc1a58429f16d8417ddb2http%3A%2F%2Fd.wanfangdata.com.cn%2FPeriodical_qxxb-e200503008.aspxhttp://d.wanfangdata.com.cn/Periodical_qxxb-e200503008.aspxIn this paper, we have preliminarily studied the application of ARGO (Array for Real-time Geostrophic Oceanography) data to the Global Ocean Data Assimilation System of National Climate Center of China (NCC-GODAS), which mainly contains 4 sub-systems such as data preprocessing, real-time wind stress calculating, variational analysis and interpolating, and ocean dynamic model. For the sake of using ARGO data, the relevant adjustment and improvement have been made at the corresponding aspects in the subsystems.Using the observation data from 1981 to 2003 including the ARGO data of 2001 to July. 2003, we have performed a series of numerical experiments on this system. Comparing with the corresponding results of NCEP, It is illustrated that using ARGO data can improve the results of NCC-GODAS in the region of the Middle Pacific, for instance SST, SSTA (SST anomalies), Nino index, sea sub-surface temperature,etc. Furthermore, it is obtained that NCC-GODAS benefits from ARGO data in the other regions such as Atlantic Ocean, Indian Ocean, and extratropical Pacific Ocean much more than in the tropical Pacific.
    Merrifield M. A., M. E. Maltrud, 2011: Regional sea level trends due to a Pacific trade wind intensification. Geophys. Res. Lett., 38,L21605, doi: 10.1029/2011GL049576.10.1029/2011GL049576587f3b5bbc6a5853f9f545b04d717a6ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2011GL049576%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2011GL049576/abstract[1] Over the past two decades, sea level trends have increased in the western tropical Pacific Ocean with rates that are approximately three times the global average. A general circulation model is used to show that the high rates are caused by a gradual intensification of Pacific trade winds since the early 1990s. The modeled sea level change captures the spatial trend pattern in satellite altimeter sea surface heights and the temporal trend shift in tide gauge observations. In addition to the sea level response, the model is used to show how other aspects of the ocean circulation have increased appreciably in amplitude as a consequence of the trade wind intensification, including tropical surface currents, the shallow meridional over-turning circulation, the Equatorial Undercurrent, and the Indonesian Throughflow. These results highlight an ongoing shift in the state of the tropical Pacific Ocean that will continue as long as the trade wind trend persists.
    Merrifield M. A., P. R. Thompson, and M. Lander, 2012: Multidecadal sea level anomalies and trends in the western tropical Pacific. Geophys. Res. Lett., 39,L13602, doi: 10.1029/ 2012GL052032.10.1029/2012GL052032bf871df79f3aebb6abfb7a8836c85d9chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL052032%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2012GL052032/fullAbstract Top of page Abstract 1.Introduction 2.Data and Methods 3.Regional Sea Level Reconstructions 4.Trade Wind Forcing and the PDO 5.Discussion Acknowledgments References [1] Tide gauge data are used to relate low frequency sea level changes over the past 60years in the western tropical Pacific, including a significant positive trend over the past two decades, to Pacific climate indices. Five-year averages of western tropical Pacific sea level are well explained by global sea-level rise and a combination of slowly varying trade wind fluctuations captured by dominant climate indices for the tropical Pacific. Implications for multidecadal sea level variations in the region are considered using reconstructions based on the climate indices.
    Moore J. K., K. Lindsay, S. C. Doney, M. C. Long, and K. Misumi, 2013: Marine ecosystem dynamics and biogeochemical cycling in the community earth system model [CESM1(BGC)]: Comparison of the 1990s with the 2090s under the RCP4. 5 and RCP8. 5 scenarios. J.Climate, 26, 9291- 9312.10.1175/JCLI-D-12-00566.13a42ec12c3ed4183c98dfd84e4d256dehttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F260631029_Marine_Ecosystem_Dynamics_and_Biogeochemical_Cycling_in_the_Community_Earth_System_Model_CESM1%28BGC%29_Comparison_of_the_1990s_with_the_2090s_under_the_RCP4.5_and_RCP8.5_Scenarioshttp://www.researchgate.net/publication/260631029_Marine_Ecosystem_Dynamics_and_Biogeochemical_Cycling_in_the_Community_Earth_System_Model_CESM1(BGC)_Comparison_of_the_1990s_with_the_2090s_under_the_RCP4.5_and_RCP8.5_ScenariosAbstract The authors compare Community Earth System Model results to marine observations for the 1990s and examine climate change impacts on biogeochemistry at the end of the twenty-first century under two future scenarios (Representative Concentration Pathways RCP4.5 and RCP8.5). Late-twentieth-century seasonally varying mixed layer depths are generally within 10 m of observations, with a Southern Ocean shallow bias. Surface nutrient and chlorophyll concentrations exhibit positive biases at low latitudes and negative biases at high latitudes. The volume of the oxygen minimum zones is overestimated. The impacts of climate change on biogeochemistry have similar spatial patterns under RCP4.5 and RCP8.5, but perturbation magnitudes are larger under RCP8.5. Increasing stratification leads to weaker nutrient entrainment and decreased primary and export production (>30% over large areas). The global-scale decreases in primary and export production scale linearly with the increases
    Nidheesh A. G., M. Lengaigne, J. Vialard, A. S. Unnikrishnan, and H. Dayan, 2013: Decadal and long-term sea level variability in the tropical Indo-Pacific Ocean. Climate Dyn.,41, 381-402, doi: 10.1007/s00382-012-1463-4.10.1007/s00382-012-1463-40c766d4f49e10b2f464bdb1526f46937http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-012-1463-4http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s00382-012-1463-4In this study, we analysed decadal and long-term steric sea level variations over 1966–2007 period in the Indo-Pacific sector, using an ocean general circulation model forced by reanalysis winds. The simulated steric sea level compares favourably with sea level from satellite altimetry and tide gauges at interannual and decadal timescales. The amplitude of decadal sea level variability (up to ~502cm standard deviation) is typically nearly half of the interannual variations (up to ~1002cm) and two to three times larger than long-term sea level variations (up to 202cm). Zonal wind stress varies at decadal timescales in the western Pacific and in the southern Indian Ocean, with coherent signals in ERA-40 (from which the model forcing is derived), NCEP, twentieth century and WASWind products. Contrary to the variability at interannual timescale, for which there is a tendency of El Ni09o and Indian Ocean Dipole events to co-occur, decadal wind stress variations are relatively independent in the two basins. In the Pacific, those wind stress variations drive Ekman pumping on either side of the equator, and induce low frequency sea level variations in the western Pacific through planetary wave propagation. The equatorial signal from the western Pacific travels southward to the west Australian coast through equatorial and coastal wave guides. In the Indian Ocean, decadal zonal wind stress variations induce sea level fluctuations in the eastern equatorial Indian Ocean and the Bay of Bengal, through equatorial and coastal wave-guides. Wind stress curl in the southern Indian Ocean drives decadal variability in the south-western Indian Ocean through planetary waves. Decadal sea level variations in the south–western Indian Ocean, in the eastern equatorial Indian Ocean and in the Bay of Bengal are weakly correlated to variability in the Pacific Ocean. Even though the wind variability is coherent among various wind products at decadal timescales, they show a large contrast in long-term wind stress changes, suggesting that long-term sea level changes from forced ocean models need to be interpreted with caution.
    Qiu B., S. M. Chen, 2012: Multidecadal sea level and gyre circulation variability in the Northwestern Tropical Pacific Ocean. J. Phys. Oceanogr., 42, 193- 206.10.1175/JPO-D-11-061.17aadb120002b5234f0220066652163dehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FJPO-D-11-061.1http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1175/JPO-D-11-061.1Abstract Sea level rise with the trend >10 mm yr 611 has been observed in the tropical western Pacific Ocean over the 1993–2009 period. This rate is 3 times faster than the global-mean value of the sea level rise. Analyses of the satellite altimeter data and repeat hydrographic data along 137°E reveal that this regionally enhanced sea level rise is thermosteric in nature and vertically confined to a patch in the upper ocean above the 12°C isotherm. Dynamically, this regional sea level trend is accompanied by southward migration and strengthening of the North Equatorial Current (NEC) and North Equatorial Countercurrent (NECC). Using a 105-layer reduced-gravity model forced by the ECMWF reanalysis wind stress data, the authors find that both the observed sea level rise and the NEC/NECC’s southward migrating and strengthening trends are largely attributable to the upper-ocean water mass redistribution caused by the surface wind stresses of the recently strengthened atmospheric Walker circulation. Based on the long-term model simulation, it is further found that the observed southward migrating and strengthening trends of the NEC and NECC began in the early 1990s. In the two decades prior to 1993, the NEC and NECC had weakened and migrated northward in response to a decrease in the trade winds across the tropical Pacific Ocean.
    Ratheesh S., R. Sharma, and S. Basu, 2014: An EnOI assimilation of satellite data in an Indian Ocean circulation model. IEEE Transactions on Geoscience and Remote Sensing, 52, 4106- 4111.10.1109/TGRS.2013.22796065e7423c9d6098af47d764d5a83997cdbhttp%3A%2F%2Fieeexplore.ieee.org%2Fxpls%2Fabs_all.jsp%3Farnumber%3D6595592http://ieeexplore.ieee.org/xpls/abs_all.jsp?arnumber=6595592Ensemble optimal interpolation, a simplified and computationally inexpensive version of ensemble Kalman filter, has been used to assimilate satellite-derived sea level anomaly and sea surface temperature in an Indian Ocean circulation model. In order to cut off the long-range spurious covariances, a localization technique has also been used where an assumption is made that only measurements located within a certain influence radius of a grid point can affect the analysis at that grid point. It has been found that the analyses fit the assimilated observations quite well. For a more stringent test of the power of the assimilation technique, selected regions have been demarcated in the analysis area. Simulated mixed layer depth, depth of the 20C isotherm, and temperature at 400-m depth have been compared with observations of these variables in these regions, and assimilation has been found to exhibit significant positive impact. Simulated surface and subsurface temperatures have been compared at isolated Research moored array for African-Asian-Australian monsoon analysis and prediction (RAMA) buoy locations, and again, the high positive impact of assimilation has been evident from these comparisons. Finally, the impact of assimilation on surface current simulation at selected RAMA buoy locations has been found to be somewhat marginal, and a significant improvement apparently needs additional assimilation of drifter data with a very high concentration of drifters.
    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., 108,4407, doi: 10.1029/2002JD002670.882a53bb-2210-4d8f-aa4e-e4a63415f6fa319ab0b7c80c1c4f801c5459d61ea21ahttp%3A%2F%2Fciteseer.ist.psu.edu%2Fshowciting%3Fcid%3D13840372refpaperuri:(3edd199780a4d7721957dbefdec34a0b)http://citeseer.ist.psu.edu/showciting?cid=13840372
    Ren L., K. Speer, and E. P. Chassignet, 2011: The mixed layer salinity budget and sea ice in the Southern Ocean. J. Geophys. Res., 116,C08031, doi: 10.1029/2010JC006634.10.1029/2010JC006634a0bc905cafe2c220cf2a288ccac9a6e3http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JC006634%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2010JC006634/full[1] The seasonal variation of the mixed layer salinity budget in the Southern Ocean is evaluated over the latitude range 45°S–62°S using Argo profiling float data, freshwater fluxes (evaporation minus precipitation (E-P)), geostrophic velocity, wind stress, and sea ice concentration observations. The seasonal cycle of the mixed layer salinity is driven by seasonality in E-P, Ekman advection, entrainment, and sea ice. Over large areas, the geostrophic advection and diffusion show smaller contributions to the seasonal variation relative to other terms. The air-sea freshwater flux and Ekman advection in this area generally result in net decreases in salinity, while the entrainment term yields increases. Residual imbalance is consistent with a sea ice effect, whose contribution is evaluated. Sea ice is found to make a significant contribution, growing in importance toward the ice edge.
    Reynolds R. W., D. C. Marsico, 1993: An improved real-time global sea surface temperature analysis. J.Climate, 6, 114- 119.10.1175/1520-0442(1993)006<0114:AIRTGS>2.0.CO;2eed548f761596a34d51192f25f9976achttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F255928833_An_improved_real-time_global_sea_surface_temperature_analysishttp://www.researchgate.net/publication/255928833_An_improved_real-time_global_sea_surface_temperature_analysisThe June 1991 eruptions of Mount Pinatubo produced new stratospheric aerosols that were greater than the aerosols from the 1982 eruptions of El Chichon. These new aerosols strongly affected the advanced very high resolution radiometer (AVHRR) retrievals of sea surface temperature in the tropics where occurred with magnitudes greater than 1[degrees]C. The time dependence of these biases are shown. In addition, a method to correct these biases is discussed and integrated into the National Meteorological Center's optimum interpolation sea surface temperature analysis.
    Reynolds R. W., T. M. Smith, C. Y. Liu, D. B. Chelton, K. S. Casey, and M. G. Schlax, 2007: Daily high-resolution-blended analyses for sea surface temperature. J.Climate, 20, 5473- 5496.10.1175/2007JCLI1824.16f04d906d78600c9650c319dd60c30a7http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F235641869_Daily_high-resolution_blended_analyseshttp://www.researchgate.net/publication/235641869_Daily_high-resolution_blended_analysesAbstract Two new high-resolution sea surface temperature (SST) analysis products have been developed using optimum interpolation (OI). The analyses have a spatial grid resolution of 0.25° and a temporal resolution of 1 day. One product uses the Advanced Very High Resolution Radiometer (AVHRR) infrared satellite SST data. The other uses AVHRR and Advanced Microwave Scanning Radiometer (AMSR) on the NASA Earth Observing System satellite SST data. Both products also use in situ data from ships and buoys and include a large-scale adjustment of satellite biases with respect to the in situ data. Because of AMSR’s near-all-weather coverage, there is an increase in OI signal variance when AMSR is added to AVHRR. Thus, two products are needed to avoid an analysis variance jump when AMSR became available in June 2002. For both products, the results show improved spatial and temporal resolution compared to previous weekly 1° OI analyses. The AVHRR-only product uses Pathfinder AVHRR data (currently available from January 1985 to December 2005) and operational AVHRR data for 2006 onward. Pathfinder AVHRR was chosen over operational AVHRR, when available, because Pathfinder agrees better with the in situ data. The AMSR–AVHRR product begins with the start of AMSR data in June 2002. In this product, the primary AVHRR contribution is in regions near land where AMSR is not available. However, in cloud-free regions, use of both infrared and microwave instruments can reduce systematic biases because their error characteristics are independent.
    Saji N. H., B. N. Goswami, P. N. Vinayachand ran, and T. Yamagata, 1999: A dipole mode in the tropical Indian Ocean. Nature, 401, 360- 363.10.1038/438541686210831d5bd363a994efe19550bc4fc1839b4http%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM16862108http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM16862108Abstract For the tropical Pacific and Atlantic oceans, internal modes of variability that lead to climatic oscillations have been recognized, but in the Indian Ocean region a similar ocean-atmosphere interaction causing interannual climate variability has not yet been found. Here we report an analysis of observational data over the past 40 years, showing a dipole mode in the Indian Ocean: a pattern of internal variability with anomalously low sea surface temperatures off Sumatra and high sea surface temperatures in the western Indian Ocean, with accompanying wind and precipitation anomalies. The spatio-temporal links between sea surface temperatures and winds reveal a strong coupling through the precipitation field and ocean dynamics. This air-sea interaction process is unique and inherent in the Indian Ocean, and is shown to be independent of the El Ni帽o/Southern Oscillation. The discovery of this dipole mode that accounts for about 12% of the sea surface temperature variability in the Indian Ocean--and, in its active years, also causes severe rainfall in eastern Africa and droughts in Indonesia--brightens the prospects for a long-term forecast of rainfall anomalies in the affected countries.
    Sakov P., F. Counillon, L. Bertino, K. A. Lis忙ter, P. R. Oke, and A. Korablev, 2012: TOPAZ4: An ocean-sea ice data assimilation system for the North Atlantic and Arctic. Ocean Science Discussions, 9, 1519- 1575.10.5194/os-8-633-201281523cbab2b527404e30922d91d2be2bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258687770_TOPAZ4_an_ocean-sea_ice_data_assimilation_system_for_the_North_Atlantic_and_Arctichttp://www.researchgate.net/publication/258687770_TOPAZ4_an_ocean-sea_ice_data_assimilation_system_for_the_North_Atlantic_and_ArcticWe present a detailed description of TOPAZ4, the latest version of TOPAZ - a coupled ocean-sea ice data assimilation system for the North Atlantic Ocean and Arctic. It is the only operational, large-scale ocean data assimilation system that uses the ensem- ble Kalman filter. This means that TOPAZ features a time-evolving, state-dependent estimate of the state error covariance. Based on results from the pilot MyOcean reanalysis for 2003-2008, we demonstrate that TOPAZ4 produces a realistic estimate of the ocean circulation and the sea ice. We find that the ensemble spread for temperature and sea-level remains fairly constant throughout the reanalysis demonstrating that the data assimilation system is robust to ensemble collapse. Moreover, the ensemble spread for ice concentration is well correlated with the actual errors. This indicates that the ensemble statistics provide reliable state-dependent error estimates - a feature that is unique to ensemble-based data assimilation systems. We demonstrate that the quality of the reanalysis changes when different sea surface temperature products are assimilated, or when in situ profiles below the ice in the Arctic Ocean are assimilated. We find that data assimilation improves the match to independent observations compared to a free model. Improvements are particularly noticeable for ice thickness, salinity in the Arctic, and temperature in the Fram Strait, but not for transport estimates or underwater temperature. At the same time, the pilot reanalysis has revealed sev20 eral flaws in the system that have degraded its performance. Finally, we show that a simple bias estimation scheme can effectively detect the seasonal or constant bias in temperature and sea-level.
    Sall茅e, J. B., K. Speer, R. Morrow, R. Lumpkin, 2008: An estimate of Lagrangian eddy statistics and diffusion in the mixed layer of the Southern Ocean. J. Mar. Res., 66( 4), 441- 463.10.1357/0022240087871574580a2124c12f6b3d8f2f2a27effeb331ebhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F228372734_An_estimate_of_Lagrangian_eddy_statistics_and_diffusion_in_the_mixed_layer_of_the_Southern_Ocean%3Fev%3Dauth_pubhttp://www.researchgate.net/publication/228372734_An_estimate_of_Lagrangian_eddy_statistics_and_diffusion_in_the_mixed_layer_of_the_Southern_Ocean?ev=auth_pubA statistical analysis of surface drifter observations is used to compute eddy length and time scales and eddy diffusion in the Southern Ocean. Eddy diffusion values of the order of 10 4 m 2 s –1 are found in the energetic western boundary currents north of the Antarctic Circumpolar Current (ACC) and secondary peaks occur where the ACC negotiates topography. The diffusivity shows an increase from the Antarctic continent to the core of the ACC, then a slight decrease or a stable plateau within the ACC. North of the ACC, diffusivity generally decreases into the interior of ocean basins, except in the western boundary regions where values are maximum. Diffusivity is also calculated from simulated trajectories based on altimetric geostrophic velocities, with and without mean flow, as well as with simulated trajectories based on Ekman currents. Ekman currents at the drogue depth (15 m) have only a small impact, and the geostrophic currents dominate the eddy diffusivity. Complementary statistical analyses confirm these results. The surface drifter cross-stream eddy diffusion is used to test a simple parameterization based on satellite altimetric observations of eddy kinetic energy (EKE). For EKE ≥ 0.015 m 2 s –2 , K = 1.35√ EKEL d m 2 s –1 , where L d is the first baroclinic Rossby radius. This parameterization holds in the energetic ACC, consistent with an eddy field in the “frozen field” regime. Over the broader areas of weaker eddy fields, mixing is fairly uniform and stable at about K = 1800 ± 1000 m 2 s –1 .
    Smith T. M., R. W. Reynolds, T. C. Peterson, and J. Lawrimore, 2008: Improvements to NOAA's historical merged land-ocean surface temperature analysis (1880-2006). J.Climate, 21, 2283- 2296.
    Talley L. D., 1993: Distribution and formation of North Pacific intermediate water. J. Phys. Oceanogr., 23, 517- 537.10.1175/1520-0485(1993)0232.0.CO;2518812044886db74cf77b0cace84cd3bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F238012528_Distribution_and_Formation_of_North_Pacific_Intermediate_Waterhttp://www.researchgate.net/publication/238012528_Distribution_and_Formation_of_North_Pacific_Intermediate_WaterAbstract The North Pacific Intermediate Water (NPIW), defined as the main salinity minimum in the subtropical North Pacific, is examined with respect to its overall property distributions. These suggest that NPIW is formed only in the northwestern subtropical gyre; that is, in the mixed water region between the Kuroshio Extension and Oyashio front. Subsequent modification along its advective path increases its salinity and reduces its oxygen. The mixed water region is studied using all bottle data available from the National Oceanographic Data Center, with particular emphasis on several winters. Waters from the Oyashio, Kuroshio, and the Tsugaru Warm Current influence the mixed water region, with a well-defined local surface water mass formed as a mixture of the surface waters from these three sources. Significant salinity minima in the mixed water region are grouped into those that are directly related to the winter surface density and are found at the base of the oxygen-saturated surface layer, and those that form deeper, around warm core rings. Both could be a source of the more uniform NPIW to the east, the former through preferential erosion of the minima from the top and the latter through simple advection. Both sources could exist all year with a narrowly defined density range that depends on winter mixed-layer density in the Oyashio region.
    Wang C. Z., X. Wang, 2013: Classifying El Nino Modoki I and II by different impacts on rainfall in Southern China and typhoon tracks. J.Climate, 26, 1322- 1338.10.1175/JCLI-D-12-00107.12752f5bc-97b3-4a38-8168-a74d7dd4eaa3700202a262be0511e6cbaa5a283819b6http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258795629_Classifying_El_Nio_Modoki_I_and_II_by_Different_Impacts_on_Rainfall_in_Southern_China_and_Typhoon_Tracksrefpaperuri:(ea405e4ef356f90719c942d6aec7adaa)http://www.researchgate.net/publication/258795629_Classifying_El_Nio_Modoki_I_and_II_by_Different_Impacts_on_Rainfall_in_Southern_China_and_Typhoon_TracksBased on their opposite influences on rainfall in southern China during boreal fall, this paper classifies El Ni01±o Modoki into two groups: El Ni01±o Modoki I and II, which show different origins and patterns of SST anomalies. The warm SST anomalies originate in the equatorial central Pacific and subtropical northeastern Pacific for El Ni01±o Modoki I and II, respectively. Thus, El Ni01±o Modoki I shows a symmetric SST anomaly distribution about the equator with the maximum warming in the equatorial central Pacific, whereas El Ni01±o Modoki II displays an asymmetric distribution with the warm SST anomalies extending from the northeastern Pacific to the equatorial central Pacific. Additionally, the warm SST anomalies in the equatorial central Pacific extend farther westward for El Ni01±o Modoki II than for El Ni01±o Modoki I. Similar to the canonical El Ni01±o, El Ni01±o Modoki I is associated with an anomalous anticyclone in the Philippine Sea that induces southwesterly wind anomalies along the south coast of China and carries the moisture for increasing rainfall in southern China. For El Ni01±o Modoki II, an anomalous cyclone resides east of the Philippines, associated with northerly wind anomalies and a decrease in rainfall in southern China. The canonical El Ni01±o and El Ni01±o Modoki I are associated with a westward extension of the western North Pacific subtropical high (WNPSH), whereas El Ni01±o Modoki II shifts the WNPSH eastward. Differing from canonical El Ni01±o and El Ni01±o Modoki I, El Ni01±o Modoki II corresponds to northwesterly anomalies of the typhoon steering flow, which are unfavorable for typhoons to make landfall in China.
    Wang D. X., Y. H. Qin, X. J. Xiao, Z. Q. Zhang, and X. Y. Wu, 2012a: El Ni帽o and El Ni帽o Modoki variability based on a new ocean reanalysis. Ocean Dynamics, 62, 1311- 1322.
    Wang D. X., Y. H. Qin, X. J. Xiao, Z. Q. Zhang, and F. M. Wu, 2012b: Preliminary results of a new global ocean reanalysis. Chinese Science Bulletin, 57, 3509- 3517.10.1007/s11434-012-5232-x3c5be697-a6ca-47c2-82c5-b5d5b88d46168b4348750f69a57e62272c09e443457ehttp%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-JXTW201226018.htmrefpaperuri:(0da3fa583a16925a89fad6426a46992e)http://www.cnki.com.cn/Article/CJFDTotal-JXTW201226018.htmABSTRACT Using a new global ocean reanalysis of the second generation Global Ocean Data Assimilation System of the Beijing Climate Center (BCC_GODAS2.0) spanning the period 1990&ndash;2009, we firstly quantify the accuracy of BCC_GODAS2.0 in representing the temperature and salinity by comparing with OISST and SODA data. The results show that the assimilation system may effectively improve the estimations of temperature and salinity by assimilating all kinds of observations, especially in the equatorial eastern Pacific. Moreover, the root mean square errors of monthly temperature and salinity are respectively reduced by 0.53掳C and 0.28 psu, compared with the model control simulation results. Then, the applicability of this ocean reanalysis for sea surface temperature (SST) anomaly variability in the tropical Pacific is evaluated with the observational HadISST data. The NINO3 index of the new reanalysis shows a good agreement with that of HadISST, with a correlation of 93.6%. Variations in SST from BCC_GODAS2.0 are similar to those obtained from HadISST data along the equator, showing the major large zonal-scale features such as the strong magnitude of seasonal cycle. The amplitude of SST anomaly standard deviation in the equatorial eastern Pacific is also closer to observations (HadISST) than NCEP GODAS does. Besides, the first two leading empirical orthogonal function (EOF) modes of the monthly SST anomalies over the tropical Pacific region are explored. The EOF1 pattern of BCC_GODAS2.0 captures a traditional El Ni&ntilde;o pattern, which improves magnitudes of the positive SST anomaly in the cold tongue of the eastern Pacific. The EOF2 pattern exhibits a El Ni&ntilde;o Modoki pattern. Comparatively, the EOF2 pattern of BCC_GODAS2.0 extends more strongly toward the subtropics. It also overcomes the problem that negative loadings are confined in the narrow equatorial eastern Pacific. Consequently, the magnitude and spatial distribution of the leading EOF patterns of BCC_GODAS2.0 are well consistent with those of HadISST.
    Wang X., D. Wang, and W. Zhou, 2009: Decadal variability of twentieth-century El Ni帽o and La Nina occurrence from observations and IPCC AR4 coupled models. Geophysical research letters, 36, L11701.
    Wang X., C. Z. Wang, 2014: Different impacts of various El Ni帽o events on the Indian Ocean Dipole. Climate Dyn., 42, 991- 1005.
    Wu, T. W., Coauthors, 2013: Progress in developing the short-range operational climate prediction system of China national climate center. Journal of Applied Meteorological Science, 24, 533- 543. (in Chinese)917289f7aacebe562fda693514680c14http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-YYQX201305004.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-YYQX201305004.htmThe progress in developing the second-generation short-range climate forecast system of National(or Beijing) Climate Center(NCC or BCC) is introduced,focusing on four items,i.e.,the global ocean data assimilation system,the land data assimilation system,the monthly-scale dynamical extended-range forecast system(DERF),and the seasonal climate forecast system.With a better assimilation of temperature and salinity than the first-generation system,the second-generation ocean data assimilation system is now at the quasi-operational level.The land data assimilation system is still under development,but the multisource precipitation merging subsystem is now quasi-operational and can produce reanalysis of precipitation as a forcing to land system.The atmospheric general circulation model BCC_AGCM2.2 and the climate system model BCC_CSM1.1(m) are the main tools for the second-generation monthly-scale DERF and the second-generation seasonal prediction system,respectively.The former has entered quasi-operational use since middle August of 2012 and conducted four-member real-time forecast jobs and 80 hindcast jobs every day,and the latter will enter its quasi-operational stage by the end of 2013.A preliminary evaluation indicates that the second-generation system shows a certain capability in predicting the pentad,ten-day,monthly,seasonal and inter-annual climate variability.It exhibits a higher prediction skill,compared to the first-generation system,in terms of precipitation,surface air temperature,atmospheric circulation and El Nino-Southern Oscillation,and so on.As shown by the hindcasts by two generations of DERF(i.e.,DERF1.0 and DERF2.0) for the monthly mean surface air temperature in January and July,DERF2.0shows overall higher prediction skill than DERFl.0,especially over the tropical Indian Ocean and Pacific and most mid-high latitude areas in the Northern Hemisphere in January,and most regions in global tropics and subtropics in July.Also,the 20-year hindcasts initialized in the end of February of each year by the two generations of seasonal climate prediction system indicate that,the second-generation system shows significant prediction skill of surface air temperature over most areas in spring,especially over the tropical Pacific,Atlantic and Indian Ocean.In contrast,the skills over most areas of the first-generation system are relative lower.
    Wu, T. W., Coauthors, 2014: An overview of BCC climate system model development and application for climate change studies. Journal of Meteorological Research, 28, 34- 56.
    Xiao X. J., D. X. Wang, C. X. Yan, and J. Zhu, 2008: Evaluation of a 3dVAR system for the South China Sea. Progress in Natural Science, 18, 547- 554.10.1016/j.pnsc.2007.12.0079d76cca7-8290-4528-9082-63016e8249c90b26004a985d87992b35bd112c6072bfhttp%3A%2F%2Fwww.cqvip.com%2FQK%2F85882X%2F200805%2F27158114.htmlrefpaperuri:(0d54933678c47b19aeadb728b91b823f)http://en.cnki.com.cn/Article_en/CJFDTOTAL-ZKJY200805005.htmThe authors evaluate a three-dimensional variational (3dVAR) system for the South China Sea (SCS) in this study. The assimilation method applied in the system takes into consideration error correlation along each ground track and uses recursive lter for optimiza- tion. Data from three R/V cruises during the spring and summer of 1998 and the summer of 2000 are used to evaluate the system. The root-mean-square error and bias are reduced signi cantly and when the altimeter data are assimilated, the distribution of the error is much closer to the Gaussian distribution. Precipitation and river discharge in the southwestern SCS are reproduced, and the variability of sea surface height is e ciently transferred to the subsurface. The 3dVAR system performs well for each of the three cruises, suggesting that it is steady for routine usage.
    Xue Y., B. Y. Huang, Z. -Z. Hu, A. Kumar, C. H. Wen, D. Behringer, and S. Nadiga, 2011: An assessment of oceanic variability in the NCEP climate forecast system reanalysis. Climate Dyn., 37, 2511- 2539.10.1007/s00382-010-0954-4744247e4da36bc31b2e33d279dc1da3chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2Fs00382-010-0954-4http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s00382-010-0954-4At the National Centers for Environmental Prediction (NCEP), a reanalysis of the atmosphere, ocean, sea ice and land over the period 1979-2009, referred to as the climate forecast system reanalysis (CFSR), was recently completed. The oceanic component of CFSR includes many advances: (a) the MOM4 ocean model with an interactive sea-ice, (b) the 6 h coupled model forecast as the first guess, (c) inclusion of the mean climatological river runoff, and (d) high spatial (0.5 0.5) and temporal (hourly) model outputs. Since the CFSR will be used by many in initializing/validating ocean models and climate research, the primary motivation of the paper is to inform the user community about the saline features in the CFSR ocean component, and how the ocean reanalysis compares with in situ observations and previous reanalysis. The net ocean surface heat flux of the CFSR has smaller biases compared to the sum of the latent and sensible heat fluxes from the objectively analyzed air-sea fluxes (OAFlux) and the shortwave and longwave radiation fluxes from the International Satellite Cloud Climatology Project (ISCCP-FD) than the NCEP/NCAR reanalysis (R1) and NCEP/DOE reanalysis (R2) in both the tropics and extratropics. The ocean surface wind stress of the CFSR has smaller biases and higher correlation with the ERA40 produced by the European Centre for Medium-Range Weather Forecasts than the R1 and R2, particularly in the tropical Indian and Pacific Ocean. The CFSR also has smaller errors compared to the QuickSCAT climatology for September 1999 to October 2009 than the R1 and R2. However, the trade winds of the CFSR in the central equatorial Pacific are too strong prior to 1999, and become close to observations once the ATOVS radiance data are assimilated in late 1998. A sudden reduction of easterly wind bias is related to the sudden onset of a warm bias in the eastern equatorial Pacific temperature around 1998/1999. The sea surface height and top 300 m heat content (HC300) of the CFSR compare with observations better than the GODAS in the tropical Indian Ocean and extratropics, but much worse in the tropical Atlantic, probably due to discontinuity in the deep ocean temperature and salinity caused by the six data streams of the CFSR. In terms of climate variability, the CFSR provides a good simulation of tropical instability waves and oceanic Kelvin waves in the tropical Pacific, and the dominant modes of HC300 that are associated with El Nino and Southern Oscillation, Indian Ocean Dipole, Pacific Decadal Oscillation and Atlantic Meridional Overturning Circulation.
    Yan C. X., J. Zhu, R. F. Li, and G. Q. Zhou, 2004: Roles of vertical correlations of background error and T-S relations in estimation of temperature and salinity profiles from sea surface dynamic height. J. Geophys. Res., 109,C08010, doi: 10.1029/2003JC002224.10.1029/2003JC0022244a93ce0e00833653ae33cfcaa9ecfc3bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2003JC002224%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2003JC002224/fullAbstract Top of page Abstract 1.Introduction 2.3DVAR-Based Data Assimilation Schemes 3.Assimilation Experiment Design 4.Experiment Results 5.Sensitivities of Estimation to T - S Diagrams 6.Weak Stratification Experiments 7.Conclusions Acknowledgments References Supporting Information [1] A data assimilation scheme based on three-dimensional variational analysis (3DVAR) is proposed to estimate temperature and salinity profiles from surface dynamic height information. The scheme takes into consideration vertical correlations for both temperature and salinity background errors and the nonlinear temperature-salinity ( T-S ) relation. In this study we designed some one-dimensional test cases to examine the separate and combined impacts of the vertical correlations and the nonlinear T-S relation on estimations of temperature and salinity profiles in comparison with a simplified scheme that considers neither vertical correlations nor T-S relations. Results show that the simplified scheme cannot simultaneously improve temperature and salinity profiles over their backgrounds in some cases and could make the correction seriously nonsmooth at different depths. The consideration of vertical correlations helps to balance the magnitude of the profile correction among all depths and produce smoother results. However, consideration of vertical correlations cannot help much in reducing the root-mean square error of estimation. The consideration of the nonlinear T-S relation can improve both temperature and salinity estimations in all test cases and can significantly reduce the root-mean square error of estimations. The combined effects of both vertical correlations and the nonlinear T-S relation are similar to those of the latter but with vertically smoother results.
    Yu J. -Y., H. -Y. Kao, 2007: Decadal changes of ENSO persistence barrier in SST and ocean heat content indices: 1958-2001. J. Geophys. Res., 112,D13106, doi: 10.1029/2006JD 007654.10.1029/2006JD0076548d9d5fb718997fe21c3e9ef079b14506http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2006JD007654%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/2006JD007654/pdfABSTRACT Decadal changes of El Ni&ntilde;o-Southern Oscillation (ENSO) persistence barriers in various indices of sea surface temperature (SST) and ocean heat content (OHC) are examined in this study using observations and ocean data assimilation products for the period 1958-2001. It is found that the SST indices in the eastern and central equatorial Pacific exhibit very different decadal barrier variability. The variability is large for the eastern Pacific SST indices (NINO1+2 and NINO3) whose persistence barriers shifted abruptly in 1976/1977 and 1989/1990. In contrast, the central Pacific SST indices (NINO3.4 and NINO4) experienced little decadal barrier variability and have had their persistence barriers fixed in spring in the past four decades. The zonal mean OHC index averaged over the equatorial Pacific shows decadal barrier changes similar to those in the eastern Pacific SST indices and always leads the NINO3 SST barrier by about one season. It is noticed that the SST persistence barrier appeared first in the eastern Pacific before 1976/1977, first in the central Pacific between 1976/1977 and 1989/1990, and almost simultaneous in both the eastern and central Pacific after 1989/1990. These timings coincide with the westward propagating, eastward propagating, and standing pattern of ENSO SST anomalies observed in these three periods. These results suggest that ENSO SST anomalies in the equatorial Pacific can be considered to consist of two different processes: a central Pacific process whose phase transition (such as onset) and barrier always happen in spring, and an eastern Pacific process whose phase transition and barrier change from decade to decade and are influenced by changes in the mean thermocline depth along the equatorial Pacific.
    Zhang Q., Y. -H. Ding, 2001: Decadal climate change and ENSO cycle. Acta Meteorologica Sinica, 59, 157- 172. (in Chinese)10.11676/qxxb2001.01740765e6e-a023-4c39-b925-ee9a40eea45355842001266488f0b078b9de6511b30e90f7cfc9c7http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-QXXB200102002.htmrefpaperuri:(d35e05c11c6b46e0408500063e081dd5)http://en.cnki.com.cn/Article_en/CJFDTOTAL-QXXB200102002.htmThe observed strongest global climate variation on the inter annual timescale is the El Niño-Southern Oscillation(ENSO) phenomenon, which is an irregular interannual oscillation resulting primarily from coupled ocean-atmosphere interaction in the tropical Pacific. Since 1970s the El Niño events have been stronger than the La Nina both on frequency and amplitude. Perhaps it links with the global climate warming. In this paper, sea surface temperature, wind stress and OLR data sets are diagnosed by using wavelet trans form. The ENSO cycle(2-7 years) signal and decadal variability(8-20 years) are filtered out to form the data sets in order to investigate the features of the decadal climate change and its effect on ENSO cycle. Then the numerical experiment sare designed to identify the effect byrunning a hybrid tropical Pacific coupled model. The main results of the paper can be outlined as follows:1) For the behavior of SSTA and wind stress in tropical Pacific by wavelet analysis, it is clear that not only the energy density of ENSO cycle(2-7 years) gets stronger during 1990s, but also does so the decadal oscillation(8-20 years). The long-term linear trend of climate change(more than 20 years) can beident ified with a catastrophe climate change in 1976 from cold climate state turning to warm state. 2) The main warming regions of the sea surfacet emperature after 1976 is the east tropical areas and extended west ward along the equator with the maximum value up to 0.6℃. The main features of wind stress pattern are the increasing westerly anomaly in west tropical Pacific and easterly anomaly at east tropical Pacific so that low level convection occurs near the date line. 3) However, the average state in the tropical Pacific in the 1990s is warmer than that in 1980s. This climate backgro und change in the 1990s may cause the asy mmetry of the warm and cold phases. On the ot her hand, the decadal oscillation has directly affected the ENSO cycle on both frequency and amplitude.
    Zhang Q., Y. Guan, and H. -J. Yang, 2008: ENSO amplitude change in observation and coupled models. Adv. Atmos. Sci.,25, 361-366, doi: 10.1007/s00376-008-0361-5.10.1007/s00376-008-0361-551bfea3116745b932382be0b3365ad23http%3A%2F%2Fd.wanfangdata.com.cn%2FPeriodical_dqkxjz-e200803003.aspxhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e200803003.aspxObservations show that the tropical El Nino-Southern Oscillation (ENSO) variability, after removing both the long term trend and decadal change of the background climate, has been enhanced by as much as 60% during the past 50 years. This shift in ENSO amplitude can be related to mean state changes in global climate. Past global warming has caused a weakening of the Walker circulation over the equatorial Indo-Pacific oceans, as well as a weakening of the trade winds and a reduction in the equatorial upwelling. These changes in tropical climatology play as stabilizing factors of the tropical coupling system. However, the shallower and strengthening thermocline in the equatorial Pacific increases the SST sensitivity to thermocline and wind stress variabilities and tend to destabilize the tropical coupling system. Observations suggest that the destabilizing factors, such as the strengthening thermocline, may have overwhelmed the stabilizing effects of the atmosphere, and played a deterministic role in the enhanced ENSO variability, at least during the past half century. This is different from the recent assessment of IPCC-AR4 coupled models.
    Zhang X. B., J. A. Church, 2012: Sea level trends, interannual and decadal variability in the Pacific Ocean. Geophys. Res. Lett., 39,L21701, doi: 10.1029/2012GL053240.10.1029/2012GL0532400d6c2d6c3c27267a772e3e0867772563http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL053240%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2012GL053240/citedby[1] Linear trend analysis is commonly applied to quantify sea level change, often over short periods because of limited data availability. However, the linear trend computed over short periods is complicated by large-scale climate variability which can affect regional sea level on interannual to inter-decadal time scales. As a result, the meaning of a local linear sea level trend over the short altimeter era (since 1993; less than 20 years) is unclear, and it is not straightforward to distinguish the regional sea level changes associated with climate change from those associated with natural climate variability. In this study, we use continuous near-global altimeter measurements since 1993 to attempt to separate interannual and decadal sea level variability in the Pacific from the sea level trend. We conclude that the rapid rates of sea level rise in the western tropical Pacific found from a single variable linear regression analysis are partially due to basin-scale decadal climate variability. The negligible sea level rise, or even falling sea level, in the eastern tropical Pacific and US west coast is a result of the combination of decreasing of sea level associated with decadal climate variability and a positive sea level trend. The single variable linear regression analysis only accounts for slightly more than 20% of the observed variance, whereas a multiple variable linear regression including filtered indices of the El Nino-Southern Oscillation and the Pacific Decadal Oscillation accounts for almost 60% of the observed variance.
    Zheng F., J. Zhu, 2015: Roles of initial ocean surface and subsurface states on successfully predicting 2006-2007 El Ni帽o with an intermediate coupled model. Ocean Science,11, 187-194, doi: 10.5194/os-11-187-2015.
    Zheng F., J. Zhu, H. Wang, and R. -H. Zhang, 2009: Ensemble hindcasts of ENSO events over the past 120 years using a large number of ensembles. Adv. Atmos. Sci.,26(2), 359-372, doi: 10.1007/s00376-009-0359-7.10.1007/s00376-009-0359-7.46b4cb1872e3da23cf3785e06025ff11http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00376-009-0359-7http://d.wanfangdata.com.cn/Periodical_dqkxjz-e200902020.aspxBased on an intermediate coupled model (ICM), a probabilistic ensemble prediction system (EPS) has been developed. The ensemble Kalman filter (EnKF) data assimilation approach is used for generating the initial ensemble conditions, and a linear, first-order Markov-Chain SST anomaly error model is embedded into the EPS to provide model-error perturbations. In this study, we perform ENSO retrospective forecasts over the 120 year period 1886-2005 using the EPS with 100 ensemble members and with initial conditions obtained by only assimilating historic SST anomaly observations. By examining the retrospective ensemble forecasts and available observations, the verification results show that the skill of the ensemble mean of the EPS is greater than that of a single deterministic forecast using the same ICM, with a distinct improvement of both the correlation and root mean square (RMS) error between the ensemble-mean hindcast and the deterministic scheme over the 12-month prediction period. The RMS error of the ensemble mean is almost 0.2C smaller than that of the deterministic forecast at a lead time of 12 months. The probabilistic skill of the EPS is also high with the predicted ensemble following the SST observations well, and the areas under the relative operating characteristic (ROC) curves for three different ENSO states (warm events, cold events, and neutral events) are all above 0.55 out to 12 months lead time. However, both deterministic and probabilistic prediction skills of the EPS show an interdecadal variation. For the deterministic skill, there is high skill in the late 19th century and in the middle-late 20th century (which includes some artificial skill due to the model training period), and low skill during the period from 1906 to 1961. For probabilistic skill, for the three different ENSO states, there is still a similar interdecadal variation of ENSO probabilistic predictability during the period 1886鈥2005. There is high skill in the late 19th century from 1886 to 1905, and a decline to a minimum of skill around 1910鈥50s, beyond which skill rebounds and increases with time until the 2000s.
    Zhou G. Q., W. W. Fu, J. Zhu, and H. J. Wang, 2004: The impact of location-dependent correlation scales in ocean data assimilation. Geophys. Res. Lett., 31,L21306, doi: 10.1029/2004 GL020579.10.1029/2004GL0205797a57ce7fbef2f5221c0e107f27357f4chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2004GL020579%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2004GL020579/citedbyAn approach is proposed to estimate the location-dependent correlation scales for background error covariance matrix used in the 3D-Var data assimilation, and the impact on ocean assimilation is examined. By a nonlinear fitting procedure, the horizontal scales are computed at each model grid point using the outputs of a tropical Pacific OGCM. The derived correlation scales are comparable to those obtained from observations, but present more complicated structures quite different from empirical functions. They vary largely in the horizontal and show a distinct distribution from surface to subsurface. Three assimilation experiments from 1982 to 2000 are performed to assess the improvement due to the location-dependent correlation scales. The overall root-mean-square errors in the upper ocean are reduced about 0.5C in the eastern equatorial Pacific and about 0.4掳C in the western equatorial Pacific. The improvement shows the location-dependent correlation scales can afford more information of the background error structures.
    Zhuang W., B. Qiu, and Y. Du, 2013: Low-frequency western Pacific Ocean sea level and circulation changes due to the connectivity of the Philippine Archipelago. J. Geophys. Res., 118, 6759- 6773.10.1002/2013JC009376bff2bc6efbe2e3d9ec1252d0a1abd804http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F2013JC009376%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/2013JC009376/pdfABSTRACT Interannual-to-decadal sea level and circulation changes associated with the oceanic connectivity around the Philippine Archipelago are studied using satellite altimeter sea surface height (SSH) data and a reduced gravity ocean model. SSHs in the tropical North Pacific, the Sulu Sea and the eastern South China Sea (ESCS) display very similar low-frequency oscillations that are highly correlated with El Ni&ntilde;o and Southern Oscillation. Model experiments reveal that these variations are mainly forced by the low-frequency winds over the North Pacific tropical gyre and affected little by the winds over the marginal seas and the North Pacific subtropical gyre. The wind-driven baroclinic Rossby waves impinge on the eastern Philippine coast and excite coastal Kelvin waves, conveying the SSH signals through the Sibutu Passage - Mindoro Strait pathway into the Sulu Sea and the ESCS. Closures of the Luzon Strait, Karimata Strait and ITF passages have little impacts on the low-frequency sea level changes in the Sulu Sea and the ESCS. The oceanic pathway west of the Philippine Archipelago modulates the western boundary current system in the tropical North Pacific. Opening of this pathway weakens the time-varying amplitudes of the North Equatorial Current bifurcation latitude and Kuroshio transport. Changes of the amplitudes can be explained by the conceptual framework of island rule that allows for baroclinic adjustment. Although it fails to capture the interannual changes in the strongly nonlinear Mindanao Current, the time-dependent island rule is nevertheless helpful in clarifying the role of the archipelago in regulating its multi-decadal variations.
  • [1] Soon-Il AN, 2018: Impact of Pacific Decadal Oscillation on Frequency Asymmetry of El Niño and La Niña Events, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 493-494.  doi: 10.1007/s00376-018-8024-7
    [2] Yuanpu LI, Wenshou TIAN, 2017: Different Impact of Central Pacific and Eastern Pacific El Niño on the Duration of Sudden Stratospheric Warming, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 771-782.  doi: 10.1007/s00376-017-6286-0
    [3] Wansuo DUAN, Chaoming HUANG, Hui XU, 2017: Nonlinearity Modulating Intensities and Spatial Structures of Central Pacific and Eastern Pacific El Niño Events, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 737-756.  doi: 10.1007/s00376-017-6148-9
    [4] Bo PANG, Zesheng CHEN, Zhiping WEN, Riyu LU, 2016: Impacts of Two Types of El Niño on the MJO during Boreal Winter, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 979-986.  doi: 10.1007/s00376-016-5272-2
    [5] N. FREYCHET, S. SPARROW, S.F. B. TETT, M.J. MINETER, G.C. HEGERL, D.C. H. WALLOM, 2018: Impacts of Anthropogenic Forcings and El Niño on Chinese Extreme Temperatures, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 994-1002.  doi: 10.1007/s00376-018-7258-8
    [6] Bin WANG, Juan LI, Qiong HE, 2017: Variable and Robust East Asian Monsoon Rainfall Response to El Niño over the Past 60 Years (1957-2016), ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1235-1248.  doi: 10.1007/s00376-017-7016-3
    [7] Feng XUE, Xiao DONG, Fangxing FAN, 2018: Anomalous Western Pacific Subtropical High during El Niño Developing Summer in Comparison with Decaying Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 360-367.  doi: 10.1007/s00376-017-7046-x
    [8] LI Jianying, LIU Boqi, LI Jiandong, MAO Jiangyu, 2015: A Comparative Study on the Dominant Factors Responsible for the Weaker-than-expected El Niño Event in 2014, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1381-1390.  doi: 10.1007/s00376-015-4269-6
    [9] Ben TIAN, Wansuo DUAN, 2016: Comparison of Constant and Time-variant Optimal Forcing Approaches in El Niño Simulations by Using the Zebiak-Cane Model, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 685-694.  doi: 10.1007/s00376-015-5174-8
    [10] Chaofan LI, Wei CHEN, Xiaowei HONG, Riyu LU, 2017: Why Was the Strengthening of Rainfall in Summer over the Yangtze River Valley in 2016 Less Pronounced than that in 1998 under Similar Preceding El Niño Events?——Role of Midlatitude Circulation in August, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1290-1300.  doi: 10.1007/s00376-017-7003-8
    [11] WangHuijun, Xue Feng, Bi Xunqiang, 1997: The Interannual Variability and Predictability in a Global Climate Model, ADVANCES IN ATMOSPHERIC SCIENCES, 14, 554-562.  doi: 10.1007/s00376-997-0073-2
    [12] YUAN Yuan, C. L. Johnny CHAN, ZHOU Wen, LI Chongyin, 2008: Decadal and Interannual Variability of the Indian Ocean Dipole, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 856-866.  doi: 10.1007/s00376-008-0856-0
    [13] FU Weiwei, 2012: Altimetric Data Assimilation by EnOI and 3DVAR in a Tropical Pacific Model: Impact on the Simulation of Variability, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 823-837.  doi: 10.1007/s00376-011-1022-7
    [14] Lin CHEN, Tim LI, Swadhin K. BEHERA, Takeshi DOI, 2016: Distinctive Precursory Air-Sea Signals between Regular and Super El Niños, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 996-1004.  doi: 10.1007/s00376-016-5250-8
    [15] Yu Yongqiang, Guo Yufu, 1995: The Interannual Variability of Climate in a Coupled Ocean-Atmosphere Model, ADVANCES IN ATMOSPHERIC SCIENCES, 12, 273-288.  doi: 10.1007/BF02656977
    [16] HU Ruijin, LIU Qinyu, WANG Qi, J. Stuart GODFREY, MENG Xiangfeng, 2005: The Shallow Meridional Overturning Circulation in the Northern Indian Ocean and Its Interannual Variability, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 220-229.  doi: 10.1007/BF02918511
    [17] Jian RAO, Rongcai REN, 2017: Parallel Comparison of the 1982/83, 1997/98 and 2015/16 Super El Niños and Their Effects on the Extratropical Stratosphere, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1121-1133.  doi: 10.1007/s00376-017-6260-x
    [18] QIAN Weihong, HU Haoran, 2006: Interannual Thermocline Signals and El Ni?no-La Ni?na Turnabout in the Tropical Pacific Ocean, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 1003-1019.  doi: 10.1007/s00376-006-1003-4
    [19] Kexin LI, Fei ZHENG, Jiang ZHU, Qing-Cun ZENG, 2024: El Niño and the AMO Sparked the Astonishingly Large Margin of Warming in the Global Mean Surface Temperature in 2023, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-3371-4
    [20] Jianjun Xu, Johnny C. L. Chan, 2002: Interannual and Interdecadal Variability of Winter Precipitation over China in Relation to Global Sea Level Pressure Anomalies, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 914-926.  doi: 10.1007/s00376-002-0055-3

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 28 April 2015
Manuscript revised: 26 July 2015
Manuscript accepted: 31 July 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

Evaluation of the Tropical Variability from the Beijing Climate Center's Real-Time Operational Global Ocean Data Assimilation System

  • 1. Laboratory for Climate Studies, National Climate Center, China Meteorological Administration, Beijing 100081
  • 2. State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou 510301
  • 3. Ministry of Education Key Laboratory for Earth System Modeling, and Center for Earth System Science, Tsinghua University, Beijing 100084
  • 4. International Center for Climate and Environment Science, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029

Abstract: The second-generation Global Ocean Data Assimilation System of the Beijing Climate Center (BCC_GODAS2.0) has been run daily in a pre-operational mode. It spans the period 1990 to the present day. The goal of this paper is to introduce the main components and to evaluate BCC_GODAS2.0 for the user community. BCC_GODAS2.0 consists of an observational data preprocess, ocean data quality control system, a three-dimensional variational (3DVAR) data assimilation, and global ocean circulation model [Modular Ocean Model 4 (MOM4)]. MOM4 is driven by six-hourly fluxes from the National Centers for Environmental Prediction. Satellite altimetry data, SST, and in-situ temperature and salinity data are assimilated in real time. The monthly results from the BCC_GODAS2.0 reanalysis are compared and assessed with observations for 1990-2011. The climatology of the mixed layer depth of BCC_GODAS2.0 is generally in agreement with that of World Ocean Atlas 2001. The modeled sea level variations in the tropical Pacific are consistent with observations from satellite altimetry on interannual to decadal time scales. Performances in predicting variations in the SST using BCC_GODAS2.0 are evaluated. The standard deviation of the SST in BCC_GODAS2.0 agrees well with observations in the tropical Pacific. BCC_GODAS2.0 is able to capture the main features of El Niño Modoki I and Modoki II, which have different impacts on rainfall in southern China. In addition, the relationships between the Indian Ocean and the two types of El Niño Modoki are also reproduced.

1. Introduction
  • Ocean initialization plays a critical role in short-term climate prediction because most predictive skill comes from the initial conditions of the upper ocean, particularly that associated with large-scale variability, such as ENSO and the Indian Ocean Dipole (IOD) (Alves et al., 2011). A common strategy to obtain the optimal initialization of the ocean is to assimilate the available ocean observations into ocean models forced by atmospheric fluxes (Xue et al., 2011; Zheng and Zhu, 2015). Various assimilation methods have been applied for the initialization of operational or quasi-operational coupled forecast systems. For example, a three-dimensional variational (3DVAR) analysis with a temporally and spatially varying background error covariance is used in the global ocean data assimilation system (GODAS) of the National Centers for Environmental Prediction (NCEP) (Behringer et al., 1998). In version 3 of the ECMWF (European Centre for Medium-Range Weather Forecasts) ocean analysis system, the incremental 3DVAR method is applied (Balmaseda et al., 2013). A more advanced data assimilation method, the EnKF (Ensemble Kalman Filter), has been employed in TOPAZ4 (the latest version of TOPAZ-a coupled ocean-sea ice data assimilation system for the North Atlantic Ocean and Arctic), which can make background error covariance vary with time (Sakov et al., 2012). Noticeably, the data assimilation in these systems is performed separately for ocean and atmosphere models in order to achieve optimal initialization (Zheng et al., 2009; Alves et al., 2011).

    High-quality in-situ and remote sensing data with fine spatial and temporal resolution are crucial for the development of a global operational ocean data assimilation system (Han et al., 2013; Ratheesh et al., 2014). Satellite observations of the ocean, which provide global coverage and real-time measurements of sea level, SST, sea ice, waves, and winds, have become a primary data source for operational oceanography. The launch of TOPEX/Poseidon (T/P) has provided an accurate description of the large-scale sea level and ocean circulation for the first time since 1992. The altimetry products at higher resolution have been supplied following the launch of ERS-1/2, Janson-1 and ENVISAT (Hunt et al., 2007). SST is an important variable for operational oceanography and for assimilation into global ocean models because SST is strongly related to air-sea interactions and can be used to correct errors in forcing fields (e.g., wind and surface heat fluxes). Satellite-derived SST products are available in two types, according to the sensors equipped on satellites. Optical/infrared data, including data from the Advanced Very High Resolution Radiometer (AVHRR), are affected by clouds and volcanic aerosols in the atmosphere (Reynolds and Marsico, 1993). Microwave data can be obtained through non-precipitating clouds and are very beneficial in terms of geographical coverage (Ratheesh et al., 2014). However, satellite observations are mainly confined to surface information. In-situ observations can complement satellite observations by providing measurements below the ocean surface. This is especially true of the Array for Real-time Geostrophic Oceanography (ARGO) project, which has been established since 2000 and now provides important in-situ observations to validate satellite data and improve the initialization of ocean models. ARGO floats measure the temperature and salinity in the upper 2000 m of the ocean and bring unquestionable advances to ocean forecast models (Huang et al., 2008). The joint use of high resolution satellite observations and high accuracy in-situ observations fulfills the operational requirements for temporal and spatial coverage and near real-time access.

    A new generation of short-term climate forecast systems is being developed at the Beijing Climate Center (BCC_ CSM1.1) (Wu et al., 2013, 2014), and the access to initial oceanic conditions depends on the development of a global ocean data assimilation system. The first generation of the global ocean data assimilation system of the Beijing Climate Center (BCC_GODAS1.0) has been developed during the period from January 1996 to December 2000, has been operational since 2002 (Liu et al., 2005), and plays an important role in climate forecast services. With the development of satellite technology and the data assimilation method, some limitations of BCC_GODAS1.0 have gradually emerged, especially with regard to not assimilating satellite observations. The physical processes in the BCC_GODAS1.0 ocean model and the ocean component used in BCC_CSM1.1 are different. A new method for assimilating altimeter and SST data under one dynamic constraint based on 3DVAR was developed to solve the problem of multivariate assimilation

    (Wang et al., 2012a, 2012b). The new method has been implemented into BCC_GODAS2.0, which now has the ability to assimilate real-time observations and to run automatically. The physical processes and grid of the ocean model used in BCC_GODAS2.0 are the same as those used in BCC_ CSM1.1. Therefore, the ocean initialization field generated by BCC_GODAS2.0 can be used in BCC_CSM1.1 directly. The global ocean reanalysis products of BCC_GODAS2.0 are being applied in BCC_CSM1.1.

    This paper is organized as follows: Section 2 provides an overview of the configurations of the ocean model and the 3DVAR analysis scheme, as well as the ocean observations used in the assimilation. Section 3 discusses the model performance and interannual-to-decadal variability in the tropical Pacific. A summary and conclusions are given in section 4.

2. The operational ocean system
  • The ocean model used in BCC_GODAS2.0 is version 4 of the Modular Ocean Model (MOM4), developed at the Geophysical Fluid Dynamics Laboratory (Griffies et al., 2003). Its zonal resolution is 1°, while the resolution in the meridional direction is 1/3° within 10° of the equator, smoothly increasing to 1° in the poleward direction of 30°. There are 50 z-levels in the vertical direction, with a 10 m resolution from the surface to a depth of 225 m, gradually increasing to approximately 366 m in the abyssal zone. The maximum depth is approximately 5.5 km. The tripolar grid is adopted to avoid a singularity at the North Pole (Griffies et al., 2005). The two north poles of the curvilinear grid are situated over land areas in North America and Eurasia, respectively. The model uses identical physical parameterization schemes as described in (Griffies et al., 2005), including the isopycnal tracer mixing and diffusion scheme, the Laplace horizontal friction scheme, and the K-Profile Parameterization vertical mixing scheme. The model is forced with 6-h averaged 10-m wind and 2-m air temperature data from the NCEP/NCAR (National Center for Atmospheric Research) Reanalysis I dataset (http://www.esrl.noaa.gov/psd/). The mean climatological river runoff is specified at the coastlines of the model. Surface temperature and salinity are relaxed to monthly Levitus climatology, with restoring time scales of 90 and 120 days, respectively.

  • A 3DVAR analysis scheme, adapted from (Yan et al., 2004), has been improved by adopting a recursive filter to explicitly avoid the inverse calculation of the background error covariance matrix (Xiao et al., 2008) and save computing resources. The analysis variables are temperature and salinity. There are two steps carried out in the 3DVAR. First, the temperature and salinity profiles along the satellite altimeter track are estimated by minimizing the cost function, defined as follows: \begin{eqnarray*} J&=&({T}-{T}_{ b})^{ T}{B}_{ T}^{-1}({T}-{T}_{ b})+\\ &&[{S}-{g}(T)]^{ T}{B}_{ S}^{-1}[{S}-{g}(T)] +\dfrac{1}{2\sigma^2}[h({T},{S})-h_0]^2 , \end{eqnarray*} where T and S are the column vectors containing the state variables of temperature and salinity, respectively; T b is the temperature background vector; and B T and B S are the background error covariance matrices in the vertical direction for temperature and salinity, respectively. B T and B S are both assumed to be diagonal matrices, and the diagonal element is given by the empirical functions used by (Behringer et al., 1998) and (Yan et al., 2004). The variance of the background error in temperature at depth z is given by $$ a_{vT}\dfrac{(d{T}/dz)^{\frac{1}{2}}}{[(d{T}/dz)^{\frac{1}{2}}]_{{ max}}} , $$ where the constant avT is determined empirically by tuning the assimilation results and setting to 2.7; dT/dz is the local vertical temperature gradient at depth z; and \([(dT/dz)^\frac12]_ max\) is the maximum value of \((dT/dz)^\frac12\) in the water column. The determination of B S is analogous to B T, except the corresponding constant avS is set equal to 0.6. g(T) is a nonlinear T-S relationship function, which is determined from datasets of historical observations. h o is the observed value of the sea surface height. σ is observational error of the sea surface height, which is set to 6 cm here. The function h(T,S) denotes an observation operator that transforms T and S to the surface dynamic height, which is defined as follows: $$ h(T,S)=-\int_0^{z_m}\dfrac{\rho(T,{S},p)-\rho_0(p)}{\rho_0(p)}dz , $$ where zm is the reference depth, set to 1000 m here. z denotes the vertical coordinate and p denotes the pressure. The function ρ(T,S,p) denotes the sea water state equation for calculating density. ρ0(p)=ρ(0,35,p) is the reference density. More details can be found in (Fu et al., 2009a) and (Xiao et al., 2008).

    Second, the estimated temperature and salinity calculated from the first step are taken as synthetic observations, which are assimilated with the in-situ temperature and salinity observations, such as those from ARGO. The cost function is defined as follows: \begin{eqnarray*} J&\!=\!&({T}-{T}_{ b})^{ T}{E}_{ T}^{-1}({T}-{T}_{ b})+({S}-{S}_{ b})^{ T}{E}_{ S}^{-1}({S}-{S}_{ b})+\\ &&({HT}-{T}_0)^{ T}{O}_{ T}^{-1}({HT}\!-\!{T}_0)\!+\!({HS}\!-\!{S}_0)^{ T}{O}_{ S}^{-1}({HS}\!-\!{S}_0) , \end{eqnarray*} where T and S are the temperature and salinity vectors at the model levels, respectively; and O T and O S are the observation error covariance matrices of temperature and salinity, respectively. The observation error covariances here are not simple in-situ observation error but include the synthetic observation error and repetitiveness errors of the variability of unresolved smaller scales. Their orders of magnitude are larger than those of the observation errors in the first step. We set the observation errors of temperature and salinity equal to 0.8°C and 0.1 psu, respectively. H is the bilinear interpolation operator that interpolates the model grid points to the observation location. T b and E T are the background temperature and its corresponding background error covariance matrix in the horizontal direction, respectively. S b and E S are the background salinity and its corresponding error variance matrix, respectively. It should be noted that E T and E S are not diagonal matrices. The background error covariance matrix is defined using the Gaussian function as follows: $$ E=A\exp\left(-\dfrac{\Delta x^2}{L_x^2}-\dfrac{\Delta y^2}{L_y^2}\right) , $$ where A denotes the background error variance of temperature (set to 2.0°C2) or salinity (set to 0.15 psu2); ∆ x and ∆ y are the distances of two horizontal grid points in the zonal and meridional directions, respectively; and Lx and Ly are correlation scales in the zonal and meridional directions, respectively. We take Lx=450 km and Ly=650 km for the temperature, and Lx=420 km and Ly=510 km for the salinity, by tuning the analysis. The detail of the method used to set appropriate values for Lx and Ly can be found in (Zhou et al., 2004).

    Figure 1.  Vertical distribution of RMSE for the temperature and salinity during 1990-2011. The global average statistics are shown for the free run (blue line) and reanalysis (red line).

  • In-situ temperature and salinity, satellite-SST and altimeter-derived sea-level anomalies (SLAs) are used for assimilation. The SST observations are from the AVHRR merged product (ftp://eclipse.ncdc.noaa.gov/pub/OI-daily/) created through optimal interpolation (Reynolds et al., 2007). SLAs are from all available satellite altimeters, including T/P, Jason-1/2, ERS-1/2 and Envisat (http://www.aviso.altimetry.fr/en/home.html), which is 0.25°× 0.25° and a 7-day average gridded dataset. In-situ temperature and salinity are from the Global Temperature and Salinity Profile Project (GTSPP) and ARGO. Quality control is critically important because erroneous values can cause spurious overturning in ocean models. The temperature and salinity profiles used in the study are mainly subject to unified quality control, including a duplicate check (similar position/time check), a depth inversion check, a temperature and salinity valid range check, an excessive gradient check, and a spike value check. The temperature and salinity valid range check applies a gross filter on observed values, which needs to include all of the expected extremes encountered in the oceans (temperature in the range of -2.0°C to 39.0°C; salinity in the range of 0.0 to 41.0 psu). The excessive gradient check requires that the ratio of the difference in adjacent values to the difference in the depths is no larger -0.2°C m-1 for temperature and 5.0 psu m-1 for salinity. The spike value check requires the test value (α) to be no larger than some pre-defined value: $$ \alpha=|V_2-(V_3+V_1)/2|-[(V_3-V_1)/2] , $$ where V2 is the measurement being tested as a spike, and V1 and V3 are the values above and below the tested layer. If α exceeds 6.0 for pressures less than 5000 hPa, or α exceeds 2.0 for pressures less than 5000 hPa, the measurement V2 is removed for temperature and salinity. Further details can be found in the ARGO quality control manual, version 2.8(http://www.argodatamgt.org/content/download/15699/102401/file/argo-quality-control-manual-version2.8.pdf).

  • The operational BCC_GODAS2.0 has been running on a Sunway4000A high-performance computer system at the China Meteorological Administration since 1 April 2014. It runs on six nodes (36 processors) and takes two hours (wall clock) to process the observations and three hours to finish the analysis. The data assimilation cycle for the system restarts the model 10 days prior to the most recent time of the NCEP forcing field, while the observations are assimilated within this time window (10 days). The shell scripts controlling the operational run have been implemented to minimize the need for manual operation. A typical daily run starts with the creation of forcing fields for the ocean model. Then, the operational system starts to run after the real-time observations of the SLAs, SST, in-depth temperature and salinity are processed. If the forcing fields are not received by the prescribed time, the model will be forced by the climatological forcing fields.

3. Validation of the operational ocean system
  • In this study, we evaluate the monthly results from the BCC_GODAS2.0 reanalysis for 1990-2011.

  • The temperature and salinity observations from World Ocean Atlas 2001 (WOA01) are used to evaluate the agreement of BCC_GODAS2.0 with observations. The vertical profiles of the root-mean-square error (RMSE) of temperature and salinity are shown in Figs. 1a and 1b, respectively. Large temperature errors are present mainly in the thermocline, and the errors are largest at a depth of approximately 200 m in the free run. The RMSE of the temperature is reduced at most vertical levels by assimilation. The strong relaxation of the SST constrains the surface temperature variation in the model so that the difference between the results of the free run and the reanalysis results is not obvious in the top layer. The overall RMSE of the temperature is reduced by 0.81°C after the cumulative effect of many assimilation cycles during integration over 22 years. Like the temperature, the RMSE of salinity is reduced by 0.24 psu by assimilation. Note that the improvement can be attributed not only to the assimilation of temperature and salinity but also to the assimilation of the SLA through the dynamic constraints (Yan et al., 2004).

    The performance of BCC_GODAS2.0 is further evaluated in the tropical Pacific Ocean (30°S-30°N, 120°E-70°W), Indian Ocean (30°S-30°N, 30°-120°E) and Atlantic Ocean (30°S-30°N, 80°W-20°E), separately. Table 1 compares the RMSEs of the free run and the reanalysis. The reduction of RMSE (RRMSE, in %) of the reanalysis relative to the free run is shown as well. The RRMSE is defined in (Counillon et al., 2014) as $$ { RRMSE}_{ reanalysis}=\dfrac{{ RMSE}_{ free}-{ RMSE}_{ reanalysis}}{{ RMSE}_{ free}} , $$ where RMSE free and RMSE reanalysis are the RMSEs of the free run and the reanalysis, respectively. The RMSEs and RRMSEs of temperature (T) and salinity (S) are calculated for surface (0-225 m) and deep (225-1007 m) waters. The RRMSEs of T and S for water depths greater than 1007 m are not examined because the variations of T/S errors with depth are relatively small below 1000 m (as shown in Fig. 1). In general, the RRMSE for deep waters is larger than the surface due to less variability than at the surface. The maximum error reductions are achieved in the Atlantic Ocean due to the large errors in the free run. The RRMSEs are greater than 60% at the surface, and greater than 70% in deep waters for both T and S. In the Indian Ocean, the RRMSE is generally larger than 70% for S, while the RRMSE for T is more than 20%. The RRMSE for the Pacific is more than 50%, while for T in the surface layer it is about 33.2%. The relatively low error reduction of RRMSE for T at the surface is mainly caused by the relatively low RMSE in the free run. Overall, the error reductions are significant.

  • The mixed layer depth (MLD) is one of the most important quantities of the upper ocean, and its variability strongly influences the physics of the upper ocean and marine biological processes (e.g., de Boyer Montégut et al., 2004; D'Ortenzio et al., 2005; Dong et al., 2007, 2009; Ren et al., 2011; Sallée et al., 2008; Moore et al., 2013). We estimate the monthly mean MLD from the temperature and salinity based on the density criterion of 0.125 (kg m-3), in which the density is supposed to change from that of the ocean surface of 0.125 kg m-3 (Levitus, 1982). The monthly minimum and maximum MLD for BCC_GODAS2.0 and WOA01 are shown in Fig. 2. In general, the MLD is deep in winter and shallow in summer for both hemispheres. The minimum MLD in BCC_GODAS2.0 is less than 30 m in the midlatitudes. This is consistent with the WOA01 minimum MLD. In the central tropical Pacific, the MLD is 50-60 m, and in the eastern upwelling region the MLD is relatively shallow (<20 m). A comparison of the minimum MLD with WOA01 illustrates the relatively deep MLD in the Southern Ocean for BCC_GODAS2.0. The possible reasons for the differences include deficiencies in model physics and the temporally constant background error covariance in the 3DVAR. It is well known that the Southern Ocean has abundant eddies. The submesoscale eddy-driven restratification of the mixed layer has not been resolved in the model due to the relatively coarse resolution (1°× 1°) at high latitudes. This is the main contribution to the deeper MLD than those of observations. Another possible reason is that the temporally constant background error covariance in the 3DVAR cannot fully capture the variability of eddies in the Southern Ocean, and thus also cannot simulate eddies well.

    A comparison of the maximum MLD highlights the deep mixing in the North Atlantic, consistent with the robust North Atlantic deep water formation. The maximum MLDs in the North Pacific are relatively shallow (<100 m), also consistent with the lack of deep-water formation there. Homogenous cold and moderately low salinity water is found east of Japan, indicating the formation of the North Pacific Intermediate Water, consistent with (Talley, 1993). In the Southern Ocean, the observed maximum MLD at 50°S is approximately 720 m, while the model MLD is over 750 m. Therefore, BCC_GODAS2.0 is generally in good agreement with WOA01.

    Figure 2.  Climatology of the monthly minimum and maximum MLD based on the density criterion 0.125 kg m$^-3$ (Levitus, 1982) in BCC_GODAS2.0 and WOA01. Note that the color scales for the minimum MLD and maximum MLD are different. Panels (a) and (c) are for BCC_GODAS2.0 and (b) and (d) are for WOA01. (units: m)

  • Sea level variability is a good indicator of dynamic ocean processes. Some recent studies have noted that the sea level in the tropical Pacific displays significant interannual-to-decadal variations in response to climate variability (e.g., Fu et al., 2009b; Merrifield et al., 2012; Qiu and Chen, 2012; Zhuang et al., 2013). To evaluate the model's performance over the tropical Pacific, empirical orthogonal function (EOF) analysis is applied to low-frequency (period longer than 13 months) SLAs from both altimeter observations and BCC_GODAS2.0 simulations. The EOF patterns and the associated principal components (PCs) reflect the spatial and temporal variations, respectively. The first EOF mode of the observed SLA accounts for 54.9% of the total variation, and its EOF-1 shows oppositely signed SLAs in the eastern and western basin (Fig. 3a). The EOF-1 of modeled SLAs (Fig. 3b), which explains 45.1% of the total variation, generally captures the observed spatial pattern. The modeled PC-1 tracks the observations quite well (Fig. 3c; correlation coefficient = 0.74, significant at the 95% confidence level), and both show a decadal increasing trend over the past two decades with embedded interannual modulations. A close inspection of the PC-1s reveals that strong sea level oscillations existed in 1994-95, 1997-98, 2006-07 and 2009-10, indicative of the impacts from El Niño-Southern Oscillation (ENSO) events.

    Figure 4 further shows the ability of BCC_GODAS2.0 to reproduce sea level trends during 1993-2011. In both BCC_GODAS2.0 and observations, the linear sea level trends decrease zonally from west to east, consistent with the results of the first EOF modes in Fig. 3. In general, BCC_GODAS2.0 satisfactorily reproduces the sea level rises in the western Pacific, and falls in the eastern Pacific. The most significant rises occur at the latitudes of 5°-15°N in the northwestern Pacific. The zonally opposite sea level trends are mainly caused by the steady intensification of the trade winds across the tropical Pacific that started in the early 1990s (Merrifield and Maltrud, 2011; Nidheesh et al., 2013). Recent studies have linked the notable sea level trends, together with the easterly trade intensification, to decadal climate modes, especially the Pacific Decadal Oscillation and the decadal variability of ENSO (Merrifield et al., 2012; Zhang and Church, 2012). It is noteworthy that the modeled sea level trends in Fig. 4b have a basin-mean rate of 1.9 mm yr-1, approximately 25% lower than the mean rate of 2.4 mm yr-1 in satellite observations, which is probably because BCC_GODAS2.0 does not include the increase in ocean mass resulting from the contributions of ice sheets, melting glaciers and continental waters.

    Figure 3.  EOF-1 of low-frequency ($>$13 months) SLAs derived from (a) altimeter observations and (b) BCC_GODAS2.0 simulations for the period 1993-2011. (c) PC-1s of observed and modeled SLA.

    Figure 4.  Linear trends of (a) altimetric and (b) BCC_GODAS2.0 simulated sea levels during 1993-2011 \small \parbox[t]13cm(units: mm yr$^-1$).

    Figure 5.  The STD of DJF SST during 1990-2011. Panels (a-d) represent SSTs in ERSST, HadISST, the free run and BCC_GODAS2.0, respectively. (units: $^\circ C^2$).

  • It is well known that the SST in the tropical Pacific shows significant interannual variability, which greatly influences global climate. In this part, the simulated SST of BCC_GODAS2.0 in the tropical Pacific is evaluated. The monthly SSTs from the Hadley Centre Sea Ice and SST dataset (HadISST) on a 1°× 1° resolution (Rayner et al., 2003), and the monthly Extended Reconstruction of SST, version 3b (ERSST.v3b), on a 2°× 2° resolution (Smith et al., 2008), from 1990 to 2011, are used to validate the analyzed SST.

    Spatial distributions of the standard deviation (STD) of the SST in the equatorial Pacific during boreal winter (December-January-February, DJF) are shown in Fig. 5. Improvements in the STD of SST by assimilation are mainly located in the tropical Pacific and Atlantic (Figs. 5c and d). In the observations (Figs. 5a and b), the distributions of the STD are symmetric about the equator, and the maximum centers of the STD are approximately 140°W. In BCC_GODAS2.0, the spatial distributions of the SST STD closely resemble the observations (Fig. 5d). However, differences between the observed and simulated SST are found near the east coast of South America, while the SST STDs from BCC_GODAS2.0 are weaker than observed, which is probably because the upwelling-favored wind stress used in the model is weaker than in reality, and the model's grid resolution cannot resolve the local bathymetry near the coast of America.

    In recent years, two types of El Niño, the canonical El Niño and El Niño Modoki, have been suggested (i.e., Larkin and Harrison, 2005; Yu and Kao, 2007; Ashok et al., 2007; Kao and Yu, 2009; Kug et al., 2009; Wang et al., 2009). The latter is also referred to as the Date Line El Niño (Larkin and Harrison, 2005), central Pacific El Niño (Yu and Kao, 2007), and warm pool El Niño (Kug et al., 2009). The performances of canonical El Niño and Niño Modoki in BCC_GODAS2.0 are presented. Here, canonical El Niño is defined by the Niño3 index, which is the area-averaged SST anomalies (SSTA) in the region (5°S-5°N, 150°-90°W). El Niño Modoki is identified by the El Niño Modoki Index (EMI), defined by (Ashok et al., 2007) as $$ { EMI}={ [SSTA]}_{ C}-0.5\times { [SSTA]}_{ E}-0.5\times{ [SSTA]}_{ W} , $$ where the square brackets with a subscript represent the area-averaged SSTA over the central Pacific region (C) (10°S-10°N, 165°E-140°W), the eastern Pacific region (E) (15°S-5°N, 110°-70°W), and the western Pacific region (W) (10°S-20°N, 125°-145°E), respectively. Figure 6 shows that the changes in both the Niño3 index and the EMI in BCC_GODAS2.0 agree well with those from HadISST and ERSST. The correlation coefficients of the two El Niño indices between BCC_GODAS2.0 and the two observations exceed 0.9, which are higher than those of the free run (Table 2). The amplitudes of the two El Niño indices in BCC_GODAS2.0 show some differences from the observations. The intensities of the Niño3 index in BCC_GODAS2.0 are somewhat underestimated.

    We next explore the possible reasons why the intensities of the Niño3 index in BCC_GODAS2.0 are weaker than those from observations. Figure 7 shows the climatology of the annual SSTs during 1990-2011 in both the simulations and observations. The mean climate state of SSTs in the eastern Pacific in BCC_GODAS2.0 (Fig. 7d) is colder than those from observations (Figs. 7a and b). The warmer climate state of SSTs will strengthen the air-sea interaction and increase the amplitudes of the ENSO warm and cold phases (Zhang and Ding, 2001; Zhang et al., 2008). The difference in the climate state of SSTs between BCC_GODAS2.0 and observations could result in weaker intensities of the Niño3 index in BCC_GODAS2.0. As is well known, the two El Niño indices show significant interannual variability. The Niño3 index in the observations has interannual variability with periods of 5 years and 3 years, while the periods of Niño3 index are 4 years in the simulations. The EMI shows the same interannual variability, with periods of 2.5 years in both the observations and simulations (Fig. 8).

    Figure 6.  Time series of DJF-mean (a) Niño3 index and (b) EMI during 1990-2011 with ERSST (green line), HadISST (blue line), free run data (black line), and BCC_GODAS2.0 (red line) (units: °C).

    Figure 7.  Climatology of annual SSTs during 1990-2011. The black box indicates the region used to calculate Niño3 index. Panels (a-d) represent SSTs in ERSST, HadISST, the free run and BCC_GODAS2.0, respectively (units: °C).

    Figure 8.  Power spectrum of the monthly (a-d) Niño3 index and (e-h) EMI during 1990-2011. The black curves and red curves represent the calculated spectrum and 90% confidence interval, respectively. The first, second, third and fourth rows represents ERSST, HadISST, the free run and BCC_GODAS2.0, respectively.

    Figure 9.  Spatial distributions of composited SSTAs for (a-c) El Niño Modoki I and (d-f) El Niño Modoki II in the boreal autumn (September-November). The first, second, and third row represent ERSST, HadISST and BCC_GODAS2.0, respectively. The composites are calculated from three El Niño Modoki I events (1990/91, 1991/92, and 2002/03), and three El Niño Modoki II events (1992/93, 2004/05 and 2009/10) (units: °C).

    According to their different impacts on rainfall in southern China and typhoon landfall activity, (Wang and Wang, 2013) separated El Niño Modoki into two subtypes, El Niño Modoki I and II, in which the warm SST anomalies originate from the central tropical Pacific and subtropical northeastern Pacific, respectively. In this study, the simulations of El Niño Modoki I and II in BCC_GODAS2.0 are evaluated. Using the classification of (Wang and Wang, 2013), there are three El Niño Modoki I events (1990/91, 1991/92, and 2002/03) and three El Niño Modoki II events (1992/93, 2004/05 and 2009/10). Figure 9 shows the composite SSTAs for El Niño Modoki I and El Niño Modoki II during the developing autumn (September-October-November) from the three datasets of ERSST, HadISST, and BCC_GODAS2.0. Similar to those in ERSST and HadISST, the warm SSTAs of El Niño Modoki I in BCC_GODAS2.0 are in the central tropical Pacific and are symmetric about the equator. Compared with those in ERSST and HadISST, the warm SSTAs of El Niño Modoki I in BCC_GODAS2.0 are weaker, and their area is smaller. The cool SSTAs are observed in the eastern tropical Pacific (Figs. 9a-c). For El Niño Modoki II, although the warm SST anomalies of BCC_GODAS2.0 show an asymmetric distribution extending and tilting from the subtropical northeastern Pacific to the central equatorial Pacific, as do the observations, their amplitudes are weaker than those in ERSST and HadISST (Figs. 9d-f). In addition, the relationships of the Indian Ocean with the two types of El Niño Modoki are reproduced. (Wang and Wang, 2014) suggested that El Niño Modoki I and II are associated with a positive and negative IOD in boreal autumn, respectively. Here, a positive IOD means warm/cool SSTAs in the western/eastern tropical Indian Ocean (Saji et al., 1999). In BCC_GODAS2.0, the positive and negative IODs are associated with El Niño Modoki I and II, respectively, which is consistent with observations. The amplitude of the positive IOD during El Niño Modoki I in BCC_GODAS2.0 is weaker than observed.

4. Summary and conclusions
  • The second-generation Global Ocean Data Assimilation System of the Beijing Climate Center (BCC_GODAS2.0) was implemented in April 2014. The ocean model of BCC_GODAS2.0 is MOM4, which adopts tripolar grids to avoid a singularity at the North Pole. An improved 3DVAR scheme is used in BCC_GODAS2.0 by adopting a recursive filter to speed up computation. Satellite SLA and SST data and in-situ observations, including ARGO and GTSPP, are assimilated in real time. BCC_GODAS2.0 has been run daily in pre-operational mode, driven by six-hourly fluxes from the NCEP. The MLD variation of BCC_GODAS2.0 is generally in agreement with that of WOA01. The simulated SST and SSH successfully depict the interannual variability of the tropical Pacific, which has a great impact on the climate of China. It is very important, for the climate monitoring services of the BCC, to evaluate the performance of the tropical Pacific variability in BCC_GODAS2.0. In the tropical Pacific, the first EOF and PC of low-frequency SLAs from BCC_GODAS2.0 are similar to those from satellite altimeter observations. BCC_GODAS2.0 can also satisfactorily reproduce the sea level rises in the western Pacific, and falls in the eastern Pacific. The STD of the SST in BCC_GODAS2.0 is similar to those of ERSST and HadISST in the tropical Pacific. The Niño3 index and EMI of BCC_GODAS2.0 show high correlation (>0.9) with those of ERSST and HadISST. A power spectrum analysis of the monthly averaged SST shows that the Niño3 index and EMI of BCC_GODAS2.0 have the most significant periods at approximately 5 and 2.5 years, respectively, which is consistent with ERSST and HadISST. BCC_GODAS2.0 can capture the main features of El Niño Modoki I and Modoki II. The performance of the IOD in the BCC_GODAS2.0 SST is similar to that observed in the Indian Ocean and generally agrees with that shown in (Wang and Wang, 2013). The global ocean reanalysis products and real-time assimilation results are being applied in the new generation of the short-term forecast system, BCC_CSM1.1.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return