Advanced Search
Article Contents

Model Uncertainty Representation for a Convection-Allowing Ensemble Prediction System Based on CNOP-P

Fund Project:

National Key Research and Development Program of China (Grant No. 2017YFC1501904)


doi: 10.1007/s00376-020-9262-z

  • Formulating model uncertainties for a convection-allowing ensemble prediction system (CAEPS) is a much more challenging problem compared to well-utilized approaches in synoptic weather forecasting. A new approach is proposed and tested through assuming that the model uncertainty should reasonably describe the fast nonlinear error growth of the convection-allowing model, due to the fast developing character and strong nonlinearity of convective events. The Conditional Nonlinear Optimal Perturbation related to Parameters (CNOP-P) is applied in this study. Also, an ensemble approach is adopted to solve the CNOP-P problem. By using five locally developed strong convective events that occurred in pre-rainy season of South China, the most sensitive parameters were detected based on CNOP-P, which resulted in the maximum variations in precipitation. A formulation of model uncertainty is designed by adding stochastic perturbations into these sensitive parameters. Through comparison ensemble experiments by using all the 13 heavy rainfall cases that occurred in the flood season of South China in 2017, the advantages of the CNOP-P-based method are examined and verified by comparing with the well-utilized stochastically perturbed physics tendencies (SPPT) scheme. The results indicate that the CNOP-P-based method has potential in improving the under-dispersive problem of the current CAEPS.
    摘要: 相对于全球集合预报,对流尺度集合预报(CAEPS)中有关模式不确定性的研究缺乏系统性和理论基础,成为目前研究的热点和难点。针对强对流天气发展迅速、非线性强等特点,本文充分考虑了对流尺度模式误差非线性快速增长特征,提出基于条件非线性最优参数扰动(CNOP-P)方法研究CAEPS模式不确定性问题的新思路。同时,在CNOP-P方法的计算中,本文采用了集合求解算法,不依赖于切线伴随模式。我们首先选取了2017年中国华南前汛期5个典型暖区暴雨个例求解CNOP-P,得到对累积降水预报产生最大影响的参数扰动集合,挑选敏感物理参数;然后,通过随机扰动敏感参数来描述模式不确定性,构造基于CNOP-P的模式扰动方案,并将该方案与目前广泛应用的随机物理参数化倾向扰动方案进行对比,展开对流尺度集合预报关于模式不确定性的敏感性试验。2017年华南地区13个强对流过程的检验结果显示:基于CNOP-P的模式扰动方案对于提高对流尺度集合离散度方面具有应用潜力。
  • 加载中
  • Figure 1.  Model forecast domain (D1: 18.5°–27.5°N, 105.0°–118.5°E) and computation domain for CNOP-P (D2: 20°–26°N, 106.5°–117°E).

    Figure 2.  The 6–24-h time-averaged differences of S/R (spread/RMSE) for 13 cases. Black lines represent EXP2’s S/R on four variables, and columns represents the difference between EXP2 and EXP1 (i.e. EXP2 minus EXP1). (a) 2-m temperature; (b) 10-m wind speed; (c) 2-m specific humidity; (d) hourly precipitation.

    Figure 3.  As in Fig. 2 but for outlier frequency.

    Figure 4.  As in Fig. 2 but for CRPS.

    Figure 5.  Vertical profiles of ensemble spread for the 24-h forecasts of 13 cases: (a) zonal wind (units: m s−1); (b) temperature (units: °C); (c) specific humidity (units: g kg−1).

    Figure 6.  The 6–24-h time-averaged S/R (spread/RMSE) for the case initialized at 1200 UTC 6 May. EXP_4 (red), EXP_6 (blue), EXP_8 (black) EXP_10 (grey), and EXP_12 (green) are five ensemble forecasts that perturbed the first 4, 6, 8, 10 and 12 parameters, respectively: (a) 2-m temperature; (b) 10-m wind speed; (c) 2-m specific humidity; (d) hourly precipitation.

    Figure 7.  As in Fig. 6 but for outlier frequency.

    Figure 8.  As in Fig. 6 but for CRPS.

    Table A1.  Summary of operational convection-allowing ensemble prediction systems.

    CenterModelHorizontal resolution
    (km)
    Forecast
    range (h)
    MembersInitial perturbationsLateral
    boundary
    conditions
    Model
    uncertainty
    DWDCOSMO-DE-EPS2.82720Downscaled and
    recentered perturbations
    from IFS, GFS,
    ICON, GSM
    Downscaled
    from IFS,
    GFS,ICON,
    GSM
    Multi-parameter
    Met OfficeMOGREPS-UK2.25412Downscaled and
    recentered
    perturbations from
    MOGREPS-G
    Downscaled from MOGREPS-GRP
    Météo-FranceAROME-EPS2.54512Downscaled and recentered perturbations from PEARPDownscaled from PEARPSPPT
    RHMCCOSMO-Ru2-EPS2.24810Downscaled from COSMO-S14-EPSDownscaled from COSMO-S14-EPS
    KMALENS34512Downscaled from EFSGDownscaled from EFSGRP
    CAPSSSEF36020Downscaled from SREFDownscaled from SREFMulti-
    modelmulti-
    physics
    NCAREnKF-based EPS34810EAKF analysisDownscaled from GFS+ random draws
    DownLoad: CSV

    Table 1.  Twenty-three typical heavy rainfall cases that occurred in South China in 2017.

    CaseInitial forecast
    time (UTC)
    Period of heavy rainfall (UTC)Weather type Trigger condition
    11200 6 May2100 6 May–0000 7 MayRainstormConvective systems in the
    warm sector
    20000 9 May0600 9 May–0900 9 MayRainstormConvective systems in the
    warm sector
    30000 15 Jun0900 15 Jun–1200 15 JunWide-area rainstormConvective systems in the
    warm sector
    40000 20 Jun0900 20 Jun–1200 20 JunWide-area rainstormConvective systems in the
    warm sector
    50000 26 Jun0900 26 Jun–1200 26 JunRainstorm, thunderstorm galeConvective systems in the
    warm sector
    60000 8 May0900 8 May–1200 8 MayWide-area rainstorm,
    thunderstorm gale
    Cold front
    71200 14 May1800 14 May–2100 14 MayWide-area rainstormCold front
    80000 23 May0900 23 May–1200 23 MayWide-Wide-area rainstorm,
    thunderstorm gale gale
    Cold front
    90000 6 Jun0600 6 Jun–0900 6 JunWide-area rainstorm,
    thunderstorm gale
    Cold front
    100000 16 Jun0700 16 Jun–1000 16 JunWide-Wide-area rainstorm,
    thunderstorm gale gale
    Cold front
    111200 20 Apr2100 20 Apr–0000 21AprRainstorm, thunderstorm gale, hailCold front
    121200 3 May1800 3 May–0000 4 MayWide-area rainstorm, thunderstorm gale, hailSquall line
    131200 23 May1700 23 May–2000 23 MayWide-area rainstormCold front
    141200 1 Jul0000 2 Jul–0300 2 JulRainstorm, thunderstorm galeConvective systems in the
    warm sector
    150000 7 Jul0700 7 Jul–1000 7 JulRainstormConvective systems in the
    warm sector
    160000 10 Jul0600 10 Jul–0900 10 JulRainstorm, thunderstorm galeConvective systems in the
    warm sector
    170000 19 Jul0600 19 Jul–0900 19 JulRainstorm, thunderstorm galeConvective systems on the edge
    of subtropical high and in
    the warm sector
    180000 21 Jul0800 21 Jul–1000 21 JulRainstorm, thunderstorm galeConvective systems on the edge
    of subtropical high and in
    the warm sectorarm sector
    190000 31 Jul0600 31 Jul–0900 31 JulWide-area rainstorm,
    thunderstorm gale, hail
    Cold front
    200000 1 Aug0600 1 Aug–0900 1 AugRainstorm, thunderstorm galeConvective systems in the
    warm sector
    210000 10 Aug0800 10 Aug–1000 10 AugRainstorm, thunderstorm galeConvective systems on the edge of subtropical high
    220000 22 Aug1200 22 Aug–1500 22 AugRainstorm, thunderstorm galeTyphoon
    230000 23 Aug0800 23 Aug–1000 23 AugRainstormTyphoon
    DownLoad: CSV

    Table 2.  The chosen physical parameters for computing CNOP-P.

    IndexSchemeParameterDescriptionDefault
    P1MRFrlamAsymptotic mixing length (m)150
    P2MRFbrcrCritical Richardson number0.5
    P3MRFpfacProfile shape exponent for calculating the momentum diffusion coefficient2
    P4MRFcfacCoefficient for Prandtl number at the top of the surface layer7.8
    P5WSM6avtrA constant for terminal velocity of rain841.9
    P6WSM6avtgA constant for terminal velocity of graupel330
    P7WSM6avtsA constant for terminal velocity of snow11.72
    P8WSM6n0rIntercept parameter of rain (m−4)8.0 × 106
    P9WSM6n0gIntercept parameter of graupel (m−4)4.0 × 106
    P10WSM6dengDensity of graupel (kgm−3)500
    P11WSM6peautMean collection efficiency0.55
    P12WSM6xncrNumber concentration of cloud water droplet (m−3)3.0 × 108
    P13WSM6dimaxMaximum value for cloud ice diameter (m)5.0 × 10−4
    P14WSM6r0Critical mean droplet radius at which auto-conversion begins (m)8.0 × 10−6
    P15WSM6qs0Threshold amount for aggregation to occur (kg kg−1)6.0 × 10−4
    DownLoad: CSV

    Table 3.  The iteration of CNOP-P computation.

    IterationSpectral projection gradientCost functionSensitivity ranking corresponding to CNOP-P
    01.192−923831.63410 13 3 11 8 5 12 4 9 6 2 15 7 14 1
    17.614 × 10−1−7120544.4511 5 3 2 14 6 11 10 12 13 8 9 15 7 4
    22.052 × 10−2−9873203.6701 5 13 6 2 14 10 12 11 8 3 7 4 9 15
    36.479 × 10−4−9735736.6031 13 5 7 10 2 12 4 3 11 9 15 6 14 8
    44.630 × 10−5−9745800.4791 13 5 7 10 3 2 4 9 12 15 11 6 8 14
    54.046 × 10−4−9856915.2101 13 5 7 10 3 2 4 9 12 15 11 6 8 14
    62.211 × 10−4−9615821.4731 13 7 5 10 3 4 2 12 9 15 11 8 14 6
    72.256 × 10−5−9657034.9761 13 7 5 10 3 2 4 12 9 15 11 8 6 14
    82.773 × 10−5−9727943.3381 13 7 5 10 3 2 4 12 9 15 11 8 6 14
    93.978 × 10−5−9758126.3461 13 7 5 3 10 4 2 12 9 15 11 8 6 14
    101.212 × 10−5−9686227.1871 13 7 5 3 10 4 2 12 9 15 11 8 14 6
    111.034 × 10−5−9658981.2411 13 7 5 3 10 4 2 12 9 15 11 8 14 6
    DownLoad: CSV

    Table 4.  Sensitive parameters and their ranges.

    IndexParameterMinDefaultMax
    P1rlam30150450
    P2brcr0.1250.51
    P3pfac123
    P4cfac3.97.815.6
    P5avtr420841.91263
    P7avts511.7218
    P10deng200500900
    P13dimax2 × 10−45 × 10−48 × 10−4
    DownLoad: CSV

    Table 5.  The iteration of CNOP-P computation using five cases under strong synoptic forcing as initial conditions.

    IterationSpectral projection gradientCost functionSensitivity ranking corresponding to CNOP-P
    01.216−1923239.88010 13 3 11 8 5 12 4 9 6 2 15 7 14 1
    13.401 × 10−1−12148995.401 3 2 5 13 7 4 9 11 14 6 8 15 12 10
    28.955 × 10−3−12753984.431 5 3 14 12 13 2 4 7 9 11 8 10 6 15
    33.236 × 10−4−12522691.231 5 13 7 2 3 14 12 11 9 4 15 10 6 8
    41.971 × 10−4−12598208.521 13 5 7 2 3 14 12 11 4 9 10 15 6 8
    51.201 × 10−4−12565818.291 13 7 5 2 3 14 12 11 4 9 10 15 8 6
    68.993 × 10−5−12602431.251 13 7 5 2 3 14 12 11 9 4 10 15 6 8
    76.820 × 10−5−12546727.081 13 7 5 2 3 14 12 11 9 4 10 15 6 8
    82.925 × 10−5−12539894.941 13 7 5 2 3 14 12 11 9 4 10 15 6 8
    91.211 × 10−4−12602381.101 13 7 5 2 3 14 12 11 9 4 10 15 6 8
    104.626 × 10−5−12584054.481 13 7 5 2 3 14 12 11 9 4 10 15 6 8
    112.772 × 10−5−12547402.311 13 5 7 2 3 14 12 11 9 4 10 15 6 8
    DownLoad: CSV
  • Baker, L. H., A. C. Rudd, S. Migliorini, and R. N. Bannister, 2014: Representation of model error in a convective-scale ensemble prediction system. Nonlinear Processes in Geophysics, 21, 19−39, https://doi.org/10.5194/npg-21-19-2014.
    Baldauf, M., A. Seifert, J. Förstner, D. Majewski, M. Raschendorfer, and T. Reinhardt, 2011: Operational convective-scale numerical weather prediction with the COSMO model: Description and sensitivities. Mon. Wea. Rev., 139, 3887−3905, https://doi.org/10.1175/MWR-D-10-05013.1.
    Berner, J., G. J. Shutts, M. Leutbecher, and T. N. Palmer, 2009: A spectral stochastic kinetic energy backscatter scheme and its impact on flow-dependent predictability in the ECMWF ensemble prediction system. J. Atmos. Sci., 66, 603−626, https://doi.org/10.1175/2008JAS2677.1.
    Birgin, E. G., J. M. Martínez, and M. Raydan, 2001: Algorithm 813: SPG-software for convex-constrained optimization. ACM Transactions on Mathematical Software, 27, 340−349, https://doi.org/10.1145/502800.502803.
    Bouttier, F., B. Vié, O. Nuissier, and L. Raynaud, 2012: Impact of stochastic physics in a convection-permitting ensemble. Mon. Wea. Rev., 140, 3706−3721, https://doi.org/10.1175/MWR-D-12-00031.1.
    Bowler, N. E., A. Arribas, K. R. Mylne, K. B. Robertson, and S. E. Beare, 2008: The MOGREPS short-range ensemble prediction system. Quart. J. Roy. Meteorol. Soc., 134, 703−722, https://doi.org/10.1002/qj.234.
    Buizza, R., M. Milleer, and T. N. Palmer, 1999: Stochastic representation of model uncertainties in the ECMWF ensemble prediction system. Quart. J. Roy. Meteorol. Soc., 125, 2887−2908, https://doi.org/10.1002/qj.49712556006.
    Charron, M., G. Pellerin, L. Spacek, P. L. Houtekamer, N. Gagnon, H. L. Mitchell, and L. Michelin, 2010: Toward random sampling of model error in the Canadian ensemble prediction system. Mon. Wea. Rev., 138, 1877−1901, https://doi.org/10.1175/2009MWR3187.1.
    Chen, F., and J. Dudhia, 2001: Coupling an advanced land surface-hydrology model with the Penn State-NCAR MM5 modeling system. Part I: Model implementation and sensitivity. Mon. Wea. Rev., 129, 569−585, https://doi.org/10.1175/1520-0493(2001)129<0569:CAALSH>2.0.CO;2.
    Christensen, H. M., S.-J. Lock, I. M. Moroz, and T. N. Palmer, 2017: Introducing independent patterns into the Stochastically Perturbed Parametrization Tendencies (SPPT) scheme. Quart. J. Roy. Meteorol. Soc., 143, 2168−2181, https://doi.org/10.1002/qj.3075.
    Clark, A. J., and Coauthors, 2012: An overview of the 2010 hazardous weather testbed experimental forecast program spring experiment. Bull. Amer. Meteorol. Soc., 93, 55−74, https://doi.org/10.1175/BAMS-D-11-00040.1.
    Clark, P., N. Roberts, H. Lean, S. P. Ballard, and C. Charlton-Perez, 2016: Convection-permitting models: A step-change in rainfall forecasting. Meteorological Applications, 23, 165−181, https://doi.org/10.1002/met.1538.
    Duan, W. S., and R. Zhang, 2010: Is model parameter error related to a significant spring predictability barrier for El Niño events? Results from a theoretical model Adv. Atmos. Sci., 27, 1003−1013, https://doi.org/10.1007/s00376-009-9166-4.
    Duan, W. S., and Z. H. Huo, 2016: An approach to generating mutually independent initial perturbations for ensemble forecasts: Orthogonal conditional nonlinear optimal perturbations. J. Atmos. Sci., 73, 997−1014, https://doi.org/10.1175/JAS-D-15-0138.1.
    Dudhia, J., 1989: Numerical study of convection observed during the winter monsoon experiment using a mesoscale two-dimensional model. J. Atmos. Sci., 46, 3077−3107, https://doi.org/10.1175/1520-0469(1989)046<3077:NSOCOD>2.0.CO;2.
    Fresnay, S., A. Hally, C. Garnaud, E. Richard, and D. Lambert, 2012: Heavy precipitation events in the Mediterranean: Sensitivity to cloud physics parameterisation uncertainties. Natural Hazards and Earth System Sciences, 12, 2671−2688, https://doi.org/10.5194/nhess-12-2671-2012.
    Gebhardt, C., S. E. Theis, M. Paulat, and Z. B. Bouallègue, 2011: Uncertainties in COSMO-DE precipitation forecasts introduced by model perturbations and variation of lateral boundaries. Atmospheric Research, 100, 168−177, https://doi.org/10.1016/j.atmosres.2010.12.008.
    Gerard, L., 2007: An integrated package for subgrid convection, clouds and precipitation compatible with meso-gamma scales. Quart. J. Roy. Meteorol. Soc., 133, 711−730, https://doi.org/10.1002/qj.58.
    Hacker, J. P., and Coauthors, 2011: The U.S. Air Force Weather Agency’s mesoscale ensemble: Scientific description and performance results. Tellus A, 63, 625−641, https://doi.org/10.1111/j.1600-0870.2010.00497.x.
    Hagelin, S., J. Son, R. Swinbank, A. McCabe, N. Roberts, and W. Tennant, 2017: The met office convective-scale ensemble, MOGREPS-UK. Quart. J. Roy. Meteorol. Soc., 143, 2846−2861, https://doi.org/10.1002/qj.3135.
    Hohenegger, C., and C. Schar, 2007: Atmospheric predictability at synoptic versus cloud-resolving scales. Bull. Amer. Meteorol. Soc., 88, 1783−1794, https://doi.org/10.1175/BAMS-88-11-1783.
    Hong, S. Y., and H. L. Pan, 1996: Nonlocal boundary layer vertical diffusion in a medium-range forecast model. Mon. Wea. Rev., 124, 2322−2339, https://doi.org/10.1175/1520-0493(1996)124<2322:NBLVDI>2.0.CO;2.
    Hong, S. Y., and J. O. J. Lim, 2006: The WRF single-moment 6-class microphysics scheme (WSM6). Journal of the Korean Meteorological Society, 42, 129−151.
    Honnert, R., F. Couvreux, V. Masson, and D. Lancz, 2016: Sampling the structure of convective turbulence and implications for grey-zone parametrizations. Bound.-Layer Meteorol., 160, 133−156, https://doi.org/10.1007/s10546-016-0130-4.
    Hou, D. C., Z. Toth, and Y. J. Zhu, 2006: A stochastic parameterization scheme within NCEP global ensemble forecast system. Extended abstract, the 18th AMS Conf. on Probability and Statistics in the Atmospheric Sciences, Atlanta, Georgia, Amer. Meteor. Soc.
    Hou, D. C., Z. Toth, Y. J. Zhu, and W. Y. Yang, 2008: Impact of a stochastic perturbation scheme on global ensemble forecast. Extended abstract, the 19th AMS Conf. on Probability and Statistics, New Orleans, Louisiana, Amer. Meteor. Soc.
    Iversen, T., A. Deckmyn, C. Santos, K. Sattler, J. B. Bremnes, H. Feddersen, and I.-L. Frogner, 2011: Evaluation of ‘GLAMEPS’-a proposed multimodel EPS for short range forecasting. Tellus A, 63, 513−530, https://doi.org/10.1111/j.1600-0870.2010.00507.x.
    Johnson, A., X. G. Wang, and M. Xue, 2014: Multiscale characteristics and evolution of perturbations for warm season convection-allowing precipitation forecasts: Dependence on background flow and method of perturbation. Mon. Wea. Rev., 142, 1053−1073, https://doi.org/10.1175/MWR-D-13-00204.1.
    Leutbecher, M., and Coauthors, 2017: Stochastic representations of model uncertainties at ECMWF: State of the art and future vision. Quart. J. Roy. Meteorol. Soc., 143, 2315−2339, https://doi.org/10.1002/qj.3094.
    Mascaro, G., E. R. Vivoni, and R. Deidda, 2010: Implications of ensemble quantitative precipitation forecast errors on distributed streamflow forecasting. Journal of Hydrometeorology, 11, 69−86, https://doi.org/10.1175/2009JHM1144.1.
    McCabe, A., R. Swinbank, W. Tennant, and A. Lock, 2016: Representing model uncertainty in the Met Office convection-permitting ensemble prediction system and its impact on fog forecasting. Quart. J. Roy. Meteorol. Soc., 142, 2897−2910, https://doi.org/10.1002/qj.2876.
    Melhauser, C., F. Q Zhang, Y. H. Weng, Y. Jin, H. Jin, and Q. Y. Zhao, 2017: A multiple-model convection-permitting ensemble examination of the probabilistic prediction of tropical cyclones: Hurricanes Sandy (2012) and Edouard (2014). Wea. Forecasting, 32, 665−688, https://doi.org/10.1175/WAF-D-16-0082.1.
    Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res.: Atmos., 102, 16 663−16 682, https://doi.org/10.1029/97JD00237.
    Mu, M., W. S. Duan, and B. Wang, 2003: Conditional nonlinear optimal perturbation and its applications. Nonlinear Processes in Geophysics, 10, 493−501, https://doi.org/10.5194/npg-10-493-2003.
    Mu, M., W. S. Duan, Q. Wang, and R. Zhang, 2010: An extension of conditional nonlinear optimal perturbation approach and its applications. Nonlinear Processes in Geophysics, 17, 211−220, https://doi.org/10.5194/npg-17-211-2010.
    Ollinaho, P., and Coauthors, 2017: Towards process-level representation of model uncertainties: Stochastically perturbed parametrizations in the ECMWF ensemble. Quart. J. Roy. Meteorol. Soc., 143, 408−422, https://doi.org/10.1002/qj.2931.
    Peralta, C., Z. Ben Bouallègue, S. E. Theis, C. Gebhardt, and M. Buchhold, 2012: Accounting for initial condition uncertainties in COSMO-DE-EPS. J. Geophys. Res.: Atmos., 117, D07108, https://doi.org/10.1029/2011JD016581.
    Qin, X. H., W. S. Duan, and H. Xu, 2020: Sensitivity to tendency perturbations of tropical cyclone short-range intensity forecasts generated by WRF. Adv. Atmos. Sci., 37, 291−306, https://doi.org/10.1007/s00376-019-9187-6.
    Qu, Y. M., W. S. Lu, R. H. Cai, Y. Yang, D. M. Jiang, and L. P. Liu, 2010: Design and experiment of GRAPES-Meso cloud analysis system. Meteorological Monthly, 36, 37−45, https://doi.org/10.7519/j.issn.1000-0526.2010.10.006. (in Chinese with English abstract)
    Romine, G. S., C. S. Schwartz, J. Berner, K. R. Fossell, C. Snyder, J. L. Anderson, and M. L. Weisman, 2014: Representing forecast error in a convection-permitting ensemble system. Mon. Wea. Rev., 142, 4519−4541, https://doi.org/10.1175/MWR-D-14-00100.1.
    Schwartz, C. S., G. S. Romine, R. A. Sobash, K. R. Fossell, and M. L. Weisman, 2015: NCAR’s experimental real-time convection-allowing ensemble prediction system. Wea. Forecasting, 30, 1645−1654, https://doi.org/10.1175/WAF-D-15-0103.1.
    Seity, Y., P. Brousseau, S. Malardel, G. Hello, P. Bénard, F. Bouttier, C. Lac, and V. Masson, 2011: The AROME-France convective-scale operational model. Mon. Wea. Rev., 139, 976−991, https://doi.org/10.1175/2010MWR3425.1.
    Shutts, G., 2005: A kinetic energy backscatter algorithm for use in ensemble prediction systems. Quart. J. Roy. Meteorol. Soc., 131, 3079−3102, https://doi.org/10.1256/qj.04.106.
    Tao, L. J., and W. S. Duan, 2019: Using a nonlinear forcing singular vector approach to reduce model error effects in ENSO forecasting. Wea. Forecasting, 34, 1321−1342, https://doi.org/10.1175/WAF-D-19-0050.1.
    Wang, B., and X. W. Tan, 2010: Conditional nonlinear optimal perturbations: Adjoint-free calculation method and preliminary test. Mon. Wea. Rev., 138, 1043−1049, https://doi.org/10.1175/2009MWR3022.1.
    Weyn, J. A., and D. R. Durran, 2018: Ensemble spread grows more rapidly in higher-resolution simulations of deep convection. J. Atmos. Sci., 75, 3331−3345, https://doi.org/10.1175/JAS-D-17-0332.1.
    Yin, X. D., J. J. Liu, and B. Wang, 2015: Nonlinear ensemble parameter perturbation for climate models. J. Climate, 28, 1112−1125, https://doi.org/10.1175/JCLI-D-14-00244.1.
    Yu, Y. S., M. Mu, and W. S. Duan, 2012: Does model parameter error cause a significant “spring predictability barrier” for El Niño events in the Zebiak-Cane model? J. Climate, 25, 1263−1277, https://doi.org/10.1175/2011JCLI4022.1.
    Yuan, Y., X. L. Li, J. Chen, and Y. Xia, 2016: Stochastic parameterization toward model uncertainty for the GRAPES mesoscale ensemble prediction system. Meteorological Monthly, 42, 1161−1175, https://doi.org/10.7519/j.issn.1000-0526.2016.10.001. (in Chinese with English abstract)
    Zhang, X. W., W. Y. Tang, L. Q. Fan, J. Sheng, X. W. Cai, T. Zhang, and X. L. Zhang, 2018: Evaluation on the application of GRAPES-CR in forecasting severe convective weather. CMA, Special Funds for GRAPES, No. 400288, 75 pp. (in Chinese)
    Zhu, L. J., J. D. Gong, L. P. Huang, D. H. Chen, Y. Jiang, and L. T. Deng, 2017: Three-dimensional cloud initial field created and applied to GRAPES numerical weather prediction nowcasting. Journal of Applied Meteorological Science, 28, 38−51, https://doi.org/10.11898/1001-7313.20170104. (in Chinese with English abstract)
  • [1] Zhizhen XU, Jing CHEN, Mu MU, Guokun DAI, Yanan MA, 2022: A Nonlinear Representation of Model Uncertainty in a Convective-Scale Ensemble Prediction System, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1432-1450.  doi: 10.1007/s00376-022-1341-x
    [2] Zhizhen XU, Jing CHEN, Zheng JIN, Hongqi LI, Fajing CHEN, 2020: Representing Model Uncertainty by Multi-Stochastic Physics Approaches in the GRAPES Ensemble, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 328-346.  doi: 10.1007/s00376-020-9171-1
    [3] SUN Guodong, MU Mu, 2011: Response of a Grassland Ecosystem to Climate Change in a Theoretical Model, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 1266-1278.  doi: 10.1007/s00376-011-0169-6
    [4] WANG Bo, and HUO Zhenhua, 2013: Extended application of the conditional nonlinear optimal parameter perturbation method in the Common Land Model, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1213-1223.  doi: 10.1007/s00376-012-2025-8
    [5] Feifan ZHOU, Wansuo DUAN, He ZHANG, Munehiko YAMAGUCHI, 2018: Possible Sources of Forecast Errors Generated by the Global/Regional Assimilation and Prediction System for Landfalling Tropical Cyclones. Part II: Model Uncertainty, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1277-1290.  doi: 10.1007/s00376-018-7095-9
    [6] LIU Liping, ZHUANG Wei, ZHANG Pengfei, MU Rong, 2010: Convective Scale Structure and Evolution of a Squall Line Observed by C-Band Dual Doppler Radar in an Arid Region of Northwestern China, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1099-1109.  doi: 10.1007/s00376-009-8217-1
    [7] Yuejian ZHU, 2005: Ensemble Forecast: A New Approach to Uncertainty and Predictability, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 781-788.  doi: 10.1007/BF02918678
    [8] Jiangshan ZHU, Fanyou KONG, Xiao-Ming HU, Yan GUO, Lingkun RAN, Hengchi LEI, 2018: Impact of Soil Moisture Uncertainty on Summertime Short-range Ensemble Forecasts, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 839-852.  doi: 10.1007/s00376-017-7107-1
    [9] Peng Jiayi, Wu Rongsheng, Wang Yuan, 2002: Initiation Mechanism of Meso-β Scale Convective Systems, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 870-884.  doi: 10.1007/s00376-002-0052-6
    [10] Youlong XIA, Zong-Liang YANG, Paul L. STOFFA, Mrinal K. SEN, 2005: Optimal Parameter and Uncertainty Estimation of a Land Surface Model: Sensitivity to Parameter Ranges and Model Complexities, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 142-157.  doi: 10.1007/BF02930878
    [11] XU Hui, DUAN Wansuo, 2008: What Kind of Initial Errors Cause the Severest Prediction Uncertainty of El Nino in Zebiak-Cane Model, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 577-584.  doi: 10.1007/s00376-008-0577-4
    [12] Fan Beifen, Ye Jiadong, William R. Cotton, Gregory J. Tripoli, 1990: Numerical Simulation of Microphysics in Meso-β-Scale Convective Cloud System Associated with a Mesoscale Convective Complex, ADVANCES IN ATMOSPHERIC SCIENCES, 7, 154-170.  doi: 10.1007/BF02919153
    [13] Zihan YIN, Panxi DAI, Ji NIE, 2021: A Two-plume Convective Model for Precipitation Extremes, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 957-965.  doi: 10.1007/s00376-021-0404-8
    [14] Shibo GAO, Haiqiu YU, Chuanyou REN, Limin LIU, Jinzhong MIN, 2021: Assimilation of Doppler Radar Data with an Ensemble 3DEnVar Approach to Improve Convective Forecasting, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 132-146.  doi: 10.1007/s00376-020-0081-z
    [15] Xiaoran ZHUANG, Jinzhong MIN, Liu ZHANG, Shizhang WANG, Naigeng WU, Haonan ZHU, 2020: Insights into Convective-scale Predictability in East China: Error Growth Dynamics and Associated Impact on Precipitation of Warm-Season Convective Events, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 893-911.  doi: 10.1007/s00376-020-9269-5
    [16] ZHANG Hanbin, CHEN Jing, ZHI Xiefei, WANG Yi, WANG Yanan, 2015: Study on Multi-Scale Blending Initial Condition Perturbations for a Regional Ensemble Prediction System, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1143-1155.  doi: 10.1007/s00376-015-4232-6
    [17] Guanshun ZHANG, Jiangyu MAO, Wei HUA, Xiaofei WU, Ruizao SUN, Ziyu YAN, Yimin LIU, Guoxiong WU, 2023: Synergistic Effect of the Planetary-scale Disturbance, Typhoon and Meso-β-scale Convective Vortex on the Extremely Intense Rainstorm on 20 July 2021 in Zhengzhou, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 428-446.  doi: 10.1007/s00376-022-2189-9
    [18] WAN Liying, ZHU Jiang, WANG Hui, YAN Changxiang, Laurent BERTINO, 2009: A ``Dressed" Ensemble Kalman Filter Using the Hybrid Coordinate Ocean Model in the Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 1042-1052.  doi: 10.1007/s00376-009-7208-6
    [19] Qian ZOU, Quanjia ZHONG, Jiangyu MAO, Ruiqiang DING, Deyu LU, Jianping LI, Xuan LI, 2023: Impact of Perturbation Schemes on the Ensemble Prediction in a Coupled Lorenz Model, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 501-513.  doi: 10.1007/s00376-022-1376-z
    [20] Xiaqiong ZHOU, Johnny C. L. CHEN, 2006: Ensemble Forecasting of Tropical Cyclone Motion Using a Baroclinic Model, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 342-354.  doi: 10.1007/s00376-006-0342-5

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 26 November 2019
Manuscript revised: 28 April 2020
Manuscript accepted: 29 April 2020
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

Model Uncertainty Representation for a Convection-Allowing Ensemble Prediction System Based on CNOP-P

    Corresponding author: Xueshun SHEN, shenxs@cma.gov.cn
  • 1. Asset Operation Centre, China Meteorological Administration, Beijing 100081, China
  • 2. State Key Laboratory of Severe Weather, Chinese Academy of Meteorological Sciences, Beijing 100081, China
  • 3. Numerical Weather Prediction Center/National Meteorological Center, China Meteorological Administration, Beijing 100081, China
  • 4. Institute of Atmosphere Physics, Chinese Academy of Sciences, Beijing 100029, China
  • 5. Center for Earth System Science, Tsinghua University, Beijing 100084, China

Abstract: Formulating model uncertainties for a convection-allowing ensemble prediction system (CAEPS) is a much more challenging problem compared to well-utilized approaches in synoptic weather forecasting. A new approach is proposed and tested through assuming that the model uncertainty should reasonably describe the fast nonlinear error growth of the convection-allowing model, due to the fast developing character and strong nonlinearity of convective events. The Conditional Nonlinear Optimal Perturbation related to Parameters (CNOP-P) is applied in this study. Also, an ensemble approach is adopted to solve the CNOP-P problem. By using five locally developed strong convective events that occurred in pre-rainy season of South China, the most sensitive parameters were detected based on CNOP-P, which resulted in the maximum variations in precipitation. A formulation of model uncertainty is designed by adding stochastic perturbations into these sensitive parameters. Through comparison ensemble experiments by using all the 13 heavy rainfall cases that occurred in the flood season of South China in 2017, the advantages of the CNOP-P-based method are examined and verified by comparing with the well-utilized stochastically perturbed physics tendencies (SPPT) scheme. The results indicate that the CNOP-P-based method has potential in improving the under-dispersive problem of the current CAEPS.

摘要: 相对于全球集合预报,对流尺度集合预报(CAEPS)中有关模式不确定性的研究缺乏系统性和理论基础,成为目前研究的热点和难点。针对强对流天气发展迅速、非线性强等特点,本文充分考虑了对流尺度模式误差非线性快速增长特征,提出基于条件非线性最优参数扰动(CNOP-P)方法研究CAEPS模式不确定性问题的新思路。同时,在CNOP-P方法的计算中,本文采用了集合求解算法,不依赖于切线伴随模式。我们首先选取了2017年中国华南前汛期5个典型暖区暴雨个例求解CNOP-P,得到对累积降水预报产生最大影响的参数扰动集合,挑选敏感物理参数;然后,通过随机扰动敏感参数来描述模式不确定性,构造基于CNOP-P的模式扰动方案,并将该方案与目前广泛应用的随机物理参数化倾向扰动方案进行对比,展开对流尺度集合预报关于模式不确定性的敏感性试验。2017年华南地区13个强对流过程的检验结果显示:基于CNOP-P的模式扰动方案对于提高对流尺度集合离散度方面具有应用潜力。

    • Convection-allowing models (CAMs) have been receiving increasing attention in recent years owing to their potential in describing the detailed characteristics of mesoscale systems (Baldauf et al., 2011; Seity et al., 2011; Clark et al., 2016). Usually, the grid size of a CAM is 1–4 km, but this range of grid spacing is regarded as the “physics grey zone” (Gerard, 2007; Honnert et al., 2016). Within this grid size range, the physical processes are partly resolved, and some remain parameterized. It is still not fine enough to resolve convective cells, subgrid-scale orographic gravity wave drag, as well as turbulence diffusion. These bring uncertainties to the model, which exist in association with physical parameterizations and parameters in physical schemes. In addition, interactions between dynamics and physics or among different physical schemes in CAMs may become even more complicated owing to the increase in number of degrees of freedom in modeling the atmospheric phenomena at convection-allowing scales. Hohenegger and Schar (2007) pointed out that there is a much stronger nonlinear interaction between dynamics and physical processes owing to the rapidly developing character of local strong events. Thus, CAMs have their limitations in predicting the initiation and development of strong events only by using a single deterministic forecast, including their location, intensity and duration.

      Due to its effective probabilistic forecast information, the ensemble approach is believed necessary for the purpose of fully exploiting the potential of CAMs. Indeed, convection-allowing ensemble prediction systems (CAEPSs) play an important role in local high-resolution numerical weather prediction. Several leading forecast centers have started to run a CAEPS operationally in recent years, including COSMO-DE-EPS (Peralta et al., 2012), MOGREPS-UK (Hagelin et al., 2017), and AROME-EPS (Bouttier et al., 2012). CAEPSs have also been developed as part of NOAA’s Hazardous Weather Spring Experiment, such as SSEF (Clark et al., 2012) and NCAR’s EnKF-based EPS (Schwartz et al., 2015). See Table A1 for further details.

      CenterModelHorizontal resolution
      (km)
      Forecast
      range (h)
      MembersInitial perturbationsLateral
      boundary
      conditions
      Model
      uncertainty
      DWDCOSMO-DE-EPS2.82720Downscaled and
      recentered perturbations
      from IFS, GFS,
      ICON, GSM
      Downscaled
      from IFS,
      GFS,ICON,
      GSM
      Multi-parameter
      Met OfficeMOGREPS-UK2.25412Downscaled and
      recentered
      perturbations from
      MOGREPS-G
      Downscaled from MOGREPS-GRP
      Météo-FranceAROME-EPS2.54512Downscaled and recentered perturbations from PEARPDownscaled from PEARPSPPT
      RHMCCOSMO-Ru2-EPS2.24810Downscaled from COSMO-S14-EPSDownscaled from COSMO-S14-EPS
      KMALENS34512Downscaled from EFSGDownscaled from EFSGRP
      CAPSSSEF36020Downscaled from SREFDownscaled from SREFMulti-
      modelmulti-
      physics
      NCAREnKF-based EPS34810EAKF analysisDownscaled from GFS+ random draws

      Table A1.  Summary of operational convection-allowing ensemble prediction systems.

      However, the above CAEPSs have a common problem, known as under-dispersion. Typically, as the forecasts proceed, the ensemble spread among the ensemble members becomes too small to ensure that the observed weather pattern could be contained within the ensemble (Mascaro et al., 2010; Weyn and Durran, 2018).

      Previous studies have pointed out that CAEPSs could benefit through improving the representation of model uncertainties (Johnson et al., 2014; Romine et al., 2014). To date, several approaches have been proposed in representing model uncertainties, including multi-model, multi-physics, multi-parameter, and stochastic physics. The multi-model approach considers the uncertainty of the dynamical core and parameterization by using different models (Iversen et al., 2011; Melhauser et al., 2017). Multi-physics and multi-parameter methods are used to describe the uncertainties embedded in the physical parameterizations by changing schemes (Charron et al., 2010; Hacker et al., 2011) or their empirical parameters (Gebhardt et al., 2011). Stochastic physics is more theoretical compared with the above approaches, in which the model uncertainties are described by introducing random perturbations into the physical processes, such as the well-utilized Stochastic Perturbed Parameterization Tendencies (SPPT) scheme (Buizza et al., 1999; Christensen et al., 2017), the Stochastic Total Tendency Perturbation scheme (Hou et al., 2006, 2008), and the Random Parameters/Stochastically Perturbed Parameterizations scheme (Bowler et al., 2008; McCabe et al., 2016; Ollinaho et al., 2017). In addition, considering the kinetic energy dissipation at the truncation scale, the Stochastic Kinetic Energy Backscatter scheme (Shutts, 2005; Berner et al., 2009) is another stochastic method that expresses the model structural uncertainty. Recently, a new approach based on the nonlinear forcing singular vector has been proposed to represent the combined effect of different sources of model uncertainties (Tao and Duan, 2019; Qin et al., 2020).

      The above-mentioned model perturbation approaches have been widely utilized in global ensemble forecasts for a long time, and recently applied in CAEPSs for forecasting severe convective weather, tropical cyclone intensity, and so on. However, it is found that the above model perturbation approaches have limited effects in dealing with the under-dispersion problem of CAEPSs (Fresnay et al., 2012; Baker et al., 2014). Such limited effects may come from deficiencies in formulating the model uncertainties at the convection-allowing scales. Compared with global ensemble prediction systems, the representation of model uncertainty in CAEPSs is thought be lacking in systematic research and a theoretical basis, therefore making it an important issue worthy of further study.

      As known, the main target of a CAEPS is to capture the probability of occurrence of rapidly evolving small- to mesoscale strong events. From the ensemble forecast viewpoint, the rapidly nonlinear growing perturbations should be described in the CAEPS’s formulations both for initial errors and model uncertainties. Mu et al. (2003) proposed the concept of conditional nonlinear optimal perturbation (CNOP). The CNOP represents the maximum nonlinear evolution of initial perturbations satisfying constraint conditions at the prediction time, which is thought to be greatly important for the predictability (Wang and Tan, 2010). Since Mu et al. (2003), CNOP has been applied in studying the various predictability issues, such as ENSO predictability, target observations, and so on. Duan and Huo (2016) also proposed the approach of orthogonal CNOPs, which extended the CNOP to ensemble forecasting. Furthermore, considering the predictability caused by the model error, Mu et al. (2010) extended the CNOP approach to search for the optimal model parameter perturbations, which resulted in the largest departure from a given reference state at a prediction time with physical constraints. This approach is referred to as CNOP-P, which has been extensively employed in understanding the climate model sensitivity to model parameters and investigating the impact of parameter errors on ENSO predictability, amongst other problems (Duan and Zhang, 2010; Yu et al., 2012; Yin et al., 2015).

      As discussed above, a good CAEPS should reasonably describe the rapidly growing nonlinear errors or maximum growing perturbations at a short prediction time either due to initial errors or model uncertainties. In this sense, the CNOP-P approach may have potential to detect and describe the maximum perturbation growth caused by the model parameter uncertainties for CAEPSs. Compared with current formulations of model uncertainties in CAEPSs, the CNOP-P approach is expected to be more objective. However, the original computation method of CNOP-P requires integration of an adjoint model to calculate the cost function gradient. This limits the practical use of CNOP-P in operational applications. Wang and Tan (2010) proposed a pragmatic method to determine the CNOP by using an ensemble approach, which is easier to implement and more time-saving than the adjoint-based method. Thus, our research seeks to exploit a novel method to formulate the model uncertainties for CAEPSs by using the CNOP-P concept and the approach to determining the CNOP proposed by Wang and Tan (2010).

      The rest of the paper is organized as follows. Section 2 briefly describes the CNOP-P approach utilized in this paper and the experimental design. The results of sensitivity tests and the impact of the CNOP-P approach on the ensemble forecasts are reported in section 3. Finally, conclusions and future work are summarized in section 4.

    2.   Methods
    • According to Mu et al. (2010), in a nonlinear prediction model M, ${{y}} = {M_\tau }\left({{P}} \right)\left({{{{U}}_0}} \right)$ can denote a prediction from the initial time 0 to τ with the parameter vector P and the initial value U0. The prediction increment y' at the prediction time τ is defined as the departure from the reference state as a result of parameter perturbations p':

      The optimal parameter perturbations p' should correspond to maximum the prediction increments y' with a constraint of || p'||≤σ (σ is a small positive number, given here as 1). The p' can be obtained through solving the following equation:

      In order to reduce the dependency of J(p') on initial values, Yin et al. (2015) introduced the ensemble-averaged type of cost function, which is constructed by using N short-term time series predictions with independent initial values U0,k (k=1,2,…,N) as follows:

      The above cost function can be rewritten into a minimization problem for convenience of solution:

      The key to this minimization problem is the calculation of its gradient with respect to the parameter perturbation, which can be expressed as:

      As derived by Wang and Tan (2010) and Yin et al. (2015), the solution of Eq. (5) can be obtained by the following ensemble approach.

      Firstly, for each initial value U0,k, n parameter perturbation samples p'i (i=1,2,…,n) are used to generate n corresponding prediction increment samples y'i,k (i=1,2,…,n; k=1,2,…,N). The parameter perturbations and prediction increment matrices are denoted as P' and Y'k, respectively:

      The second step is to obtain a series of orthogonal basis functions of P' by using singular value decomposition:

      where, Ap is the left singular vector of P', which is taken as the orthogonal basis function.

      The third step is to estimate the gradient of the cost function based on ensemble projection. With Ap as the projection matrix, each parameter perturbation can be projected to the ensemble space:

      where α is the parameter perturbation after projection.

      According to the tangent linear relationship between p' and y'k, each prediction increment can also be projected to the same ensemble space:

      where M' is the tangent linear model of M.

      Then, a statistical model between p' and y'k can be established:

      According to the orthogonality of Ap, we can obtain ${\left({{{\partial {{{{y}}'_k}}} / {\partial {{p'}}}}} \right)^{\rm{T}}} = {{{A}}_p}{\left({{{{A}}_{y,k}}} \right)^{\rm{T}}}$. Then, the gradient of the cost function can be denoted as follows:

      The last step is to compute CNOP-P and detect the most sensitive parameters. The Spectral Projected Gradient 2 (SPG2) optimization algorithm of (Birgin et al., 2001) is used to obtain a result iteratively. To guarantee nonlinearity of the optimal solution, GRAPES-Meso (Global/Regional Assimilation and Prediction System, mesoscale version), with a horizontal grid size of 3 km, is used to calculate the cost function value. Meanwhile, the gradient of the cost function and constraint condition of parameter perturbations are required during iteration. The constraint condition is expressed as:

      There are m parameters involved in the calculation of CNOP-P. Perturbations are standardized by dividing their corresponding default values. In this way, the greater the perturbation weight of a parameter in the CNOP-P, the greater the relative sensitivity of this parameter.

    • The model utilized here is the operational GRAPES-Meso, version 4.1, with a convection-allowing resolution setup. This model solves the fully compressible nonhydrostatic equations using a semi-implicit, semi-Lagrangian integration scheme on a spherical Arakawa-C grid horizontally, and Charney–Phillips staggered grids vertically. It has 50 vertical levels up to 10 hPa. For the 3 km resolution, the 30-s time step is adopted.

      Physical parameterization schemes used here include the Rapid Radiation Transfer Model longwave radiation scheme (Mlawer et al., 1997), the Dudhia shortwave radiation scheme (Dudhia, 1989), the Monin–Obukhov surface layer scheme, the Noah land surface scheme (Chen and Dudhia, 2001), the MRF boundary layer scheme (Hong and Pan, 1996), and the WSM6 cloud microphysics scheme (Hong and Lim, 2006). The cumulus convection scheme is not applied.

      The initial conditions are generated by the GRAPES Cloud Analysis System (GCAS) (Qu et al., 2010) with a 3-km resolution, which has been tested in GRAPES-Meso_3km from 2015 to the present (Zhu et al., 2017). By using GCAS, less model spin-up time in the precipitation forecast was observed. NCEP Final Operational Global Analysis data at a spatial resolution of 1° are used as the background. Based on this background, GCAS integrates data sources from the Doppler weather radar 3D mosaic reflectivity data at a spatial resolution of 0.01°, blackbody temperature and total cloud amount data from the FY-2G geostationary satellite. The lateral boundary conditions are also provided by the NCEP Final Operational Global Analysis with a 1° horizontal resolution.

      Because CAEPSs are mainly designed for obtaining the forecast probability of convection, a region with frequent occurrence of locally developed convection is preferred as the modeling area. As is known, most of the precipitation during the pre-summer rainy season in South China is of a convective nature with mesoscale organizational characteristics. Therefore, the model domain covers South China (18.5°–27.5°N, 105.0°–118.5°E) in this study. In order to reduce the influences from the boundaries on the CNOP-P computation, a smaller area (20°–26°N, 106.5°–117°E) is selected for calculating the CNOP-P, as shown in Fig. 1.

      Figure 1.  Model forecast domain (D1: 18.5°–27.5°N, 105.0°–118.5°E) and computation domain for CNOP-P (D2: 20°–26°N, 106.5°–117°E).

    • For the purpose of better isolating the physical processes influencing the initiation and development of convection, five locally developed rainstorm cases in 2017 that occurred in South China are chosen for constructing the ensemble to compute the CNOP-P, i.e., N=5 in Eq. (3). These five cases occurred under weak synoptic forcing. Here, the 12-h cumulative precipitation is used to calculate the cost function for detecting the most sensitive parameters, which means that CNOP-P could lead to the maximum variations in precipitation during the first 12 h of model integration. Details of the five cases are given as Cases 1–5 in Table 1. Then, a supplementary experiment is carried out to confirm the robustness of the above parameter detection. Five heavy rainfall cases under strong synoptic forcing are selected to compute the CNOP-P, which are given as Cases 6–10 in Table 1. It is worth mentioning that all the forecasts in Table 1 are initialized about 6 h before the onset of rainfall events. These cases are classified as well as documented by the forecasters in the Central Meteorological Bureau (Zhang et al., 2018).

      CaseInitial forecast
      time (UTC)
      Period of heavy rainfall (UTC)Weather type Trigger condition
      11200 6 May2100 6 May–0000 7 MayRainstormConvective systems in the
      warm sector
      20000 9 May0600 9 May–0900 9 MayRainstormConvective systems in the
      warm sector
      30000 15 Jun0900 15 Jun–1200 15 JunWide-area rainstormConvective systems in the
      warm sector
      40000 20 Jun0900 20 Jun–1200 20 JunWide-area rainstormConvective systems in the
      warm sector
      50000 26 Jun0900 26 Jun–1200 26 JunRainstorm, thunderstorm galeConvective systems in the
      warm sector
      60000 8 May0900 8 May–1200 8 MayWide-area rainstorm,
      thunderstorm gale
      Cold front
      71200 14 May1800 14 May–2100 14 MayWide-area rainstormCold front
      80000 23 May0900 23 May–1200 23 MayWide-Wide-area rainstorm,
      thunderstorm gale gale
      Cold front
      90000 6 Jun0600 6 Jun–0900 6 JunWide-area rainstorm,
      thunderstorm gale
      Cold front
      100000 16 Jun0700 16 Jun–1000 16 JunWide-Wide-area rainstorm,
      thunderstorm gale gale
      Cold front
      111200 20 Apr2100 20 Apr–0000 21AprRainstorm, thunderstorm gale, hailCold front
      121200 3 May1800 3 May–0000 4 MayWide-area rainstorm, thunderstorm gale, hailSquall line
      131200 23 May1700 23 May–2000 23 MayWide-area rainstormCold front
      141200 1 Jul0000 2 Jul–0300 2 JulRainstorm, thunderstorm galeConvective systems in the
      warm sector
      150000 7 Jul0700 7 Jul–1000 7 JulRainstormConvective systems in the
      warm sector
      160000 10 Jul0600 10 Jul–0900 10 JulRainstorm, thunderstorm galeConvective systems in the
      warm sector
      170000 19 Jul0600 19 Jul–0900 19 JulRainstorm, thunderstorm galeConvective systems on the edge
      of subtropical high and in
      the warm sector
      180000 21 Jul0800 21 Jul–1000 21 JulRainstorm, thunderstorm galeConvective systems on the edge
      of subtropical high and in
      the warm sectorarm sector
      190000 31 Jul0600 31 Jul–0900 31 JulWide-area rainstorm,
      thunderstorm gale, hail
      Cold front
      200000 1 Aug0600 1 Aug–0900 1 AugRainstorm, thunderstorm galeConvective systems in the
      warm sector
      210000 10 Aug0800 10 Aug–1000 10 AugRainstorm, thunderstorm galeConvective systems on the edge of subtropical high
      220000 22 Aug1200 22 Aug–1500 22 AugRainstorm, thunderstorm galeTyphoon
      230000 23 Aug0800 23 Aug–1000 23 AugRainstormTyphoon

      Table 1.  Twenty-three typical heavy rainfall cases that occurred in South China in 2017.

      Fifteen parameters as shown in Table 2 are selected from the boundary layer and microphysics, which are thought to be the most important two processes in influencing convective events. Theoretically, all the physical parameters should be selected to perturb for calculating the CNOP-P. However, due to the high computing costs, we only selected 15 parameters in this study. According to previous research on CAEPSs (Baker et al., 2014; Gebhardt et al., 2011; McCabe et al., 2016), it was believed that this selection would not lose the generality. When we generate the parameter perturbation samples, only one parameter is changed each time, and the perturbed value is ±50% of its default value.

      IndexSchemeParameterDescriptionDefault
      P1MRFrlamAsymptotic mixing length (m)150
      P2MRFbrcrCritical Richardson number0.5
      P3MRFpfacProfile shape exponent for calculating the momentum diffusion coefficient2
      P4MRFcfacCoefficient for Prandtl number at the top of the surface layer7.8
      P5WSM6avtrA constant for terminal velocity of rain841.9
      P6WSM6avtgA constant for terminal velocity of graupel330
      P7WSM6avtsA constant for terminal velocity of snow11.72
      P8WSM6n0rIntercept parameter of rain (m−4)8.0 × 106
      P9WSM6n0gIntercept parameter of graupel (m−4)4.0 × 106
      P10WSM6dengDensity of graupel (kgm−3)500
      P11WSM6peautMean collection efficiency0.55
      P12WSM6xncrNumber concentration of cloud water droplet (m−3)3.0 × 108
      P13WSM6dimaxMaximum value for cloud ice diameter (m)5.0 × 10−4
      P14WSM6r0Critical mean droplet radius at which auto-conversion begins (m)8.0 × 10−6
      P15WSM6qs0Threshold amount for aggregation to occur (kg kg−1)6.0 × 10−4

      Table 2.  The chosen physical parameters for computing CNOP-P.

    • By applying the CNOP-P, a group of the most sensitive parameters can be detected from the 15 parameters, which correspond to the maximum forecast error growth caused by the parameter uncertainties. Based on these most sensitive parameters, a novel formulation of model uncertainty is designed, and the ensemble experiment design is given below.

      As is known, currently existing systems typically use ensemble sizes of around 10 or more members, like MOGREPS-UK, AROME-EPS, NCAR’s EnKF-based EPS, etc. Due to the available computer resources, a 10-member ensemble run was designed in this study, which included one control member and nine perturbed members. For the purpose of investigating the advantages of this new approach in describing model uncertainty, we established the ensemble prediction system by only taking into account the model uncertainty. The well-utilized SPPT scheme is adopted as the benchmark, which represents the typical model uncertainty formulation in current operational global ensemble and regional ensemble systems (Bouttier et al., 2012; Leutbecher et al., 2017). Two sets of ensemble forecasts are carried out with different model perturbation methods. The first set of ensemble forecasts uses SPPT scheme, which is called EXP1 for short. The second uses the CNOP-P-based approach, which is called EXP2 for short. Next, the details of the two model uncertainty formulations are described.

      In the SPPT scheme, a stochastic perturbation is added to the net tendency term of physical parameterizations:

      where X0 is the net tendency term, Xp is the perturbed net tendency term, and φ is a 3D random field. The scheme firstly introduces a random field generator with certain space structures and time-correlated characteristics. This generator is an expansion of spherical harmonics along the horizontal directions based on first-order autoregressive random processes. Then, the random number is multiplied into the net tendency of all the physical processes. Aimed at maintaining the numerical stability, vertical profiles are introduced to reduce the perturbations near the surface and the model top. The following parameter configurations are based on Yuan et al. (2016): the random field follows a Gaussian distribution with a mean value of 0, a standard deviation of 0.27, a perturbation range of [0.2, 1.8], and temporal decorrelation scale of 6 h. Because of time limitations, this paper does not consider the tuning of the SPPT scheme, or its reformulation, e.g., by using the iSPPT formulation (Christensen et al., 2017). It should be noted that the SPPT scheme used here has been well tuned for the GRAPES mesoscale ensemble forecast system with a 15-km resolution, and has been operationally implemented since 2014 (Yuan et al., 2016).

      For the CNOP-P-based approach, based on the detected sensitive parameters as described above, a random parameter algorithm (McCabe et al., 2016) is introduced to construct the CNOP-P-based formulation of the model uncertainty.

      By giving a physically reasonable range [Pmin, Pmax] for varying to each parameter, the detected sensitive parameters are first mapped onto the range [−1, 1]:

      where η is the mapped value, and Pmin and Pmax are the minimum and maximum values.

      By using the first-order auto-regression model, the mapped parameters are stochastically perturbed and updated at time step t. The parameters are perturbed stochastically and independently at regular intervals throughout the relations:

      where ηt is the value of mapped parameters at time t, r is the autocorrelation coefficient, and εt is the stochastic shock term. εt is formulated in terms of r and stochastic number At. r is expressed by the update time interval Δt and the characteristic time scale τp. According to McCabe et al. (2016), the update time interval is set to 5 min and the characteristic time scale is set to 7 d. Then, ηt will be mapped back to original range using Eq. (14).

    • For the purpose of systematically comparing the forecast skills using the two different approaches of model uncertainty formulation, 13 heavy rainfall cases that occurred during the flood season of South China in 2017 were selected to carry out reforecast experiments. These cases are given as Cases 11–23 in Table 1, including heavy rainfall events with strong and weak synoptic forcing.

      The near-surface variables, including 2-m specific humidity (Q2m), 10-m wind speed (V10m) and 2-m temperature (T2m), as well as hourly precipitation (R), are mainly evaluated in this study. Hourly observations from automatic weather stations are used in the verification.

      The verification metrics include the ratio of spread to root-mean-square error (RMSE), outlier frequency, and continuous ranked probability score (CRPS). In an ideal ensemble, the ratio of spread to RMSE (hereafter, S/R) should be as close to 1 as possible; that is, the spread can represent the error evolution characteristics of the system. In reality, spread is often less than the RMSE owing to the observation error. Here, observation error is not taken into account in the verification in our study:

      where (¯) denotes the mean value of all grid points, fi is the forecast of ensemble member i, n is the number of ensemble members, fmean is the ensemble mean, f0 is the control forecast, and o is the observation value.

      Outlier frequency is used to evaluate the reliability of the CAEPS. Suppose there are N valid samples and n ensemble members. After sorting the forecast of ensemble members from small to large, n+1 intervals are generated. Si denotes the occurrence frequency of observation in the ith interval. The Talagrand distribution map can be drawn from the probability distribution:

      In an ideal Talagrand distribution, there is not much difference between the probabilities of each interval, but in reality, the two ends are larger. Outlier frequency is a measure of the value at both ends of the distribution, which denotes the average probability that the observations fall outside the range of the forecasts. The ideal value is 1/(n+1).

      CRPS is a probabilistic skill score to measure the differences in cumulative distribution functions between observations and forecasts. It can be expressed as the integral of the Brier score for all possible thresholds of the meteorological variables:

      where F(x) and Fo(x) represent the cumulative distribution functions of the ensemble forecast and observation, respectively. The value for an ideal forecast of CRPS is 0. A lower CRPS indicates a more skillful forecast.

    3.   Results
    • As described in section 2.1, iteration is necessary in solving Eq. (5) for detecting the sensitive parameters. In this study, the iteration is stopped when the cost function becomes almost unchanged (< 1%), and the spectral projection gradient decreases by five to six orders of magnitude. Table 3 shows the results of the iterative process. It is found that the above criteria start being met from step 5 of iteration, and then the parameters remain stably ranked. Here, the top eight parameters are regarded as the most sensitive ones, denoted as P1, P13, P7, P5, P3, P10, P4 and P2 in Table 2. We attempt to provide a physical explanation of these eight sensitive parameters.

      IterationSpectral projection gradientCost functionSensitivity ranking corresponding to CNOP-P
      01.192−923831.63410 13 3 11 8 5 12 4 9 6 2 15 7 14 1
      17.614 × 10−1−7120544.4511 5 3 2 14 6 11 10 12 13 8 9 15 7 4
      22.052 × 10−2−9873203.6701 5 13 6 2 14 10 12 11 8 3 7 4 9 15
      36.479 × 10−4−9735736.6031 13 5 7 10 2 12 4 3 11 9 15 6 14 8
      44.630 × 10−5−9745800.4791 13 5 7 10 3 2 4 9 12 15 11 6 8 14
      54.046 × 10−4−9856915.2101 13 5 7 10 3 2 4 9 12 15 11 6 8 14
      62.211 × 10−4−9615821.4731 13 7 5 10 3 4 2 12 9 15 11 8 14 6
      72.256 × 10−5−9657034.9761 13 7 5 10 3 2 4 12 9 15 11 8 6 14
      82.773 × 10−5−9727943.3381 13 7 5 10 3 2 4 12 9 15 11 8 6 14
      93.978 × 10−5−9758126.3461 13 7 5 3 10 4 2 12 9 15 11 8 6 14
      101.212 × 10−5−9686227.1871 13 7 5 3 10 4 2 12 9 15 11 8 14 6
      111.034 × 10−5−9658981.2411 13 7 5 3 10 4 2 12 9 15 11 8 14 6

      Table 3.  The iteration of CNOP-P computation.

      In the MRF scheme, the PBL and free atmosphere vertical diffusion is calculated based on the local K approach. The asymptotic mixing length (P1) is a physical quantity describing the size of the large energy-containing eddies in turbulence mixing, and is thus crucial in determining the vertical diffusion intensity. The mixing length scale is detected as the number one sensitive parameter, illustrating the rather important role of boundary layer processes in affecting locally developed convections. P2, the critical Richardson number, is a key parameter for calculating boundary layer height. P3 is a profile shape exponent for calculating the momentum diffusion coefficient, which determines the magnitude of the vertical diffusion as well as the height at which the diffusion reaches a maximum. P4, the Prandtl number coefficient, is involved in computing eddy diffusivity for temperature and moisture within the boundary layer. In the WSM6 microphysics scheme, the constants for the terminal velocity of rain and snow (P5, P7) play important roles in the vertical distribution of hydrometeors and auto-conversion in clouds. P10, the density of graupel, varies greatly, and can lead to obviously different characteristics of cloud microphysics. Setting P10 as a constant may underestimate or overestimate the precipitation efficiency of clouds. The maximum value of the cloud ice diameter, named P13 here, is related to the maximum mass of ice crystals, which is critical in auto-conversion from ice to snow, and further influences the rainfall intensity for the strong convection cases.

      These eight sensitive parameters are applied to address the model uncertainty using Eqs. (14)–(17). Table 4 gives the range of perturbations for these parameters. It should be noted that this range is a little different from the perturbed range in the calculation of CNOP. To ensure the representativeness and independence of samples, we perturbed ±50% of the defaults to generate the parameter perturbation samples in computing CNOP-P. Here, however, the perturbations need to be closer to the actual values in the whole variation ranges, which should be large enough to add reliable variability into the forecasts, and small enough to make the forecasts reasonable. Although these exact thresholds are not clear, we define the values by referring to previous research (McCabe et al., 2016; Ollinaho et al., 2017).

      IndexParameterMinDefaultMax
      P1rlam30150450
      P2brcr0.1250.51
      P3pfac123
      P4cfac3.97.815.6
      P5avtr420841.91263
      P7avts511.7218
      P10deng200500900
      P13dimax2 × 10−45 × 10−48 × 10−4

      Table 4.  Sensitive parameters and their ranges.

      The following are the results of the supplementary experiment, which, using five heavy rainfall cases under strong synoptic forcing, was in sharp contrast to the above in which we selected five cases with weak synoptic forcing. The sensitive parameters detected by this experiment are P1, P13, P5, P7, P2, P3, P14 and P12, as shown in Table 5. Only two parameters changed compared with the experiments shown in Table 3. The result indicates that the detection and sensitivity ranking of parameters is relatively stable and robust.

      IterationSpectral projection gradientCost functionSensitivity ranking corresponding to CNOP-P
      01.216−1923239.88010 13 3 11 8 5 12 4 9 6 2 15 7 14 1
      13.401 × 10−1−12148995.401 3 2 5 13 7 4 9 11 14 6 8 15 12 10
      28.955 × 10−3−12753984.431 5 3 14 12 13 2 4 7 9 11 8 10 6 15
      33.236 × 10−4−12522691.231 5 13 7 2 3 14 12 11 9 4 15 10 6 8
      41.971 × 10−4−12598208.521 13 5 7 2 3 14 12 11 4 9 10 15 6 8
      51.201 × 10−4−12565818.291 13 7 5 2 3 14 12 11 4 9 10 15 8 6
      68.993 × 10−5−12602431.251 13 7 5 2 3 14 12 11 9 4 10 15 6 8
      76.820 × 10−5−12546727.081 13 7 5 2 3 14 12 11 9 4 10 15 6 8
      82.925 × 10−5−12539894.941 13 7 5 2 3 14 12 11 9 4 10 15 6 8
      91.211 × 10−4−12602381.101 13 7 5 2 3 14 12 11 9 4 10 15 6 8
      104.626 × 10−5−12584054.481 13 7 5 2 3 14 12 11 9 4 10 15 6 8
      112.772 × 10−5−12547402.311 13 5 7 2 3 14 12 11 9 4 10 15 6 8

      Table 5.  The iteration of CNOP-P computation using five cases under strong synoptic forcing as initial conditions.

    • The purpose of this paper is to introduce the CNOP-P-based model uncertainty formulation and understand its advantages for CAEPSs. For a relatively fair comparison, we chose a total of 13 heavy rainfall cases to understand the systematic effects, and calculated the averaged differences of metrics over a forecasting time of 6–24 h between EXP1 and EXP2, i.e., EXP2 minus EXP1, to investigate the added value of the CNOP-P-based approach. According to the definition of these metrics, EXP2 performs better when the difference of S/R is greater than 0, or the differences of outlier frequency and CRPS are less than 0, and vice versa. Figure 2 shows the 6–24-h averaged differences of S/R on four variables for the 13 cases, with T2m, V10m, Q2m and R shown in Figs. 2ad, respectively. From all the positive values of S/R differences in Fig. 2, it is obvious that the CNOP-P-based method, i.e., EXP2, exhibits systematic improvement with respect to the S/R relationship for all the near-surface variables and precipitation. This result indicates that the CNOP-P-based method has great potential in improving the under-dispersive problem of current CAEPSs.

      Figure 2.  The 6–24-h time-averaged differences of S/R (spread/RMSE) for 13 cases. Black lines represent EXP2’s S/R on four variables, and columns represents the difference between EXP2 and EXP1 (i.e. EXP2 minus EXP1). (a) 2-m temperature; (b) 10-m wind speed; (c) 2-m specific humidity; (d) hourly precipitation.

      Figure 3.  As in Fig. 2 but for outlier frequency.

      Similar to Fig. 2, Fig. 3 gives the 6–24-h averaged differences of outlier frequency for each variable. As mentioned above, outlier frequency denotes the average probability that the observations fall outside the range of the forecasts. As shown in Fig. 3, the time-averaged differences of outlier frequency for T2m, V10m and Q2m exhibit negative values for all 13 cases. Only two exceptions are found in the precipitation outlier frequency difference (Fig. 3d). It can be concluded that, on average, the CNOP-P-based method can reduce the probability that the observations fall outside the range of the forecasts, thus enhancing the reliability of the CAEPS.

      The CRPS is frequently used to assess the respective accuracy of two probabilistic forecasting models. The 6–24-h time-averaged differences of CRPS for the two probabilistic forecasts are shown in Fig. 4 for each variable. According to the definition of CRPS, a negative time-averaged difference means a better probabilistic forecast. From Fig. 4, it can be found that, for all four variables, about 11 out of 13 cases show negative differences, indicating a better forecast corresponding to the cumulative distribution function of the observations.

      Figure 4.  As in Fig. 2 but for CRPS.

      As shown in Figs. 24, the above verification results indicate that the CNOP-P-based method can bring more spread, and has relatively reliable and skillful probabilistic forecasts compared with the SPPT scheme. That is, perturbing the most sensitive parameters or physics may have more potential to improve CAEPSs than perturbing the net physics tendency.

      Next, we check the different effects of the two approaches in the vertical direction. Figure 5 shows the vertical profiles of ensemble spread for the 24-h forecast of zonal wind, temperature and specific humidity, the profiles of which are the averages of the 13 cases. For the zonal wind field in Fig. 5a, EXP2 can obtain larger spread within the boundary layer, while EXP1 seems to be more dispersive in the middle atmosphere. This may partly reflect the characteristics of the two approaches in formulating the model uncertainties, but on the whole the difference is not significant. For temperature and specific humidity in Fig. 5b and Fig. 5c, EXP2 has consistently and statistically significant larger spread over the troposphere compared with EXP1. This again confirms the merits of the CNOP-P-based approach, because through detecting the most sensitive parameters from CNOP-P, the process-specific and event-oriented model uncertainties can be realized. More specifically, focusing on the CAEPS of convective events in this study, the most sensitive parameters embedded in the PBL and microphysics are detected by using CNOP-P. Thus, the formulation of model uncertainty of these two key processes for convection, which has solid theoretical basis, can bring more spread for temperature and humidity over the troposphere.

      Figure 5.  Vertical profiles of ensemble spread for the 24-h forecasts of 13 cases: (a) zonal wind (units: m s−1); (b) temperature (units: °C); (c) specific humidity (units: g kg−1).

      Finally, it should be noted that some ensembles might not necessarily be under-dispersive in terms of all parameters, so that our study may be more about creating “a right kind of spread” rather than increasing the spread in absolute terms. A sensitivity test was conducted to explain why we chose to perturb the first eight sensitive parameters in the ranking of Table 3. Five comparison experiments were implemented, each of which perturbed the first 4, 6, 8, 10 and 12 parameters in the sensitivity ranking of Table 3. We chose Case 1 in Table 1 for ensemble prediction, and calculated the differences of three metrics over the 6–24-h forecasting time (Figs. 68). From the results of the three verification metrics, we found that the perturbation scheme with eight parameters is more effective than that with four or six parameters. However, increasing to more than eight parameters does not obviously improve the forecasts. It can be seen that the uncertainty of parameters is mainly due to the important sensitive parameters.

      Figure 6.  The 6–24-h time-averaged S/R (spread/RMSE) for the case initialized at 1200 UTC 6 May. EXP_4 (red), EXP_6 (blue), EXP_8 (black) EXP_10 (grey), and EXP_12 (green) are five ensemble forecasts that perturbed the first 4, 6, 8, 10 and 12 parameters, respectively: (a) 2-m temperature; (b) 10-m wind speed; (c) 2-m specific humidity; (d) hourly precipitation.

      Figure 7.  As in Fig. 6 but for outlier frequency.

      Figure 8.  As in Fig. 6 but for CRPS.

    4.   Summary and remarks
    • The model perturbation methods in CAEPSs lack consideration of the rapidly nonlinear growing errors at the convective scale, which is partly responsible for the under-dispersion problem of CAEPSs. This paper seeks to address this problem by designing a more reasonable formulation of model uncertainty based on the CNOP-P approach. The CNOP-P can search for the optimal model parameter perturbations, which result in the largest departure from a given reference state at a prediction time with physical constraints. Focusing on locally developed strong convective events that occurred in the pre-rainy season of South China, the most sensitive parameters were detected using CNOP-P, which resulted in the largest departure of heavy rainfall forecasts for five heavy rainfall cases. Then, a formulation of model uncertainty was designed by adding stochastic perturbations into these sensitive parameters. Comparison ensemble experiments between the CNOP-P-based method and the well-utilized SPPT scheme were conducted, in which the ensemble system only considered the model uncertainties. These experiments were carried out using all the 13 heavy rainfall cases that occurred in the flood season of South China in 2017. Our conclusions and remarks are as follows:

      The CNOP-P approach is very robust in detecting the sensitive parameters when applied to the convection-allowing-scale model.

      The CNOP-P-based method has potential in improving the under-dispersive problem of current CAEPSs, and has more reliable forecast skill for T2m, V10m, Q2m and precipitation compared with the SPPT scheme.

      The CNOP-P-based method brings more spread for humidity and temperature over the troposphere owing to its process-specific and event-oriented formulation of model uncertainties.

      Although the cases we selected in this study are specific to South China and may therefore limit the conclusions, the experiments with these 13 heavy rainfall cases in the 2017 flood season strongly suggest a very bright future in terms of the application of the CNOP-P-based approach in CAEPSs.

      There is, however, still a practical and critical issue to resolve in terms of how to select the parameters if two of the eight parameters are different when different cases are used. The CNOP-P approach has been proven theoretically in previous studies and in some climate model applications, and it is believed in practical NWP applications that a stable group comprising the most sensitive parameters could be obtained if the cases are chosen to share the similar physical features or processes. The supplementary experiment in our study illustrates this issue, i.e., the parameters based on the “strong forcing” cases were similar to those based on the “weak forcing” cases. Much research is needed in which the CNOP-P approach is further applied in operational CAEPSs—such as, its application in heavy rainfall events in the mid-to-high latitudes, the impact of parameter perturbation constraints, the use of spatiotemporal random perturbation approach when perturbing the parameters, and the choice of cost functions.

      Acknowledgements. We sincerely appreciate the constructive comments and suggestions of the anonymous reviewers. This work was supported by the National Key Research and Development Program of China (Grant No. 2017YFC150 1904).

      APPENDIX

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return