Advanced Search
Article Contents

LICOM Model Datasets for the CMIP6 Ocean Model Intercomparison Project

Fund Project:

National Key R&D Program for Developing Basic Sciences (Grant Nos. 2016YFC1401401 and 2016YFC1401601), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDC01000000), and the National Natural Science Foundation of China (Grants Nos. 41576026, 41576025, 41776030, 41931183 and 41976026)


doi: 10.1007/s00376-019-9208-5

  • The datasets of two Ocean Model Intercomparison Project (OMIP) simulation experiments from the LASG/IAP Climate Ocean Model, version 3 (LICOM3), forced by two different sets of atmospheric surface data, are described in this paper. The experiment forced by CORE-II (Co-ordinated Ocean–Ice Reference Experiments, Phase II) data (1948–2009) is called OMIP1, and that forced by JRA55-do (surface dataset for driving ocean–sea-ice models based on Japanese 55-year atmospheric reanalysis) data (1958–2018) is called OMIP2. First, the improvement of LICOM from CMIP5 to CMIP6 and the configurations of the two experiments are described. Second, the basic performances of the two experiments are validated using the climatological-mean and interannual time scales from observation. We find that the mean states, interannual variabilities, and long-term linear trends can be reproduced well by the two experiments. The differences between the two datasets are also discussed. Finally, the usage of these data is described. These datasets are helpful toward understanding the origin system bias of the fully coupled model.
    摘要: 本文描述了大气所大气科学和地球流体力学数值模拟国家重点实验室(LASG/IAP) 气候系统海洋模式第三版(LICOM3)参与海洋模式比较计划的两个模拟试验数据集。两个模拟试验采用不同大气表面强迫数据,采用规范的海洋-海冰参考试验第二阶段 (Co-ordinated Ocean–Ice Reference Experiments, Phase II, CORE-II) 作为外强迫的试验称为海洋模式比较计划1阶段(OMIP1,1948-2009),采用来自日本55年大气再分析表面数据驱动海洋-海冰模式 (JRA55-do)强迫试验称为海洋模式比较计划第二阶段(OMIP2,1958-2018)。首先,文章描述了LICOM从CMIP5到CMIP6的改进和试验基本设置。其次,利用来自观测气侯态和年际变化数据,对两个试验的模拟性能进行检验。结果表明模式很好再现了两个试验中的平均态、年际变化和长期趋势。同时,讨论了两个模拟试验数据间的差异。最后,介绍了数据集的使用情况。这两个数据集有助于了解全耦合模式的系统偏差来源。
  • 加载中
  • Figure 1.  Annual global mean (a) SST (units: °C), (b) volume ocean temperature (VOT; units: °C), (c) SSS (units: psu), and (d) volume ocean salinity (VOS; units: psu) for OMIP1 (cyan) and OMIP2 (purple) during all the six cycles. The black lines in the figure indicate the reference value calculated from WOA13 observation.

    Figure 2.  (a) Annual global mean SST (units: °C), (b) annual mean SIC in the NH and SH, (c) steric sea level, and (d) AMOC, for OMIP1 (cyan), OMIP2 (purple) and ERSST.v5/NSIDC/Argo/RAPID (black) during the sixth cycles. The STDs (removing linear trend), linear trends and correlation coefficients (SST during 1958–2009, SIC for 1980–2009, 2005–2018 for Argo, 2004–16 for RAPID) are provided in the figures.

    Figure 3.  Simulations (contours) and biases (shaded) of SST (units: °C) for (a) OMIP1 and (b) OMIP2, and (c) the difference betweeen OMIP2 and OMIP1 (OMIP2 minus OMIP1). (d–f) As in (a–c), respectively, but for SSS (units: psu). The annual SST and SSS distributions employ the data from 1980–2009 of the sixth cycle. The RMSEs, mean values, and minimim and maximal values are provided in the figures.

    Figure 4.  The (a) observed and (b, c) OMIP1- and OMIP2-simulated SSH (units: m) during 1993–2009 of the sixth cycle. The spatial correlations between the observation and simulation are noted in the top right.

    Figure 5.  The simulated AMOC (units: Sv; 1Sv = 1 × 106 m3 s−1) for (a) OMIP1 and (b) OMIP2 during the last 30 years (1980–2009) of the sixth cycle. (c) AMOC values at the latitude of 26.5°N for OMIP1, OMIP2 and RAPID (2005–09). The maximal values and STD (+/-) between 2005–09 are noted. The y-axises represent ocean depth (units: km).

    Figure 6.  The STDs (units °C) of observed SST anomalies from (a) ERSST.v5 and (b, c) OMIP1- and OMIP2-simulated SST anomalies during 1960–2009. Before calculating the STD, both the linear trend and annual cycle were removed from the SST. The spatial correlation coefficients (R) with the observation are denoted in the top middle. The STDs in the Niño3 region (5°S–5°N, 150°–90°W) are denoted in the top right.

    Table 1.  Comparison of model configurations between two versions of LICOM (LICOM2.0 and LICOM3).

    ConfigurationLICOM2.0LICOM3
    GridGridLongitude/LatitudeTripolar
    Resolution~1°, 30 levels~1°, 30 or 80 levels
    Dynamic coreTracer advectionCentral differential schemePreserved shape scheme (Yu, 1994)
    Momentum time integrationExplicitImplicit
    PhysicsDiapycnal mixingMixing in the mixed layer (Canuto et al., 2001, 2002)Mixing in the mixed layer (Canuto et al., 2001) and internal tide mixing (St. Laurent et al., 2002)
    Isopycnal mixingIsopycnal mixing (Redi, 1982) and advection (Gent and McWilliams, 1990)Isopycnal mixing (Redi, 1982) and advection (Gent and McWilliams, 1990) with N2 thickness diffusivity (Ferreira et al., 2005)
    Computing technicsCoupler interfaceNCAR Flux Coupler 6NCAR Flux Coupler 7
    Parallel1D MPI and OMP2D MPI and OMP
    DataInitial conditionWOA01 (Conkright et al., 2002)PHC3.0 (Steele et al., 2001)
    BathymetryDBDB5 (https://www.bodc.ac.uk/resources/inventories/edmed/report/356/)ETOPO2 (https://ngdc.noaa.gov/mgg/global/etopo2.html)
    DownLoad: CSV

    Table 2.  Depths of model levels for the T- and W-grid. Positive is upward.

    LevelDepth for TDepths for W
    1−50
    2−15−10
    3−25−20
    4−35−30
    5−45−40
    6−55−50
    7−65−60
    8−75−70
    9−85−80
    10−95−90
    11−105−100
    12−115−110
    13−125−120
    14−135−130
    15−145−140
    16−156.9303−150
    17−178.4277−163.8606
    18−222.5018−192.9948
    19−303.1057−252.0088
    20−432.5961−354.2027
    21−621.1931−510.9896
    22−876.5334−731.3966
    23−1203.337−1021.67
    24−1603.2−1385.003
    25−2074.526−1821.396
    26−2612.596−2327.656
    27−3209.772−2897.536
    28−3855.835−3522.009
    29−4538.428−4189.662
    30−5243.597−4887.194
    31−5600
    DownLoad: CSV

    Table 3.  Descriptions of the experiments of LICOM3.

    Experiment_idModelInitial condition (TS/currents)Forcing dataPeriodFrequency
    OMIP1LICOM3/CICE4.0PHC3.0/ZeroCORE-II1948–2009 (6 cycles)6 h
    OMIP2LICOM3/CICE4.0PHC3.0/ZeroJRA55-do1958–2018 (6 cycles)6 h
    DownLoad: CSV

    Table 4.  Descriptions of dataset variables of Priority 1.

    NameDescriptionFrequency
    hfbasinNorthward ocean heat transportMonthly
    hfbasinpmadvNorthward ocean heat transport due to parameterized mesoscale advectionMonthly
    hfbasinpmdiffNorthward ocean heat transport due to parameterized mesoscale diffusionMonthly
    hfdsDownward heat flux at sea water surfaceMonthly
    masscelloOcean grid-cell mass per areaMonthly
    mlotstOcean mixed-layer thickness defined by sigma TMonthly
    msftbarotOcean barotropic mass streamfunctionMonthly
    msftmzOcean meridional overturning mass streamfunctionMonthly
    msftmzmpaOcean meridional overturning mass streamfunction due to parameterized mesoscale advectionMonthly
    obvfsqSquare of Brunt–Vaisala frequency in sea waterMonthly
    pboSea water pressure at sea floorMonthly
    psoSea water pressure at sea water surfaceMonthly
    soSea water salinityMonthly
    sobSea water salinity at sea floorMonthly
    sogaGlobal mean sea water salinityMonthly
    sosSea surface salinityMonthly
    sosgaGlobal average sea surface salinityMonthly
    thetaoSea water potential temperatureMonthly
    thetaogaGlobal average sea water potential temperatureMonthly
    tobSea water potential temperature at sea floorMonthly
    tosSea surface temperatureMonthly
    tosgaGlobal average sea surface temperatureMonthly
    uoSea water X velocityMonthly
    umoOcean mass X transportMonthly
    voSea water Y velocityMonthly
    vmoOcean mass Y transportMonthly
    wfoWater flux into sea waterMonthly
    woSea water vertical velocityMonthly
    wmoUpward ocean mass transportMonthly
    zosSea surface height above geoidMonthly
    omldamaxMean daily maximum ocean mixed-layer thickness defined by mixing schemeDaily
    areacelloGrid-cell area for ocean variablesFixed
    depthoSea floor depth below geoidFixed
    thkcelloOcean model cell thicknessFixed
    ugridoUGRID grid specificationFixed
    volcelloOcean grid-cell volumeFixed
    DownLoad: CSV
  • Bao, Q., and Coauthors, 2013: The flexible global ocean-atmosphere-land system model, spectral version 2: FGOALS-s2. Adv. Atmos. Sci., 30, 561−576, https://doi.org/10.1007/s00376-012-2113-9.
    Canuto, V. M., A. Howard, Y. Cheng, and M. S. Dubovikov, 2001: Ocean turbulence. Part I: One-point closure model—Momentum and heat vertical diffusivities. J. Phys. Oceanogr., 31, 1413−1426, https://doi.org/10.1175/1520-0485(2001)031<1413:OTPIOP>2.0.CO;2.
    Canuto, V. M., A. Howard, Y. Cheng, and M. S. Dubovikov, 2002: Ocean turbulence. Part II: Vertical diffusivities of momentum, heat, salt, mass, and passive scalars. J. Phys. Oceanogr., 32, 240−264, https://doi.org/10.1175/1520-0485(2002)032<0240:otpivd>2.0.co;2.
    Conkright, M. E., R. A. Locarnini, H. E. Garcia, T. D. O’Brien, T. P. Boyer, C. Stephens, and J. I. Antonov, 2002: World ocean atlas 2001: Objective analyses, data statistics, and figures, CD-ROM documentation. National Oceanographic Data Center, Silver Spring, MD, 17 pp.
    Craig, A. P., M. Vertenstein, and R. Jacob, 2012: A new flexible coupler for Earth system modeling developed for CCSM4 and CESM1. The International Journal of High Performance Computing Applications, 26, 31−42, https://doi.org/10.1177/1094342011428141.
    Cunningham, S. A., and Coauthors, 2007: Temporal variability of the Atlantic meridional overturning circulation at 26.5°N. Science, 317, 935−938, https://doi.org/10.1126/science.1141304.
    Danabasoglu, G., and Coauthors, 2014: North Atlantic simulations in coordinated ocean-ice reference experiments phase II (CORE-II). Part I: Mean states. Ocean Modelling, 73, 76−107, https://doi.org/10.1016/j.ocemod.2013.10.005.
    Danabasoglu, G., and Coauthors, 2016: North Atlantic simulations in coordinated ocean-ice reference experiments phase II (CORE-II). Part II: Inter-annual to decadal variability. Ocean Modelling, 97, 65−90, https://doi.org/10.1016/j.ocemod.2015.11.007.
    Ferreira, D., J. Marshall, and P. Heimbach, 2005: Estimating eddy stresses by fitting dynamics to observations using a residual-mean ocean circulation model and its adjoint. J. Phys. Oceanogr., 35, 1891−1910, https://doi.org/10.1175/JPO2785.1.
    Fetterer, F., K. Knowles, W. Meier, M. Savoie, and A. Windnagel, 2017: Updated daily. Sea Ice Index, Version 3. [Indicate subset used]. NSIDC, Boulder, Colorado, USA, https://doi.org/10.7265/N5K072F8.
    Gent, P. R., and J. C. McWilliams, 1990: Isopycnal mixing in ocean circulation models. J. Phys. Oceanogr., 20, 150−155, https://doi.org/10.1175/1520-0485(1990)020<0150:IMIOCM>2.0.CO;2.
    Griffies, S. M., and Coauthors, 2009: Coordinated ocean-ice reference experiments (COREs). Ocean Modelling, 26, 1−46, https://doi.org/10.1016/j.ocemod.2008.08.007.
    Huang, B., and Coauthors, 2017: Extended Reconstructed Sea Surface Temperature, Version 5 (ERSSTv5): Upgrades, validations, and intercomparisons. J. Climate, 30(20), 8179−8205, https://doi.org/10.1175/JCLI-D-16-0836.1.
    Large, W. G., and S. G. Yeager, 2004: Diurnal to decadal global forcing for ocean and sea-ice models: The data sets and flux climatologies. NCAR/TN-460+STR, CGD Division of the National Center for Atmospheric Research, https://doi.org/10.5065/D6KK98Q6.
    Lee, S. K., R. Lumpkin, M. O. Baringer, C. S. Meinen, M. Goes, S. F. Dong, H. Lopez, and S. G. Yeager, 2019: Global meridional overturning circulation inferred from a data-constrained ocean & sea-ice model. Geophys. Res. Lett., 46, 1521−1530, https://doi.org/10.1029/2018GL080940.
    Li, L. J., and Coauthors, 2013: The flexible global ocean-atmosphere-land system model, grid-point version 2: FGOALS-g2. Adv. Atmos. Sci., 30, 543−560, https://doi.org/10.1007/s00376-012-2140-6.
    Li, Y. W., H. L. Liu, and P. F. Lin, 2019: The role of thickness diffusivity coefficients in a climate ocean model. PhD dissertation, 142 pp.
    Lin, P. F., Y. Q. Yu, and H. L. Liu, 2013a: Long-term stability and oceanic mean state simulated by the coupled model FGOALS-s2. Adv. Atmos. Sci., 30, 175−192, https://doi.org/10.1007/s00376-012-2042-7.
    Lin, P. F., Y. Q. Yu, and H. L. Liu, 2013b: Oceanic climatology in the coupled model FGOALS-g2: Improvements and biases. Adv. Atmos. Sci., 30(3), 819−840, https://doi.org/10.1007/s00376-012-2137-1.
    Lin, P. F., and Coauthors, 2016: A coupled experiment with LICOM2 as the ocean component of CESM1. Journal of Meteorological Research, 30(1), 76−92, https://doi.org/10.1007/s13351-015-5045-3.
    Liu, H. L., P. F. Lin, Y. Q. Yu, and X. H. Zhang, 2012: The baseline evaluation of LASG/IAP Climate system Ocean Model (LICOM) version 2.0. Acta Meteorologica Sinica, 26(3), 318−329, https://doi.org/10.1007/s13351-012-0305-y.
    Lumpkin, R., and K. Speer, 2007: Global ocean meridional overturning. J. Phys. Oceanogr., 37, 2550−2562, https://doi.org/10.1175/JPO3130.1.
    Madec, G., and M. Imbard, 1996: A global ocean mesh to overcome the north pole singularity. Climate Dyn., 12(6), 381−388, https://doi.org/10.1007/BF00211684.
    Murray, R. J., 1996: Explicit generation of orthogonal grids for ocean models. J. Comput. Phys., 126(2), 251−273, https://doi.org/10.1006/jcph.1996.0136.
    Ohlmann, J. C., 2003: Ocean radiant heating in climate models. J. Climate, 16, 1337−1351, https://doi.org/10.1175/1520-0442-16.9.1337.
    Redi, M. H., 1982: Oceanic isopycnal mixing by coordinate rotation. J. Phys. Oceanogr., 12, 1154−1158, https://doi.org/10.1175/1520-0485(1982)012<1154:OIMBCR>2.0.CO;2.
    St. Laurent, L. C., H. L. Simmons, and S. R. Jayne, 2002: Estimating tidally driven mixing in the deep ocean. Geophys. Res. Lett., 29, 2106, https://doi.org/10.1029/2002GL015633.
    Steele, M., R. Morley, and W. Ermold, 2001: PHC: A global ocean hydrography with a high-quality Arctic Ocean. J. Climate, 14, 2079−2087, https://doi.org/10.1175/1520-0442(2001)014<2079:PAGOHW>2.0.CO;2.
    Tsujino, H., and Coauthors, 2018: JRA-55 based surface dataset for driving ocean—sea-ice models (JRA55-do). Ocean Modelling, 130, 79−139, https://doi.org/10.1016/j.ocemod.2018.07.002.
    Xiao, C., 2006: Adoption of a two-step shape-preserving advection scheme in an OGCM and its coupled experiment. M.S. thesis, Institute of Atmospheric Physics, Chinese Academy of Sciences, 89 pp. (in Chinese)
    Yu, R. C., 1994: A two-step shape-preserving advection scheme. Adv. Atmos. Sci., 11(4), 479−490, https://doi.org/10.1007/BF02658169.
    Yu, Y. Q., S. L. Tang, H. L. Liu, P. F. Lin, and X. L. Li., 2018: Development and evaluation of the dynamic framework of an ocean general circulation model with arbitrary orthogonal curvilinear coordinate. Chinese Journal of Atmospheric Sciences, 42(4), 877−889, https://doi.org/10.3878/j.issn.1006-9895.1805.17284. (in Chinese)
    Yu, Z. P., H. L. Liu, and P. F. Lin, 2017: A numerical study of the influence of tidal mixing on Atlantic meridional overturning circulation (AMOC) Simulation. Chinese Journal of Atmospheric Sciences, 41(5), 1087−1100, https://doi.org/10.3878/j.issn.1006-9895.1702.16263. (in Chinese)
    Zhang, X. H., and X. Z. Liang, 1989: A numerical world ocean general circulation model. Adv. Atmos. Sci., 6(1), 43−61, https://doi.org/10.1007/BF02656917.
  • [1] Yiwen LI, Hailong LIU, Mengrong DING, Pengfei LIN, Zipeng YU, Yongqiang YU, Yao MENG, Yunlong LI, Xiaodong JIAN, Jinrong JIANG, Kangjun CHEN, Qian YANG, Yaqi WANG, Bowen ZHAO, Jilin WEI, Jinfeng MA, Weipeng ZHENG, Pengfei WANG, 2020: Eddy-resolving Simulation of CAS-LICOM3 for Phase 2 of the Ocean Model Intercomparison Project, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1067-1080.  doi: 10.1007/s00376-020-0057-z
    [2] Jie ZHANG, Tongwen WU, Fang ZHANG, Kalli FURTADO, Xiaoge XIN, Xueli SHI, Jianglong LI, Min CHU, Li ZHANG, Qianxia LIU, Jinghui Yan, Min WEI, Qiang MA, 2021: BCC-ESM1 Model Datasets for the CMIP6 Aerosol Chemistry Model Intercomparison Project (AerChemMIP), ADVANCES IN ATMOSPHERIC SCIENCES, 38, 317-328.  doi: 10.1007/s00376-020-0151-2
    [3] Yaqi WANG, Zipeng YU, Pengfei LIN, Hailong LIU, Jiangbo JIN, Lijuan LI, Yanli TANG, Li DONG, Kangjun CHEN, Yiwen LI, Qian YANG, Mengrong DING, Yao MENG, Bowen ZHAO, Jilin WEI, Jinfeng MA, Zhikuo SUN, 2020: FGOALS-g3 Model Datasets for CMIP6 Flux-Anomaly-Forced Model Intercomparison Project, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1093-1101.  doi: 10.1007/s00376-020-2045-8
    [4] Binghao JIA, Longhuan WANG, Yan WANG, Ruichao LI, Xin LUO, Jinbo XIE, Zhenghui XIE, Si CHEN, Peihua QIN, Lijuan LI, Kangjun CHEN, 2021: CAS-LSM Datasets for the CMIP6 Land Surface Snow and Soil Moisture Model Intercomparison Project, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 862-874.  doi: 10.1007/s00376-021-0293-x
    [5] Yajuan SONG, Xinfang LI, Ying BAO, Zhenya SONG, Meng WEI, Qi SHU, Xiaodan YANG, 2020: FIO-ESM v2.0 Outputs for the CMIP6 Global Monsoons Model Intercomparison Project Experiments, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1045-1056.  doi: 10.1007/s00376-020-9288-2
    [6] Liang ZHAO, Yunwen LIU, Yihui DING, Qingquan LI, Wei DONG, Xinyong SHEN, Wei CHENG, Haoxin YAO, Ziniu XIAO, 2023: The Warm Arctic–Cold Eurasia Pattern and Its Key Region in Winter in CMIP6 Model Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 2138-2153.  doi: 10.1007/s00376-022-2201-4
    [7] Yuyang GUO, Yongqiang YU, Pengfei LIN, Hailong LIU, Bian HE, Qing BAO, Shuwen ZHAO, Xiaowei WANG, 2020: Overview of the CMIP6 Historical Experiment Datasets with the Climate System Model CAS FGOALS-f3-L, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1057-1066.  doi: 10.1007/s00376-020-2004-4
    [8] Xinyao RONG, Jian LI, Haoming CHEN, Jingzhi SU, Lijuan HUA, Zhengqiu ZHANG, Yufei XIN, 2021: The CMIP6 Historical Simulation Datasets Produced by the Climate System Model CAMS-CSM, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 285-295.  doi: 10.1007/s00376-020-0171-y
    [9] Shuai HU, Bo WU, Yiming WANG, Tianjun ZHOU, Yongqiang YU, Bian HE, Pengfei LIN, Qing BAO, Hailong LIU, Kangjun CHEN, Shuwen ZHAO, 2023: CAS FGOALS-f3-L Model Datasets for CMIP6 DCPP Experiment, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1911-1922.  doi: 10.1007/s00376-023-2122-x
    [10] Bian HE, Qing BAO*, Xiaocong WANG, Linjiong ZHOU, Xiaofei WU, Yimin LIU, Guoxiong WU, Kangjun CHEN, Sicheng HE, Wenting HU, Jiandong LI, Jinxiao LI, Guokui NIAN, Lei WANG, Jing YANG, Minghua ZHANG, Xiaoqi ZHANG, 2019: CAS FGOALS-f3-L Model Datasets for CMIP6 Historical Atmospheric Model Intercomparison Project Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, , 771-778.  doi: 10.1007/s00376-019-9027-8
    [11] Jiangbo JIN, He ZHANG, Xiao DONG, Hailong LIU, Minghua ZHANG, Xin GAO, Juanxiong HE, Zhaoyang CHAI, Qingcun ZENG, Guangqing ZHOU, Zhaohui LIN, Yi YU, Pengfei LIN, Ruxu LIAN, Yongqiang YU, Mirong SONG, Dongling ZHANG, 2021: CAS-ESM2.0 Model Datasets for the CMIP6 Flux-Anomaly-Forced Model Intercomparison Project (FAFMIP), ADVANCES IN ATMOSPHERIC SCIENCES, 38, 296-306.  doi: 10.1007/s00376-020-0188-2
    [12] Shuwen ZHAO, Yongqiang YU, Pengfei LIN, Hailong LIU, Bian HE, Qing BAO, Yuyang GUO, Lijuan HUA, Kangjun CHEN, Xiaowei WANG, 2021: Datasets for the CMIP6 Scenario Model Intercomparison Project (ScenarioMIP) Simulations with the Coupled Model CAS FGOALS-f3-L, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 329-339.  doi: 10.1007/s00376-020-0112-9
    [13] Bian HE, Yimin LIU, Guoxiong WU, Qing BAO, Tianjun ZHOU, Xiaofei WU, Lei WANG, Jiandong LI, Xiaocong WANG, Jinxiao LI, Wenting HU, Xiaoqi ZHANG, Chen SHENG, Yiqiong TANG, 2020: CAS FGOALS-f3-L Model Datasets for CMIP6 GMMIP Tier-1 and Tier-3 Experiments, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 18-28.  doi: 10.1007/s00376-019-9085-y
    [14] Bian HE, Xiaoqi ZHANG, Anmin DUAN, Qing BAO, Yimin LIU, Wenting HU, Jinxiao LI, Guoxiong WU, 2021: CAS FGOALS-f3-L Large-ensemble Simulations for the CMIP6 Polar Amplification Model Intercomparison Project, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1028-1049.  doi: 10.1007/s00376-021-0343-4
    [15] Ye PU, Hongbo LIU, Ruojing YAN, Hao YANG, Kun XIA, Yiyuan LI, Li DONG, Lijuan LI, He WANG, Yan NIE, Mirong SONG, Jinbo XIE, Shuwen ZHAO, Kangjun CHEN, Bin WANG, Jianghao LI, Ling ZUO, 2020: CAS FGOALS-g3 Model Datasets for the CMIP6 Scenario Model Intercomparison Project (ScenarioMIP), ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1081-1092.  doi: 10.1007/s00376-020-2032-0
    [16] Xiao DONG, Jiangbo JIN, Hailong LIU, He ZHANG, Minghua ZHANG, Pengfei LIN, Qingcun ZENG, Guangqing ZHOU, Yongqiang YU, Mirong SONG, Zhaohui LIN, Ruxu LIAN, Xin GAO, Juanxiong HE, Dongling ZHANG, Kangjun CHEN, 2021: CAS-ESM2.0 Model Datasets for the CMIP6 Ocean Model Intercomparison Project Phase 1 (OMIP1), ADVANCES IN ATMOSPHERIC SCIENCES, 38, 307-316.  doi: 10.1007/s00376-020-0150-3
    [17] Yuanxin LIU, Lijing CHENG, Yuying PAN, Zhetao TAN, John ABRAHAM, Bin ZHANG, Jiang ZHU, Junqiang SONG, 2022: How Well Do CMIP6 and CMIP5 Models Simulate the Climatological Seasonal Variations in Ocean Salinity?, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1650-1672.  doi: 10.1007/s00376-022-1381-2
    [18] Lanying CHEN, Renhao WU, Qi SHU, Chao MIN, Qinghua YANG, Bo HAN, 2023: The Arctic Sea Ice Thickness Change in CMIP6’s Historical Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 2331-2343.  doi: 10.1007/s00376-022-1460-4
    [19] Shengzhe CHEN, Jiping LIU, Yifan DING, Yuanyuan ZHANG, Xiao CHENG, Yongyun HU, 2021: Assessment of Snow Depth over Arctic Sea Ice in CMIP6 Models Using Satellite Data, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 168-186.  doi: 10.1007/s00376-020-0213-5
    [20] Bruno FERRERO, Marcos TONELLI, Fernanda MARCELLO, Ilana WAINER, 2021: Long-term Regional Dynamic Sea Level Changes from CMIP6 Projections, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 157-167.  doi: 10.1007/s00376-020-0178-4

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 29 September 2019
Manuscript revised: 19 November 2019
Manuscript accepted: 05 December 2019
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

LICOM Model Datasets for the CMIP6 Ocean Model Intercomparison Project

    Corresponding author: Hailong LIU, lhl@lasg.iap.ac.cn
  • 1. LASG, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029, China
  • 2. Computer Network Information Center, Chinese Academy of Sciences, Beijing 100190, China
  • 3. Department of Computer Science and Technology, Tsinghua University, Beijing 100084, China
  • 4. College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing 100049, China

Abstract: The datasets of two Ocean Model Intercomparison Project (OMIP) simulation experiments from the LASG/IAP Climate Ocean Model, version 3 (LICOM3), forced by two different sets of atmospheric surface data, are described in this paper. The experiment forced by CORE-II (Co-ordinated Ocean–Ice Reference Experiments, Phase II) data (1948–2009) is called OMIP1, and that forced by JRA55-do (surface dataset for driving ocean–sea-ice models based on Japanese 55-year atmospheric reanalysis) data (1958–2018) is called OMIP2. First, the improvement of LICOM from CMIP5 to CMIP6 and the configurations of the two experiments are described. Second, the basic performances of the two experiments are validated using the climatological-mean and interannual time scales from observation. We find that the mean states, interannual variabilities, and long-term linear trends can be reproduced well by the two experiments. The differences between the two datasets are also discussed. Finally, the usage of these data is described. These datasets are helpful toward understanding the origin system bias of the fully coupled model.

摘要: 本文描述了大气所大气科学和地球流体力学数值模拟国家重点实验室(LASG/IAP) 气候系统海洋模式第三版(LICOM3)参与海洋模式比较计划的两个模拟试验数据集。两个模拟试验采用不同大气表面强迫数据,采用规范的海洋-海冰参考试验第二阶段 (Co-ordinated Ocean–Ice Reference Experiments, Phase II, CORE-II) 作为外强迫的试验称为海洋模式比较计划1阶段(OMIP1,1948-2009),采用来自日本55年大气再分析表面数据驱动海洋-海冰模式 (JRA55-do)强迫试验称为海洋模式比较计划第二阶段(OMIP2,1958-2018)。首先,文章描述了LICOM从CMIP5到CMIP6的改进和试验基本设置。其次,利用来自观测气侯态和年际变化数据,对两个试验的模拟性能进行检验。结果表明模式很好再现了两个试验中的平均态、年际变化和长期趋势。同时,讨论了两个模拟试验数据间的差异。最后,介绍了数据集的使用情况。这两个数据集有助于了解全耦合模式的系统偏差来源。

    • The Ocean Model Intercomparison Project (OMIP) is one of the endorsed Model Intercomparison Projects in phase 6 of the Coupled Model Intercomparison Project (CMIP6). The purpose of OMIP is to understand the origin of systematic model biases. The subject also belongs to one of the scientific questions of CMIP6. The primary contribution of OMIP to the World Climate Research Programme’s Grand Challenges in Climate Science is in regional sea level change and near-term (climate/decadal) prediction. The OMIP experiment is a hindcast experiment for global coupled ocean–sea-ice models, inheriting the protocol from the Co-ordinated Ocean–Ice Reference Experiments, Phase II (CORE-II; Griffies et al., 2009). The experiments are driven by the prescribed atmospheric forcings from modified reanalysis and observational datasets. The turbulence momentum and heat fluxes are computed through the method of Large and Yeager (2004).

      The LASG/IAP Climate Ocean Model (LICOM) was developed at the State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics (LASG), Institute of Atmospheric Physics (IAP), Chinese Academy of Sciences (CAS) around 30 years in the late 1980s (Zhang and Liang, 1989). The latest version of LICOM is LICOM3, preparing for CMIP6. Currently, LICOM3 is the ocean component of two climate system models of CMIP6: the CAS Flexible Global Ocean–Atmosphere–Land System model, finite-volume version 3 (CAS FGOALS-f3) and the CAS Flexible Global Ocean–Atmosphere–Land System model, grid-point version 3 (CAS FGOALS-g3), which in CMIP5 were FGOALS-s2 (Bao et al., 2013, Lin et al., 2013a) and FGOALS-g2 (Li et al., 2013; Lin et al., 2013b), respectively. In addition, the previous version of LICOM, LICOM2.0 (Liu et al., 2012), is also employed as the ocean component in CAS-ESM (the CAS Earth System Model), version 1.0 (Dr. Minghua ZHANG, personal communication, 2016).

      The OMIP simulations of LICOM3 were finished in June 2019 and the data have submitted to the Earth System Grid (ESG) data server (https://esgf-nodes.llnl.gov/projects/cmip6/). The purpose of the present paper is to provide a comprehensive description of the OMIP datasets of LICOM3 for a variety of users. We document detailed descriptions of the model configurations and experiments, as well as the algorithm of the diagnostic variables. Section 2 presents the model descriptions and experiment designs. Section 3 presents a basic technical validation of the LICOM3 experiments. Section 4 describes the datasets. Section 5 provides usage notes.

    2.   Model and experiments
    • Since LICOM2.0 (Liu et al., 2012), LICOM has been substantially upgraded in the interface with the flux coupler, the dynamic core, and the physical packages (Table 1). First, we upgraded the LICOM interface from the NCAR flux coupler 6 to coupler 7 (Lin et al., 2016), because the new version has been optimized for high-resolution modeling (Craig et al., 2012). Here, LICOM3 coupled with the Community Ice Code, version 4 (CICE4), i.e., the ocean–ice coupled model, is used to conduct the OMIP experiments. The prescribed atmospheric data have been input from the atmospheric data model and then passed to the coupler to drive the ocean–sea-ice coupled model.

      ConfigurationLICOM2.0LICOM3
      GridGridLongitude/LatitudeTripolar
      Resolution~1°, 30 levels~1°, 30 or 80 levels
      Dynamic coreTracer advectionCentral differential schemePreserved shape scheme (Yu, 1994)
      Momentum time integrationExplicitImplicit
      PhysicsDiapycnal mixingMixing in the mixed layer (Canuto et al., 2001, 2002)Mixing in the mixed layer (Canuto et al., 2001) and internal tide mixing (St. Laurent et al., 2002)
      Isopycnal mixingIsopycnal mixing (Redi, 1982) and advection (Gent and McWilliams, 1990)Isopycnal mixing (Redi, 1982) and advection (Gent and McWilliams, 1990) with N2 thickness diffusivity (Ferreira et al., 2005)
      Computing technicsCoupler interfaceNCAR Flux Coupler 6NCAR Flux Coupler 7
      Parallel1D MPI and OMP2D MPI and OMP
      DataInitial conditionWOA01 (Conkright et al., 2002)PHC3.0 (Steele et al., 2001)
      BathymetryDBDB5 (https://www.bodc.ac.uk/resources/inventories/edmed/report/356/)ETOPO2 (https://ngdc.noaa.gov/mgg/global/etopo2.html)

      Table 1.  Comparison of model configurations between two versions of LICOM (LICOM2.0 and LICOM3).

      Second, the orthogonal curvilinear coordinate (Madec and Imbard, 1996; Murray, 1996) has been introduced in LICOM. The tripolar grid can be used with the North Pole split into two poles on the land in the Northern Hemisphere (NH) at (65°N, 65°E) and (65°N, 115°W). This improvement has solved the problem of the singularity of the North Pole in the normal longitude–latitude grid. Meanwhile, the spatial filter in the high latitudes of LICOM has also been eliminated, and the scalability and efficiency of the parallel algorithm has been extensively improved. Besides, preserved shape advection in the tracer formulation (Xiao, 2006) and implicit vertical viscosity (Yu et al., 2018) are also employed in the present version.

      Third, the tidal mixing of St. Laurent et al. (2002) and the buoyancy frequency (N2)–related thickness diffusivity of Ferreira et al. (2005) have been introduced into LICOM after CMIP5. The effects of tidal mixing on the Atlantic meridional overturning circulation (AMOC) was preliminary evaluated by Yu et al. (2017). The effects of thickness diffusivity with temporal and spatial variation were evaluated by Li et al. (2019). Besides, the chlorophyll-a-dependent solar penetration of Ohlmann (2003), the vertical diffusivity of Canuto et al. (2001, 2002), and the isopycnal mixing of Redi (1982) and Gent and McWilliams (1990, GM90 hereafter) are also used in LICOM3. The isopycnal mixing coefficient is constant, with a value of 300 m2 s−1. The thickness mixing coefficient (i.e., diffusivity) for GM90 employs the scheme of Ferreira et al. (2005), in which the coefficient is dependent on the spatial distribution of N2 and varies with location and time. The coefficient is set to 300 m2 s−1 within the mixed layer or in the coastal region where the water depth is shallower than 60 m, while it varies between 300 and 2000 m2 s−1 in other places.

      The virtual salinity flux is computed as the freshwater flux multiplied by a constant salinity of 34.7 psu. A restoring term with a piston velocity of 20 m yr−1 has been applied to the virtual salinity flux. If there is sea ice, a piston velocity of 50 m (20 d)−1 is applied under sea ice.

      Here, the low-resolution LICOM3, which is used both for CMIP6 and OMIP, has 360 and 218 grid numbers for the zonal and the meridional directions, respectively. LICOM3 has two resolutions in the vertical (30 and 80 levels), but only the 30-level resolution is used for OMIP and CMIP6 to save computing resources. The depths of the W-grid and T-grid are shown in Table 2. At the same time, an eddy-resolving version of LICOM3 has also been developed, with a horizontal resolution of about 10 km at the equator and 2.7 km around the Antarctic, and 55 levels in the vertical. This version has also been implemented in an ocean forecast system for short-term ocean prediction.

      LevelDepth for TDepths for W
      1−50
      2−15−10
      3−25−20
      4−35−30
      5−45−40
      6−55−50
      7−65−60
      8−75−70
      9−85−80
      10−95−90
      11−105−100
      12−115−110
      13−125−120
      14−135−130
      15−145−140
      16−156.9303−150
      17−178.4277−163.8606
      18−222.5018−192.9948
      19−303.1057−252.0088
      20−432.5961−354.2027
      21−621.1931−510.9896
      22−876.5334−731.3966
      23−1203.337−1021.67
      24−1603.2−1385.003
      25−2074.526−1821.396
      26−2612.596−2327.656
      27−3209.772−2897.536
      28−3855.835−3522.009
      29−4538.428−4189.662
      30−5243.597−4887.194
      31−5600

      Table 2.  Depths of model levels for the T- and W-grid. Positive is upward.

    • The OMIP experiment follows the protocol of CORE-II, which is an ocean–sea-ice coupled hindcast simulation forced by about 60 years of modified reanalysis atmospheric variables with diurnal to decadal signals. Usually, the experiment is conducted for five or six cycles to reach an equilibrium state. Details of the experiments are shown in Table 3. Here, LICOM3 is coupled with CICE4 using the NCAR flux coupler 7. Two standard OMIP experiments have been conducted: one forced with CORE-II data derived from NCEP–NCAR reanalysis (Large and Yeager, 2004), named OMIP1; and the other forced with the surface dataset for driving ocean–sea-ice models based on Japanese 55-year atmospheric reanalysis (JRA55-do, Tsujino et al., 2018), named OMIP2. The atmospheric variables include atmospheric surface wind vectors at 10 m, temperature at 10 m, specific humidity at 10 m, air density at 10 m, precipitation (both rain and snow), surface downward shortwave radiation, downward longwave radiation, and sea level pressure. The period of the forcing data is 62 years (1948–2009) for CORE-II, while the period is 61 years (1958–2018) for JRA55-do. Although the frequencies of the two datasets are different, we use 6-h intervals for both datasets to force the model. Here, experiments with six cycles (one cycle corresponds to 1948-2009 for CORE-II and 1958−2018 for JRA55-do) have been completed and uploaded onto ESG nodes for both OMIP1 and OMIP2. Details of the primary output and diagnostic variables are given in Table 4.

      Experiment_idModelInitial condition (TS/currents)Forcing dataPeriodFrequency
      OMIP1LICOM3/CICE4.0PHC3.0/ZeroCORE-II1948–2009 (6 cycles)6 h
      OMIP2LICOM3/CICE4.0PHC3.0/ZeroJRA55-do1958–2018 (6 cycles)6 h

      Table 3.  Descriptions of the experiments of LICOM3.

      NameDescriptionFrequency
      hfbasinNorthward ocean heat transportMonthly
      hfbasinpmadvNorthward ocean heat transport due to parameterized mesoscale advectionMonthly
      hfbasinpmdiffNorthward ocean heat transport due to parameterized mesoscale diffusionMonthly
      hfdsDownward heat flux at sea water surfaceMonthly
      masscelloOcean grid-cell mass per areaMonthly
      mlotstOcean mixed-layer thickness defined by sigma TMonthly
      msftbarotOcean barotropic mass streamfunctionMonthly
      msftmzOcean meridional overturning mass streamfunctionMonthly
      msftmzmpaOcean meridional overturning mass streamfunction due to parameterized mesoscale advectionMonthly
      obvfsqSquare of Brunt–Vaisala frequency in sea waterMonthly
      pboSea water pressure at sea floorMonthly
      psoSea water pressure at sea water surfaceMonthly
      soSea water salinityMonthly
      sobSea water salinity at sea floorMonthly
      sogaGlobal mean sea water salinityMonthly
      sosSea surface salinityMonthly
      sosgaGlobal average sea surface salinityMonthly
      thetaoSea water potential temperatureMonthly
      thetaogaGlobal average sea water potential temperatureMonthly
      tobSea water potential temperature at sea floorMonthly
      tosSea surface temperatureMonthly
      tosgaGlobal average sea surface temperatureMonthly
      uoSea water X velocityMonthly
      umoOcean mass X transportMonthly
      voSea water Y velocityMonthly
      vmoOcean mass Y transportMonthly
      wfoWater flux into sea waterMonthly
      woSea water vertical velocityMonthly
      wmoUpward ocean mass transportMonthly
      zosSea surface height above geoidMonthly
      omldamaxMean daily maximum ocean mixed-layer thickness defined by mixing schemeDaily
      areacelloGrid-cell area for ocean variablesFixed
      depthoSea floor depth below geoidFixed
      thkcelloOcean model cell thicknessFixed
      ugridoUGRID grid specificationFixed
      volcelloOcean grid-cell volumeFixed

      Table 4.  Descriptions of dataset variables of Priority 1.

    3.   Validation
    • The basic results of OMIP1 and OMIP2 from LICOM3 are evaluated before submitting the whole datasets. Here, some primary aspects are selected to demonstrate the performance of the model. After six cycles of integration, the surface fluxes are balanced by the adjustment of intrinsic processes and the model reaches a quasi-steady state. Figure 1 shows six cycles of the global mean sea surface temperature (SST), the volume mean temperature, the global mean sea surface salinity (SSS), and the volume mean ocean temperature for both OMIP1 (blue curve) and OMIP2 (red curve). The global mean SST value quickly reaches an equilibrium state (Fig. 1a). The global volume mean temperature increases at a rate of 0.09°C (100 yr)−1 due to the input of net heat flux at the sea surface (Fig. 1b). The gradual decrease in net heat flux (from 0.4 W m−2 for the first cycle to 0.17 W m−2 for the sixth cycle), due to the SST increase, leads to the decrease in the temperature trend from cycle to cycle. The simulated global mean temperature reaches a quasi-stable state after the fifth to the sixth cycle. The global mean SSS value has a small trend, with a value of 0.02–0.03 psu (100 yr)−1, and reaches a near steady state after the second cycle. Meanwhile, the global mean volume ocean salinity has a trend of 0.01 psu (100 yr)−1 due to the positive virtual salt flux at the sea surface.

      Figure 1.  Annual global mean (a) SST (units: °C), (b) volume ocean temperature (VOT; units: °C), (c) SSS (units: psu), and (d) volume ocean salinity (VOS; units: psu) for OMIP1 (cyan) and OMIP2 (purple) during all the six cycles. The black lines in the figure indicate the reference value calculated from WOA13 observation.

      The global annual mean SSTs from the last cycle of the two experiments are compared with Extended Reconstructed SST, version 5 (ERSST.v5) data (Huang et al., 2017), which serve as the observed reference values (Fig. 2a). The SSTs for OMIP1 and OMIP2 follow the observation well. The strong, warm ENSO events can be captured very well; for instance, 1982–83, 1997–98 and 2015–16. The correlation coefficient between OMIP2 and the observation is 0.95, which is much higher than that (0.81) between OMIP1 and the observation during the period of 1958–2009. After removing the linear trend, the standard deviations (STDs) of SST, indicative of the amplitude of interannual–decadal variability, are 0.08°C and 0.07°C for OMIP1 and OMIP2, respectively. However, the simulated global warming trends [calculated by linear trends; 0.06°C (10 yr)−1] are smaller than the observed ones [0.09°C (10 yr)−1] during 1958–2009, indicating a relatively fast vertical heat exchange in LICOM3.

      Figure 2.  (a) Annual global mean SST (units: °C), (b) annual mean SIC in the NH and SH, (c) steric sea level, and (d) AMOC, for OMIP1 (cyan), OMIP2 (purple) and ERSST.v5/NSIDC/Argo/RAPID (black) during the sixth cycles. The STDs (removing linear trend), linear trends and correlation coefficients (SST during 1958–2009, SIC for 1980–2009, 2005–2018 for Argo, 2004–16 for RAPID) are provided in the figures.

      Associated with global warming, the sea-ice areas in the polar regions also shrink significantly. The change in sea ice will further affect the albedo and greatly perturb the heat entering the ocean. The simulated annual mean NH and Southern Hemisphere (SH) sea-ice cover (SIC) are shown along with their observed values from the National Snow and Ice Data Center (NSIDC; Fetterer et al., 2017; Fig. 2b). The annual mean values in the NH for models and observations match very well. The correlation coefficients between OMIP1 and OMIP2 and observations during 1980–2009 are 0.89 and 0.96, respectively. The simulated amplitude of interannual–decadal variability for OMIP2 is also better than that for OMIP1. The STDs are both 0.2 × 106 km2 for NSIDC and OMIP2, but much larger for OMIP1 (0.34 × 106 km2).

      In the SH, the results from OMIP2 are also much closer to observation than that of OMIP1. The annual mean SIC values are overestimated by about 1.0–2.0 × 106 km2 in OMIP1 and OMIP2, although the latter is closer to the observation. The correlation coefficient between the simulation and the observation is 0.81 for OMIP2, but only 0.41 for OMIP1, during 1980–2009. The STDs are 0.24, 0.47 and 0.26 × 106 km2 for NSIDC, OMIP1 and OMIP2, respectively.

      The simulated linear trends of SIC are negative for both the NH and SH. However, both OMIP1 and OMIP2 underestimate the magnitude, which is similar to the situation for SST. The observed linear trends are −0.50 and 0.15 × 106 km2 (10 yr)−1 for the NH and SH, respectively. For comparison, the simulated trends are −0.48 and 0.08 × 106 km2 (10 yr)−1 for OMIP1 and −0.38 and 0.14 × 106 km2 (10 yr)−1 for OMIP2.

      The global mean thermosteric sea level and the volume transport at 26.5°N in the Atlantic, which is used to measure the magnitude of AMOC, are also shown in Fig. 2 against the observational data from the Array for Real-time Geostrophic Oceanography (Argo; http://apdrc.soest.hawaii.edu/projects/Argo/data/gridded/On_standard_levels/index-1.html) and RAPID (RAPID/MOCHA/WBTS array; Cunningham et al., 2007). The global mean thermosteric sea level reflects the heat uptake of the ocean. Although the simulation can reproduce the interannual and interdecadal variations of the whole system, the increasing/decreasing trend magnitudes in the model are also smaller than those in the observation, which is consistent with the trends of both SST and SIC. The simulations of AMOC for OMIP1 and OMIP2 are larger than that from RAPID, although the simulated values for OMIP2 are closer to the observation. The simulations for both OMIP1 and OMIP2 can capture the weakened AMOC from 2005 to 2009, but the simulation for OMIP2 fails to reproduce the increasing AMOC transport after 2012. This bias can also be found in other OMIP2 simulations, indicating a possible issue of OMIP2 forcing in the northern Atlantic Ocean (Dr. Hiroyuki TSUJINO, personal communication, 2019).

    • To understand the origin of systematic biases of ocean model, the biases of SST and SSS are shown in Figs. 3ab and Figs. 3de, respectively. In general, the SST biases for the two experiments are similar, including large cold SST biases located in the regions of strong western boundary currents and the Arctic connected with the North Atlantic, and warm SST biases in the eastern boundary regions. The root-mean-square errors (RMSEs) of SST are 0.66°C for OMIP1 and 0.63°C for OMIP2. The improvements in SST simulation for OMIP2 occur in the eastern boundary, the warm pool, southern Indian Ocean, the Norwegian Sea, and Barents Sea (Fig. 3c).

      Figure 3.  Simulations (contours) and biases (shaded) of SST (units: °C) for (a) OMIP1 and (b) OMIP2, and (c) the difference betweeen OMIP2 and OMIP1 (OMIP2 minus OMIP1). (d–f) As in (a–c), respectively, but for SSS (units: psu). The annual SST and SSS distributions employ the data from 1980–2009 of the sixth cycle. The RMSEs, mean values, and minimim and maximal values are provided in the figures.

      The SSS for OMIP1 and OMIP2 have the same bias patterns and the same RMSEs (0.45 psu). The common biases include larger than 1.5 psu salty biases in the East Siberian Sea, Chukchi Sea and Beaufort Sea of the Arctic Ocean; 0.2–0.8 psu salty biases in the tropical Atlantic Ocean; ~0.2 psu salty biases in the tropical Pacific and western Indian Ocean; fresher than −1.5 psu biases in the Gulf Stream region; and fresher biases of about −0.8 to −0.2 psu in the Greenland Sea, Barents Sea, Southern Ocean south of 40°S, and high-latitude North Pacific north of 30°N.

      The improvements in SSS simulation for OMIP2 occur in the South China Sea, eastern Indian Ocean, western Atlantic Ocean, and in the Southern Ocean (south of 40°S; Fig. 3f). However, the SSS biases for OMIP2 increase in the western Indian Ocean. In the eastern Indian Ocean and western Pacific warm pool near the western coast, the salty SSS bias becomes the fresh SSS bias from OMIP1 to OMIP2. Meanwhile, the salty biases for OMIP2 also increase slightly in the subtropics of the South Pacific.

    • The simulated sea surface height (SSH) is compared with satellite data from AVISO (Archiving, Validation and Interpretation of Satellite Oceanographic Data; http://www.aviso.altimetry.fr/) in Fig. 4. The observed high SSH is located in the western Pacific and Atlantic, eastern Indian Ocean, and southwest Indian Ocean, while the low SSH is located in the Southern Ocean, Arctic Ocean, and sub-polar gyres in the high-latitude Atlantic and Pacific. The model can capture all the high or low features very well. The spatial correlations between the simulations and observation are about 0.98. This indicates that the upper-layer horizontal circulations are also simulated well.

      Figure 4.  The (a) observed and (b, c) OMIP1- and OMIP2-simulated SSH (units: m) during 1993–2009 of the sixth cycle. The spatial correlations between the observation and simulation are noted in the top right.

    • The AMOC results are displayed in Fig. 5. Both OMIP1 and OMIP2 can capture the features of that derived from the World Ocean Circulation Experiment hydrographic sections (Lumpkin and Speer, 2007). In the upper ocean (0–400 m), there are wind-driven cells connecting the tropics and subtropics. The cell is northward (southward) in the NH (SH). In the NH, the cell includes an equatorial current system and western boundary currents (Gulf Stream). In the middle and high latitudes of the NH, the upper-ocean cell is southward, corresponding to the sub-polar gyre. The maximal volume transport of North Atlantic Deep Water (NADW), which is located at the latitude of 38°N and a depth of about 1000 m, can reach 22 Sv. The upper branch of NADW can enter the Arctic Ocean (> 60°N). The low branch of NADW returns southward. The minimal volume transport of Antarctic Bottom Water (AABW) is −6 Sv and limited below 3500 m.

      Figure 5.  The simulated AMOC (units: Sv; 1Sv = 1 × 106 m3 s−1) for (a) OMIP1 and (b) OMIP2 during the last 30 years (1980–2009) of the sixth cycle. (c) AMOC values at the latitude of 26.5°N for OMIP1, OMIP2 and RAPID (2005–09). The maximal values and STD (+/-) between 2005–09 are noted. The y-axises represent ocean depth (units: km).

      Compared with the observed one at 26.5°N during the period 2005–09, the depth of maximal AMOCs is close to the observed. The depth from NADW shifting to AABW is located at about 3000 m for both simulations, which is shallower than the 4500 m for the observed and robust diagnostic values from Lee et al. (2019). The upward shift of the transition depth is a common problem for surface forced ocean-ice simulations (Danabasoglu et al., 2014, 2016). The simulated magnitudes of maximal AMOC transport are likely overestimated by 0.5–2.2 Sv than observation at 26.5°N. At 26.5°N, the AMOC for OMIP2 is about 18.24 Sv, which is slightly larger than the observed value from RAPID (17.57 Sv), but smaller than the value for OMIP1 (20.75 Sv).

    • The interannual–decadal SST amplitude is denoted as the STD. The STDs of SST are presented in Fig. 6. The simulated amplitudes from OMIP1 and OMIP2 can capture well the observed one, with spatial correlations of 0.96. The large amplitudes are located in the tropical Pacific, the North Atlantic (> 40°N) and North Pacific (> 30°N) for the observation and simulations. However, the amplitudes are overestimated in the tropical Pacific and North Atlantic both in OMIP1 and OMIP2. Further comparisons show the simulated amplitude is closer to the observed in OMIP2 than that in OMIP1, including the tropical Pacific and North Atlantic. In the Niño3 region, the amplitudes are overestimated by about 32% and 16% for OMIP1 and OMIP2, respectively, compared with the observed (0.91). This indicates the OMIP2 forcing is better for simulating interannual–decadal variabilities than that of OMIP1.

      Figure 6.  The STDs (units °C) of observed SST anomalies from (a) ERSST.v5 and (b, c) OMIP1- and OMIP2-simulated SST anomalies during 1960–2009. Before calculating the STD, both the linear trend and annual cycle were removed from the SST. The spatial correlation coefficients (R) with the observation are denoted in the top middle. The STDs in the Niño3 region (5°S–5°N, 150°–90°W) are denoted in the top right.

    4.   Data records
    • The two sets of data, OMIP1 and OMIP2, have been uploaded onto ESG nodes and can be found at https://esgf-nodes.llnl.gov/projects/cmip6/. The dataset format is Network Common Data Form (NetCDF), version 4. Although the model outputs are double precision, we converted all the variables into single precision for analysis. These data can be easily handled by common computer programming languages and professional software, such as Climate Data Operators (https://code.mpimet.mpg.de/projects/cdo/) or NetCDF Operator (http://nco.sourceforge.net).

    5.   Usage notes
    • The original model outputs are on a tripolar grid with two poles set on land in the NH. The horizontal grid numbers are 360 and 218 in the zonal and meridional directions, respectively. The original grid distribution is kept and the format is slightly changed to CMOR (Climate Model Output Rewriter) file structure as required by OMIP. The data have 30 vertical levels and the original vertical level is not changed on ESG nodes. The first level is at the depth of 5 m with the thickness of 10 m. The depths of each level can be found in Table 3.

      The variables of Priority 1 for OMIP include monthly ocean temperature, salinity, ocean velocities, SSH, etc. (Table 4). There are 31 Priority 1 variables. Additionally, the daily maximal ocean mixed-layer thickness, defined as the depth that potential density is 0.03 kg m−3 larger than that on the sea surface, is also provided. Data grid information, such as area, mask, cell thickness and volume for every ocean grid, is also provided.

      Acknowledgements. This study was supported by the National Key R&D Program for Developing Basic Sciences (Grant Nos. 2016YFC1401401 and 2016YFC1401601), the Strategic Priority Research Program of the Chinese Academy of Sciences (Grant No. XDC01000000), and the National Natural Science Foundation of China (Grants Nos. 41576026, 41576025, 41776030, 41931183 and 41976026).

    Data availability statement
    • The data that support the findings of this study are available from https://esgf-nodes.llnl.gov/projects/cmip6/. The citation for OMIP1 is “CAS FGOALS-f3-L model output prepared for CMIP6 OMIP omip1. Earth System Grid Federation. http://doi.org/10.22033/ESGF/CMIP6.3413”. The citation for OMIP2 is “CAS FGOALS-f3-L model output prepared for CMIP6 OMIP omip2. Earth System Grid Federation. http://doi.org/10.22033/ESGF/CMIP6.3419”.

    Disclosure statement
    • No potential conflict of interest was reported by the authors.

      Open Access This article is distributed under the terms of the Creative Commons Attribution License which permits any use, distribution, and reproduction in any medium, provided the original author(s) and the source are credited.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return