Advanced Search
Article Contents

Possible Sources of Forecast Errors Generated by the Global/Regional Assimilation and Prediction System for Landfalling Tropical Cyclones. Part I: Initial Uncertainties


doi: 10.1007/s00376-016-5238-4

  • This paper investigates the possible sources of errors associated with tropical cyclone (TC) tracks forecasted using the Global/Regional Assimilation and Prediction System (GRAPES). The GRAPES forecasts were made for 16 landfalling TCs in the western North Pacific basin during the 2008 and 2009 seasons, with a forecast length of 72 hours, and using the default initial conditions ("initials", hereafter), which are from the NCEP-FNL dataset, as well as ECMWF initials. The forecasts are compared with ECMWF forecasts. The results show that in most TCs, the GRAPES forecasts are improved when using the ECMWF initials compared with the default initials. Compared with the ECMWF initials, the default initials produce lower intensity TCs and a lower intensity subtropical high, but a higher intensity South Asia high and monsoon trough, as well as a higher temperature but lower specific humidity at the TC center. Replacement of the geopotential height and wind fields with the ECMWF initials in and around the TC center at the initial time was found to be the most efficient way to improve the forecasts. In addition, TCs that showed the greatest improvement in forecast accuracy usually had the largest initial uncertainties in TC intensity and were usually in the intensifying phase. The results demonstrate the importance of the initial intensity for TC track forecasts made using GRAPES, and indicate the model is better in describing the intensifying phase than the decaying phase of TCs. Finally, the limit of the improvement indicates that the model error associated with GRAPES forecasts may be the main cause of poor forecasts of landfalling TCs. Thus, further examinations of the model errors are required.
  • 加载中
  • Brennan M. J., S. J. Majumdar, 2011: An examination of model track forecast errors for Hurricane Ike (2008) in the Gulf of Mexico. Wea.Forecasting, 26, 848- 867.10.1175/WAF-D-10-05053.142f56357e1ef12d995fceaebeeffe3a8http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2011WtFor..26..848Bhttp://adsabs.harvard.edu/abs/2011WtFor..26..848BNot Available
    Chan J. C. L., 2010: Movement of tropical cyclones. Global Perspectives on Tropical Cyclones, J. C. L. Chan and J. D. Kepert, Eds., World Science, Singapore, 133- 148.eb876eca-9c62-49f1-a698-08e14fc88e0734354391809de1417c4080bf08cca688http%3A%2F%2Fwww.worldscientific.com%2Fdoi%2Fpdf%2F10.1142%2F9789814293488_0004refpaperuri:(b5e3b532939661585d75d40a295524d0)http://www.worldscientific.com/doi/pdf/10.1142/9789814293488_0004
    Chen L.S., 1965: Flow patterns in westerlies in relation to the East-Asia typhoon tracks. Acta Meteorologica Sinica (in Chinese), 35, 476- 485.
    Chen, D. H., Coauthors, 2008: New generation of multi-scale NWP system (GRAPES): General scientific design. Chinese Science Bulletin, 53, 3433- 3445.10.1007/s11434-008-0494-z426cc55bf84622f194822ec3b646ebafhttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs11434-008-0494-zhttp://www.cnki.com.cn/Article/CJFDTotal-JXTW200822003.htmA new generation of numerical prediction system GRAPES (a short form of Global/Regional Assimilation and PrEdiction System) was set up in China Meteorological Administration (CMA). This paper focuses on the scientific design and preliminary results of the numerical prediction model in GRAPES, including basic idea and strategy of the general scientific design, multi-scale dynamic core, physical package configuration, architecture and parallelization of the codes. A series of numerical experiments using the real data with horizontal resolutions from 10 to 280 km and idealized experiments with very high resolution up to 100 m are conducted, giving encouraging results supporting the multi-scale application of GRAPES. The results of operational implementation of GRAPES model in some NWP centers are also presented with stress at evaluations of the capability to predict the main features of precipitation in China. Finally the issues to be dealt with for further development are discussed.
    Dai, Y. J., Coauthors, 2003: The common land model. Bull. Amer. Meteor. Soc., 84, 1013- 1023.10.1175/BAMS-84-8-10134ec35d76-3852-483f-a426-b6051d2d504e20771a7b6c6eb603520971725d92ee14http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F80016174835refpaperuri:(9322c11e3e9ec412be88573e84d889b7)http://ci.nii.ac.jp/naid/80016174835Abstract The Common Land Model (CLM) was developed for community use by a grassroots collaboration of scientists who have an interest in making a general land model available for public use and further development. The major model characteristics include enough unevenly spaced layers to adequately represent soil temperature and soil moisture, and a multilayer parameterization of snow processes; an explicit treatment of the mass of liquid water and ice water and their phase change within the snow and soil system; a runoff parameterization following the TOPMODEL concept; a canopy photosynthesis-onductance model that describes the simultaneous transfer of CO 2 and water vapor into and out of vegetation; and a tiled treatment of the subgrid fraction of energy and water balance. CLM has been extensively evaluated in offline mode and coupling runs with the NCAR Community Climate Model (CCM3). The results of two offline runs, presented as examples, are compared with observations and with the simulation of three other land models [the Biosphere- tmosphere Transfer Scheme (BATS), Bonan's Land Surface Model (LSM), and the 1994 version of the Chinese Academy of Sciences Institute of Atmospheric Physics LSM (IAP94)].
    Ding Y.H., E.R. Wright, 1983: The large scale circulation condition for the western Pacific typhoon genesi. Acta Oceanologica Sinica (in Chinese), 5, 561- 574.
    Duan W. S., F. F. Zhou, 2013: Nonlinear forcing singular vector of a two-dimensional quasi-geostrophic model. Tellus A.,65, 18452, .v65i0. 18452.http://dx.doi.org/10.3402/tellusa
    Duan W. S., P. Zhao, 2015: Revealing the most disturbing tendency error of Zebiak-Cane model associated with El Niño predictions by nonlinear forcing singular vector approach. Climate Dyn.,44, 2351-2367, doi: 10.1007/s00382-014-2369-0.
    Fiorino M., 2009: Record-setting performance of the ECMWF IFS in medium-range tropical cyclone track prediction. ECMWF Newsletter, 118, 20- 27.
    Galarneau T. J.Jr., C. A. Davis, 2013: Diagnosing forecast errors in tropical cyclone motion. Mon. Wea. Rev., 141, 405- 430.10.1175/MWR-D-12-00071.10d0e795c1f97155ba61b404d68c084b7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2013MWRv..141..405Ghttp://adsabs.harvard.edu/abs/2013MWRv..141..405GNot Available
    Galarneau, T. J. Jr., T. M. Hamill, 2015: Diagnosis of track forecast errors for tropical cyclone Rita (2005) using GEFS reforecasts. Wea. Forecasting,30, 1334-1354, doi: 10.1175/ WAF-D-15-0036.1.10.1175/WAF-D-15-0036.1c9fcfc2ba7b9107bdfe3b4cc027a2438http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015WtFor..30.1334Ghttp://adsabs.harvard.edu/abs/2015WtFor..30.1334GNot Available
    Green B. W., F. Q. Zhang, 2014: Sensitivity of tropical cyclone simulations to parametric uncertainties in air-sea fluxes and implications for parameter estimation. Mon. Wea. Rev. , 142, 2290- 2308.10.1175/MWR-D-13-00208.1490b5c92-1661-4a61-ad47-b259750831b10f9258e989eac75470b00ec29f8e8964http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FMWR-D-13-00208.1http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1175/MWR-D-13-00208.1Not Available
    Han J., H. L. Pan, 2011: Revision of convection and vertical diffusion schemes in the NCEP global forecast system. Wea.Forecasting, 26, 520- 533.10.1175/WAF-D-10-05038.1b268077d000babacfc600f4b4b31ac76http%3A%2F%2Fconnection.ebscohost.com%2Fc%2Farticles%2F64842702%2Frevision-convection-vertical-diffusion-schemes-ncep-global-forecast-systemhttp://connection.ebscohost.com/c/articles/64842702/revision-convection-vertical-diffusion-schemes-ncep-global-forecast-systemAbstract A new physics package containing revised convection and planetary boundary layer (PBL) schemes in the National Centers for Environmental Prediction- Global Forecast System is described. The shallow convection (SC) scheme in the revision employs a mass flux parameterization replacing the old turbulent diffusion-based approach. For deep convection, the scheme is revised to make cumulus convection stronger and deeper to deplete more instability in the atmospheric column and result in the suppression of the excessive grid-scale precipitation. The PBL model was revised to enhance turbulence diffusion in stratocumulus regions. A remarkable difference between the new and old SC schemes is seen in the heating or cooling behavior in lower-atmospheric layers above the PBL. While the old SC scheme using the diffusion approach produces a pair of layers in the lower atmosphere with cooling above and heating below, the new SC scheme using the mass-flux approach produces heating throughout the convection layers. In particular, the new SC scheme does not destroy stratocumulus clouds off the west coasts of South America and Africa as the old scheme does. On the other hand, the revised deep convection scheme, having a larger cloud-base mass flux and higher cloud tops, appears to effectively eliminate the remaining instability in the atmospheric column that is responsible for the excessive grid-scale precipitation in the old scheme. The revised PBL scheme, having an enhanced turbulence mixing in stratocumulus regions, helps prevent too much low cloud from forming. An overall improvement was found in the forecasts of the global 500-hPa height, vector wind, and continental U.S. precipitation with the revised model. Consistent with the improvement in vector wind forecast errors, hurricane track forecasts are also improved with the revised model for both Atlantic and eastern Pacific hurricanes in 2008.
    Heming J., J. Goerss, 2010: Track and structure forecasts of tropical cyclones. Global Perspectives on Tropical Cyclones, J. C. L. Chan and J. D. Kepert, Eds., World Science, Singapore, 287- 323.666bc98b-f17f-49df-bae4-db90f23bddfd71a5e2609d049620778dd961d3ef0583http%3A%2F%2Fwww.worldscientific.com%2Fdoi%2Fabs%2F10.1142%2F9789814293488_0010%3Fmi%3D6egmof%26af%3DR%26Contrib%3DKepert%252C%2BJeffrey%2BD%26content%3DarticlesChapters%26countTerms%3Dtrue%26target%3Ddefaultrefpaperuri:(8b0fc69a65f1220152c53132a4759465)http://www.worldscientific.com/doi/abs/10.1142/9789814293488_0010?mi=6egmof&af=R&Contrib=Kepert%2C+Jeffrey+D&content=articlesChapters&countTerms=true&target=default
    Hong S. Y., J. O. J. Lim, 2006: The WRF single-moment 6-class microphysics scheme (WSM6). Asia-Pacific Journal of Atmospheric Sciences, 42, 129- 151.7308c59e0fe08d8147ff5b2869261e63http%3A%2F%2Fwww.dbpia.co.kr%2FJournal%2FArticleDetail%2F773025http://www.dbpia.co.kr/Journal/ArticleDetail/773025This study examines the performance of the Weather Research and Forecasting (WRF)-Single-Moment- Microphysics scheme (WSMMPs) with a revised ice-microphysics of the Hong et al. In addition to the simple (WRF Single-Moment 3-class Microphysics scheme; WSM3) and mixed-phase (WRF Single-Moment 5-class Microphysics scheme; WSM5) schemes of the Hong et al., a more complex scheme with the inclusion of graupel as another predictive variable (WRF Single-Moment 6-class Microphysics scheme; WSM6) was developed. The characteristics of the three categories of WSMMPs were examined for an idealized storm case and a heavy rainfall event over Korea. In an idealized thunderstorm simulation, the overall evolutionary features of the storm are not sensitive to the number of hydrometeors in the WSMMPs; however, the evolution of surface precipitation is significantly influenced by the complexity in microphysics. A simulation experiment for a heavy rainfall event indicated that the evolution of the simulated precipitation with the inclusion of graupel (WSM6) is similar to that from the simple (WSM3) and mixed-phase (WSM5) microphysics in a low-resolution grid; however, in a high-resolution grid, the amount of rainfall increases and the peak intensity becomes stronger as the number of hydrometeors increases.
    Hong S. Y., H. L. Pan, 1996: Nonlocal boundary layer vertical diffusion in a medium-range forecast model. Mon. Wea. Rev., 124, 2322- 2339.10.1175/1520-0493(1996)1242.0.CO;2ef05aa5667be0bc7e5538f034b3cba21http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F10012348693http://ci.nii.ac.jp/naid/10012348693Not Available
    Huang Y. Y., J. S. Xue, Q. L. Wan, Z. T. Chen, W. Y. Ding, and C. Z. Zhang, 2013: Improvement of the surface pressure operator in GRAPES and its application in precipitation forecasting in South China. Adv. Atmos. Sci.,30(2), 354-366, doi: 10.1007/s00376-012-1270-1.10.1007/s00376-012-1270-10190822dbf2d968d55e551bf55a34e81http%3A%2F%2Fwww.cqvip.com%2FQK%2F84334X%2F201302%2F44841864.htmlhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e201302009.aspxIn this study we investigated the problems involved in assimilating surface pressure in the current global and regional assimilation and prediction system,GRAPES.A new scheme of assimilating surface pressure was proposed,including a new interpolation scheme and a refreshed background covariance.The new scheme takes account of the differences between station elevation and model topography,and it especially deals with stations located at elevations below that of the first model level.Contrast experiments were conducted using both the original and the new assimilation schemes.The influence of the new interpolation scheme and the updated background covariance were investigated.Our results show that the new interpolation scheme utilized more observations and improved the quality of the mass analysis.The background covariance was refreshed using statistics resulting from the technique proposed by Parrish and Derber in 1992.Experiments show that the updated vertical covariance may have a positive influence on the analysis at higher levels of the atmosphere when assimilating surface pressure.This influence may be more significant if the quality of the background field at high levels is poor.A series of assimilation experiments were performed to test the validity of the new scheme.The corresponding simulation experiments were conducted using the analysis of both schemes as initial conditions.The results indicated that the new scheme leads to better forecasting of sea level pressure and precipitation in South China,especially the forecast of moderate and heavy rain.
    Huang B., D. H. Chen, X. L. Li, and C. Li, 2014: Improvement of the semi-Lagrangian advection scheme in the GRAPES model: Theoretical analysis and idealized tests. Adv. Atmos. Sci. ,31(3), 693-704, doi:10.1007/s00376-013-3086-z.10.1007/s00376-013-3086-z24824911c54cf5b4afaee97abf0c9c92http%3A%2F%2Fd.wanfangdata.com.cn%2FPeriodical_dqkxjz-e201403019.aspxhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e201403019.aspxThe Global/Regional Assimilation and PrEdiction System (GRAPES) is the new-generation numerical weather prediction (NWP) system developed by the China Meteorological Administration. It is a fully compressible non-hydrostatical global/regional unified model that uses a traditional semi-Lagrangian advection scheme with cubic Lagrangian interpolation (referred to as the SL_CL scheme). The SL_CL scheme has been used in many operational NWP models, but there are still some deficiencies, such as the damping effects due to the interpolation and the relatively low accuracy. Based on Reich- semi-Lagrangian advection scheme (referred to as the R2007 scheme), the Re_R2007 scheme that uses the low- and high-order B-spline function for interpolation at the departure point, is developed in this paper. One- and two-dimensional idealized tests in the rectangular coordinate system with uniform grid cells were conducted to compare the Re_R2007 scheme and the SL_CL scheme. The numerical results showed that: (1) the damping effects were remarkably reduced with the Re_R2007 scheme; and (2) the normalized errors of the Re_R2007 scheme were about 7.5 and 3 times smaller than those of the SL_CL scheme in one- and two-dimensional tests, respectively, indicating the higher accuracy of the Re_R2007 scheme. Furthermore, two solid-body rotation tests were conducted in the latitude-longitude spherical coordinate system with nonuniform grid cells, which also verified the Re_R2007 scheme- advantages. Finally, in comparison with other global advection schemes, the Re_R2007 scheme was competitive in terms of accuracy and flow independence. An encouraging possibility for the application of the Re_R2007 scheme to the GRAPES model is provided.
    Iacono M. J., J. S. Delamere, E. J. Mlawer, M. W. Shephard, S. A. Clough, and W. D. Collins, 2008: Radiative forcing by long-lived greenhouse gases: Calculations with the AER radiative transfer models. J. Geophys. Res., 113,D13103, doi: 10.1029/2008JD009944.10.1029/2008JD0099440b25c1c2a104d51c498700a19269e7f0http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008JD009944%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2008JD009944/fullA primary component of the observed, recent climate change is the radiative forcing from increased concentrations of long-lived greenhouse gases (LLGHGs). Effective simulation of anthropogenic climate change by general circulation models (GCMs) is strongly dependent on the accurate representation of radiative processes associated with water vapor, ozone and LLGHGs. In the context of the increasing application of the Atmospheric and Environmental Research, Inc. (AER) radiation models within the GCM community, their capability to calculate longwave and shortwave radiative forcing for clear sky scenarios previously examined by the radiative transfer model intercomparison project (RTMIP) is presented. Forcing calculations with the AER line-by-line (LBL) models are very consistent with the RTMIP line-by-line results in the longwave and shortwave. The AER broadband models, in all but one case, calculate longwave forcings within a range of -0.20 to 0.23 W m{sup -2} of LBL calculations and shortwave forcings within a range of -0.16 to 0.38 W m{sup -2} of LBL results. These models also perform well at the surface, which RTMIP identified as a level at which GCM radiation models have particular difficulty reproducing LBL fluxes. Heating profile perturbations calculated by the broadband models generally reproduce high-resolution calculations within a few hundredths K d{sup more 禄 -1} in the troposphere and within 0.15 K d{sup -1} in the peak stratospheric heating near 1 hPa. In most cases, the AER broadband models provide radiative forcing results that are in closer agreement with high 20 resolution calculations than the GCM radiation codes examined by RTMIP, which supports the application of the AER models to climate change research. less
    Kepert J. D., 2011: Choosing a boundary layer parameterization for tropical cyclone modeling. Mon. Wea. Rev., 140, 1427- 1445.10.1175/MWR-D-11-00217.1c39b89e2bcc3074a96034d187373d05ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012MWRv..140.1427Khttp://adsabs.harvard.edu/abs/2012MWRv..140.1427KNot Available
    Liu C. H., Z. M. Tan, and Y. Zhang, 2010: On the deviations of approximate model of forecasting errors growth. Journal of Nanjing University (Natural Sciences), 46, 607- 615. (in Chinese)10.3788/gzxb20103906.0998ed15914dcd8a3426123dd28723611132http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-NJDZ201006002.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-NJDZ201006002.htmThe deviations of accumulative approximate model and linear approximate model of errors' growth are discussed in theoretical and numerical experiments.Theoretical results show that the deviations of accumulative approximation for the initial errors,as well as for the model errors,grow linearly with time,but the growth rate of the deviations for model errors is about one half of that for initial errors.The deviations of linear approximation of initial or model errors are proportional to the norm of errors' linear part.The results are illustrated by experiments in Lorenz'96 system.Experimental results also show that the deviations of accumulative approximation and linear approximation are both insensitive to background state,and the deviations of accumulative approximation of model errors are insensitive to model errors' sources.These results provide theoretical basis for understanding the applicable range of approximate models and the separation of error growth phases.
    McTaggart-Cowan R., L. F. Bosart, J. R. Gyakum, and E. H. Atallah, 2006: Hurricane Juan (2003). Part II. Forecasting and numerical simulation. Mon. Wea. Rev., 134, 1748- 1771.3253b667-7439-4f5f-8cc1-7b5c56dbc0b60db9ad80cbcfe9413c80d0cbc6f40d76http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2006MWRv..134.1748M%26db_key%3DPHY%26link_type%3DEJOURNALrefpaperuri:(09a15e0a3b71c17594e001347b7c9985)/s?wd=paperuri%3A%2809a15e0a3b71c17594e001347b7c9985%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2006MWRv..134.1748M%26db_key%3DPHY%26link_type%3DEJOURNAL&ie=utf-8&sc_us=1576694230275247338
    Nicolis C., R. A. P. Perdigao, and S. Vannitsem, 2009: Dynamics of prediction errors under the combined effect of initial condition and model errors. J. Atmos. Sci., 66, 766- 778.d6a5c098c9900e362ef61bcd7a1302efhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009JAtS...66..766N/s?wd=paperuri%3A%287d72fae4b74a50d8342320ac1bef546c%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009JAtS...66..766N&ie=utf-8&sc_us=17168213878657832876
    Orrell D., 2003: Model error and predictability over different timescales in the Lorenz-96 systems. J. Atmos. Sci., 60, 2219- 2228.10.1175/1520-0469%282003%29060%3C2219%3AMEAPOD%3E2.0.CO%3B246ed5886-8d47-4d81-a461-81250a383618a4a776463ef4e63a96d0dcbb94df1c7ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2003JAtS...60.2219Orefpaperuri:(a8fd3e9a094b34b0f312011c8efd700b)http://adsabs.harvard.edu/abs/2003JAtS...60.2219OABSTRACT Prediction problems have been described by Lorenz as falling into two categories. Problems that depend on the initial condition, such as short-to medium-range weather forecasting, are described as ''predictions of the first kind,'' while problems that depend on boundary rather than initial conditions, such as, in many cases, the longer-term climatology, are referred to as predictions of the second kind. Both kinds of prediction will be affected by error in the model equations used to approximate the true system. In this paper, predictability over different timescales for the medium-dimensional Lorenz '96 systems is examined. Models are constructed for the purposes of optimizing both short-range prediction and climatological behavior, and studied over a range of forcings for which they show periodic, quasi-periodic, or chaotic behavior. It is shown that, for the models discussed here, there is a link between short-and long-range predictability, which is held independent of the effects of chaos. The role of stochastic terms is considered, and the possible implications for atmospheric or oceanographic modeling are discussed.
    Orrell D., L. Smith, J. Barkmeijer, and T. N. Palmer, 2001: Model error in weather forecasting. Nonlinear Processes in Geophysics, 8, 357- 371.896a2dbe-4288-4a17-b0c5-c4c927458d5f14aa8390bb2c8e94a3d39cec664fb283http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2001NPGeo...8..357O%26db_key%3DPHY%26link_type%3DABSTRACTrefpaperuri:(91ab95621c6dfcc0021a74af0585e6fb)/s?wd=paperuri%3A%2891ab95621c6dfcc0021a74af0585e6fb%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2001NPGeo...8..357O%26db_key%3DPHY%26link_type%3DABSTRACT&ie=utf-8&sc_us=14807337250055183968
    Rappaport, E. N., Coauthors, 2009: Advances and challenges at the National Hurricane Center. Wea.Forecasting, 24, 395- 419, doi. 10.1175/2008WAF2222128.1.10.1175/2008WAF2222128.14d87fc2a18c68ff8a54160435ca8b554http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009WtFor..24..395Rhttp://adsabs.harvard.edu/abs/2009WtFor..24..395RNot Available
    Roy C., R. Kovordanyi, 2012: Tropical cyclone track forecasting techniques-A review. Atmospheric Research,104-105, 40- 69.10.1016/j.atmosres.2011.09.012ac0ddb85-ef5d-487a-9ec3-15bc7cef4e2d3e2a5647f5c4f65a48a2dc3342116f49http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0169809511002973refpaperuri:(cc7e04b280569a50048c23052031b1cd)http://www.sciencedirect.com/science/article/pii/S0169809511002973Delivering accurate cyclone forecasts in time is of key importance when it comes to saving human lives and reducing economic loss. Difficulties arise because the geographical and climatological characteristics of the various cyclone formation basins are not similar, which entail that a single forecasting technique cannot yield reliable performance in all ocean basins. For this reason, global forecasting techniques need to be applied together with basin-specific techniques to increase the forecast accuracy. As cyclone track is governed by a range of factors variations in weather conditions, wind pressure, sea surface temperature, air temperature, ocean currents, and the earth's rotational force-he coriolis force, it is a formidable task to combine these parameters and produce reliable and accurate forecasts. In recent years, the availability of suitable data has increased and more advanced forecasting techniques have been developed, in addition to old techniques having been modified. In particular, artificial neural network based techniques are now being considered at meteorological offices. This new technique uses freely available satellite images as input, can be run on standard PCs, and can produce forecasts with good accuracy. For these reasons, artificial neural network based techniques seem especially suited for developing countries which have limited capacity to forecast cyclones and where human casualties are the highest.
    Ryan D. T., J. S. Whitaker, P. Pegion, T. M. Hamill, and G. J. Hakim, 2015: Diagnosis of the source of GFS medium-range track errors in Hurricane Sandy (2012). Mon. Wea. Rev., 143, 132- 152.10.1175/MWR-D-14-00086.119f75d5ac5504244c27b65e9ba138457http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015MWRv..143..132Thttp://adsabs.harvard.edu/abs/2015MWRv..143..132TAbstract Medium-range forecasts of Hurricane Sandy- track were characterized by widely diverging solutions, with some suggesting that Sandy would make landfall over the mid-Atlantic region of the United States, while others forecast the storm to move due east to the north of Bermuda. Here, dynamical processes responsible for the eastward-tracking forecasts are diagnosed using an 80-member ensemble of experimental Global Forecast System (GFS) forecasts initialized five days prior to landfall. Comparing the ensemble members with tracks to the east against those with tracks to the west indicates that the eastern members were characterized by a lower-amplitude upper-tropospheric anticyclone on the poleward side of Sandy during the first 24 h of the forecast, which in turn was associated with a westerly perturbation steering wind. The amplification of this ridge in each set of members was modulated by differences in the advection of potential vorticity (PV) by the irrotational wind associated with Sandy- secondary circulation and isentropic lift along a warm front that formed on the poleward side of Sandy. The amplitude of the irrotational wind in this region was proportional to the 0-h water vapor mixing ratio, and to a lesser extent the 0-h upper-tropospheric horizontal divergence. These two quantities modulated the vertical profile of grid-scale condensation within the model and subsequent upper-tropospheric divergence. The results from this study suggest that additional observations within regions of large-scale precipitation outside the tropical cyclone (TC) core could benefit TC track forecasts, particularly when the TC is located near an upper-tropospheric PV gradient.
    Webster P. J., G. J. Holland , J. A. Curry, and H. R. Chang, 2005: Changes in tropical cyclone number,duration, and intensity in a warming environment. Science , 309, 1844-1846, doi:10.1126/science.1116448.10.1126/science.1116448161665144b3da70b2014d037261218dece85528dhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FPMED%3Fid%3D16166514http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM16166514Abstract We examined the number of tropical cyclones and cyclone days as well as tropical cyclone intensity over the past 35 years, in an environment of increasing sea surface temperature. A large increase was seen in the number and proportion of hurricanes reaching categories 4 and 5. The largest increase occurred in the North Pacific, Indian, and Southwest Pacific Oceans, and the smallest percentage increase occurred in the North Atlantic Ocean. These increases have taken place while the number of cyclones and cyclone days has decreased in all basins except the North Atlantic during the past decade.
    Xu H., M. Mu, and D. H. Luo, 2004: Application of nonlinear optimization method to sensitivity analysis of numerical model. Progress in Natural Science, 14, 546- 549.10.1080/100200704123313439216924370f5e92568ce3fab73b1e81671ahttp%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-ZKJY200406013.htmhttp://d.wanfangdata.com.cn/Periodical_zrkxjz-e200406014.aspxA nonlinear optimization method is applied to sensitivity analysis of a numerical model. Theoretical analysis and numerical experiments indicate that this method can give not only a quantitative assessment whether the numerical model is able to simulate the observations or not, but also the initial field that yields the optimal simulation. In particular, when the simulation results are apparently satisfactory, and sometimes both model error and initial error are considerably large, the nonlinear optimization method, under some conditions, can identify the error that plays a dominant role.
    Yamaguchi M., S. J. Majumdar, 2010: Using TIGGE data to diagnose initial perturbations and their growth for tropical cyclone ensemble forecasts. Mon. Wea. Rev., 138, 3634- 3655.72a16dc7-181e-4f9f-8bef-78d38c27450bd4448d47547f48fac5251570d4d66f20http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2010MWRv..138.3634Y%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D15332refpaperuri:(1a26b5d2f0de3fb9f40b4ae5208c5771)/s?wd=paperuri%3A%281a26b5d2f0de3fb9f40b4ae5208c5771%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2010MWRv..138.3634Y%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D15332&ie=utf-8&sc_us=8067665282700353199
    Yamaguchi M., T. Nakazawa, and K. Aonashi, 2012: Tropical cyclone track forecasts using JMA model with ECMWF and JMA initial conditions. Geophys. Res. Lett., 39, L09801.10.1029/2012GL05147380d1902548460ca99885127c4610dd21http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL051473%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2012GL051473/fullThe JMA's Global Spectral Model (JMA/GSM) was run from the initial conditions of ECMWF, which are available in the YOTC data set, to distinguish between TC track prediction errors attributable to the initial conditions and those attributable to the NWP model. The average position error was reduced by about 10% by replacing the initial conditions, and in some cases, the predictions were significantly improved. In these cases, the low wavenumber component of the ECMWF analysis was found to account for most of the improvement. In addition, the observed tracks were captured by the JMA Typhoon Ensemble Prediction System (TEPS), which deals with initial condition uncertainties. In some cases, however, the replacement of the initial conditions did not improve the prediction even when the ECMWF forecast was accurate. In these cases, TEPS could not capture the observed track either, implying the need for dealing with uncertainties associated with the NWP model.
    Ying M., W. Zhang, H. Yu, X. Lu, J. Feng, Y. Fan, Y. Zhu, and D. Chen, 2014: An overview of the China Meteorological Administration tropical cyclone database. J. Atmos. Oceanic Technol., 31, 287- 301. doi: 10.1175/JTECH-D-12-00119.110.1175/JTECH-D-12-00119.1d13a05b6042b59ed8aa6180668d48875http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014JAtOT..31..287Yhttp://adsabs.harvard.edu/abs/2014JAtOT..31..287YNot Available
    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.
    Zhang R. H., X. S. Shen, 2008: On the development of GRAPES-A new generation of the national operational NWP system in China. Chinese Science Bulletin, 53, 3429- 3432.10.1007/s11434-008-0462-7cce987f7-0030-41b7-893f-4b9f7315dbd1e360f29cfcb8dbc80334cfd494da6110http%3A%2F%2Flink.springer.com%2F10.1007%2Fs11434-008-0462-7refpaperuri:(e7f1c255e9e8e8051d9a963b605870ee)http://www.cnki.com.cn/Article/CJFDTotal-JXTW200822002.htmNumerical weather prediction (NWP) has become one of the most important means for weather forecasts in the world. It also mirrors a nation’s comprehensive strength in meteorology. In 2000, China Meteorological Administration (CMA) established the National Innovative Base for Meteorological Numerical Prediction in the Chinese Academy of Meteorological Sciences (CAMS), to work on developing a new generation of the national operational NWP system—Global/Regional Assimilation and PrEdiction System (GRAPES), to enhance meteorological services in China in the new century. In recent years, the GRAPES has witnessed a fast development. The GRAPES has been set up as an integration of the model framework, data assimilation, regional and global NWP system, which can be commonly used for both operation and research. In this paper, a brief review is made for illustrating the GRAPES system, including the advanced designs of the GRAPES, its diverse applications in multi-fields, and efficiencies of the regional and global GRAPES in operational applications based on hindcast results.
  • [1] 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
    [2] Chang-Hoi HO, Joo-Hong KIM, Hyeong-Seog KIM, Woosuk CHOI, Min-Hee LEE, Hee-Dong YOO, Tae-Ryong KIM, Sangwook PARK, 2013: Technical Note on a Track-pattern-based Model for Predicting Seasonal Tropical Cyclone Activity over the Western North Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 1260-1274.  doi: 10.1007/s00376-013-2237-6
    [3] GAO Feng*, Peter P. CHILDS, Xiang-Yu HUANG, Neil A. JACOBS, and Jinzhong MIN, 2014: A Relocation-based Initialization Scheme to Improve Track-forecasting of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 27-36.  doi: 10.1007/s00376-013-2254-5
    [4] Shaoying LI, Jun PENG, Weimin ZHANG, Jianping WU, Qiang YAO, Xiangrong YANG, Tengling LUO, 2023: Effects of a Dry-Mass Conserving Dynamical Core on the Simulation of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 464-482.  doi: 10.1007/s00376-022-2085-3
    [5] Chenxi WANG, Zhihua ZENG, Ming YING, 2020: Uncertainty in Tropical Cyclone Intensity Predictions due to Uncertainty in Initial Conditions, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 278-290.  doi: 10.1007/s00376-019-9126-6
    [6] QIN Xiaohao, MU Mu, 2014: Can Adaptive Observations Improve Tropical Cyclone Intensity Forecasts?, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 252-262.  doi: 10.1007/s00376-013-3008-0
    [7] HUANG Hong, JIANG Yongqiang, CHEN Zhongyi, LUO Jian, WANG Xuezhong, 2014: Effect of Tropical Cyclone Intensity and Instability on the Evolution of Spiral Bands, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1090-1100.  doi: 10.1007/s00376-014-3108-5
    [8] MA Zhanhong, FEI Jianfang, HUANG Xiaogang, CHENG Xiaoping, 2014: Impacts of the Lowest Model Level Height on Tropical Cyclone Intensity and Structure, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 421-434.  doi: 10.1007/s00376-013-3044-9
    [9] Zhenhua HUO, Wansuo DUAN, Feifan ZHOU, 2019: Ensemble Forecasts of Tropical Cyclone Track with Orthogonal Conditional Nonlinear Optimal Perturbations, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 231-247.  doi: 10.1007/s00376-018-8001-1
    [10] Tong XIE, Liguang WU, Yecheng FENG, Jinghua YU, 2024: Alignment of Track Oscillations during Tropical Cyclone Rapid Intensification, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 655-670.  doi: 10.1007/s00376-023-3073-y
    [11] Kexin CHEN, Guanghua CHEN, Donglei SHI, 2023: Modulation of the Wind Field Structure of Initial Vortex on the Relationship between Tropical Cyclone Size and Intensity, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1707-1721.  doi: 10.1007/s00376-023-2233-4
    [12] LI Wei-Wei, WANG Chunzai, WANG Dongxiao, YANG Lei, DENG Yi, 2012: Modulation of Low-Latitude West Wind on Abnormal Track and Intensity of Tropical Cyclone Nargis (2008) in the Bay of Bengal, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 407-421.  doi: 10.1007/s00376-011-0229-y
    [13] Jian YUE, Zhiyong MENG, Cheng-Ku YU, Lin-Wen CHENG, 2017: Impact of Coastal Radar Observability on the Forecast of the Track and Rainfall of Typhoon Morakot (2009) Using WRF-based Ensemble Kalman Filter Data Assimilation, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 66-78.  doi: 10.1007/s00376-016-6028-8
    [14] LUO Xia, FEI Jianfang, HUANG Xiaogang, CHENG Xiaoping, YU Kun, 2015: Remote Impact of Blocking Highs on the Sudden Track Reversal of Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1517-1532.  doi: 10.1007/s00376-015-4284-7
    [15] Hong HUANG, Dan WU, Yuan WANG, Zhen WANG, Yu LIU, 2024: Track-Pattern-Based Characteristics of Extratropical Transitioning Tropical Cyclones in the Western North Pacific, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-2330-4
    [16] Jiawei YAO, Wansuo DUAN, Xiaohao QIN, 2021: Which Features of the SST Forcing Error Most Likely Disturb the Simulated Intensity of Tropical Cyclones?, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 581-602.  doi: 10.1007/s00376-020-0073-z
    [17] Suk-Jin CHOI, Dong-Kyou LEE, 2016: Impact of Spectral Nudging on the Downscaling of Tropical Cyclones in Regional Climate Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 730-742.  doi: 10.1007/s00376-016-5061-y
    [18] Yanchen ZHOU, Jiuwei ZHAO, Ruifen ZHAN, Peiyan CHEN, Zhiwei WU, Lan WANG, 2021: A Logistic-growth-equation-based Intensity Prediction Scheme for Western North Pacific Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1750-1762.  doi: 10.1007/s00376-021-0435-1
    [19] Qu Xiaobo, Julian Heming, 2002: The Impact of Dropsonde Data on Forecasts of Hurricane Debby by the Meteorological Office Unified Model, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 1029-1044.  doi: 10.1007/s00376-002-0062-4
    [20] Meng Zhiyong, Chen Lianshou, Xu Xiangde, 2002: Recent Progress on Tropical Cyclone Research in China, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 103-110.  doi: 10.1007/s00376-002-0037-5

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 12 November 2015
Manuscript revised: 02 February 2016
Manuscript accepted: 07 March 2016
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Possible Sources of Forecast Errors Generated by the Global/Regional Assimilation and Prediction System for Landfalling Tropical Cyclones. Part I: Initial Uncertainties

  • 1. Laboratory of Cloud-Precipitation Physics and Severe Storms, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029
  • 2. Meteorological Research Institute of the Japan Meteorological Agency, Tsukuba 305-0052, Japan
  • 3. State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029

Abstract: This paper investigates the possible sources of errors associated with tropical cyclone (TC) tracks forecasted using the Global/Regional Assimilation and Prediction System (GRAPES). The GRAPES forecasts were made for 16 landfalling TCs in the western North Pacific basin during the 2008 and 2009 seasons, with a forecast length of 72 hours, and using the default initial conditions ("initials", hereafter), which are from the NCEP-FNL dataset, as well as ECMWF initials. The forecasts are compared with ECMWF forecasts. The results show that in most TCs, the GRAPES forecasts are improved when using the ECMWF initials compared with the default initials. Compared with the ECMWF initials, the default initials produce lower intensity TCs and a lower intensity subtropical high, but a higher intensity South Asia high and monsoon trough, as well as a higher temperature but lower specific humidity at the TC center. Replacement of the geopotential height and wind fields with the ECMWF initials in and around the TC center at the initial time was found to be the most efficient way to improve the forecasts. In addition, TCs that showed the greatest improvement in forecast accuracy usually had the largest initial uncertainties in TC intensity and were usually in the intensifying phase. The results demonstrate the importance of the initial intensity for TC track forecasts made using GRAPES, and indicate the model is better in describing the intensifying phase than the decaying phase of TCs. Finally, the limit of the improvement indicates that the model error associated with GRAPES forecasts may be the main cause of poor forecasts of landfalling TCs. Thus, further examinations of the model errors are required.

1. Introduction
  • Tropical cyclones (TCs), especially those that make landfall, have long been recognized as a cause of considerable damage and disruption in coastal areas (Webster et al., 2005; Roy and Kovordanyi, 2012). Forecasting the tracks of these storms has been a longstanding concern for meteorologists. Over recent decades, ongoing research has led to significant improvements in the forecasting of TC tracks by NWP models (Rappaport et al., 2009; Heming and Goerss, 2010). However, large track-forecast errors still occur (Ryan et al., 2015), and there have been instances where the forecast errors have exceeded 1000 km over 3 days (Yamaguchi et al., 2012). The impact of such large errors is exacerbated if they are associated with landfalling TCs. To improve disaster prevention and mitigation measures, more accurate TC track forecasts are required, especially for landfalling TCs. As the current theory of TC motion is mature and widely accepted (Chan, 2010), error diagnosis has become an important aspect of attempts to further improve TC forecasts (Galarneau and Davis, 2013).

    Forecast errors derive from both uncertainties in the state of the atmosphere at the beginning of the forecast period and model imperfections, and many studies have examined the impacts of these factors on TC forecasts (Yamaguchi and Majumdar, 2010; Brennan and Majumdar, 2011; Kepert, 2011; Green and Zhang, 2014; Galarneau and Hamill, 2015; Ryan et al., 2015). Generally, both the initial and model uncertainties have significant impacts on TC forecasts. However, in a real forecast the forecast errors usually result simultaneously from both the model and initial uncertainties. It is important to consider whether we can identify the forecast errors derived from initial errors, and those derived from the model errors. If this is possible, then we can reduce the uncertainty associated with the largest sources of error.

    Given the above, it is important to consider how best to distinguish the contributions of the initial errors from those of the model errors. This task is generally difficult, despite previous efforts. For example, (Orrell et al., 2001) and (Orrell, 2003) compared the evolution of real forecast errors with the evolution of an approximate accumulated model error to determine the impact of the model errors. They pointed out that at the start of the forecast the evolution of the model errors is similar to the evolution of the forecast errors, and therefore the forecast errors result mainly from the model errors. However, as the simulation proceeds, the approximate accumulated model error begins to underestimate the forecast error. A similar experiment was carried out by (Liu et al., 2010) to evaluate error development in the Lorenz-96 system. (Nicolis et al., 2009) classified the leading timescales for initial errors and model errors by comparing their development rates, and one of the key points in their approach was determining the crossover time at which the contribution of the initial condition errors matches that of the model errors under the assumptions of small uncorrelated initial errors and small parameter errors. (Yu et al., 2012) used the conditional nonlinear optimal perturbations of initials (CNOP-I) and parameters (CNOP-P) to identify the source of the spring predictability barrier (SPB) for El Niño events. They found that the CNOP-P errors do not cause a significant SPB for El Niño events, but that the CNOP-I errors generate a significant SPB. On the basis that model systems contain many parameters, and that it is difficult to consider all of these parameters, (Duan and Zhou, 2013) proposed a nonlinear forcing singular vector (NFSV) approach to evaluate the effect of tendency errors on predications, and the NFSV was found to be another important source of the El Niño prediction error (Duan and Zhao, 2015). A similar method was proposed by (Xu et al., 2004), who applied a nonlinear optimization method to identify whether the initial errors or the model errors are dominant. However, this method requires an adjoint model and an optimization algorithm, and is limited by the development of the adjoint model. (Yamaguchi et al., 2012) distinguished TC track prediction errors attributable to the initial conditions ("initials", hereafter) and those attributable to the NWP model by replacing the initials of the JMA Global Spectral Model with those of the ECMWF. They found that by doing so, the average position error was reduced by about 10%. In some cases the predictions were significantly improved, whereas in others the replacement of the initials did not improve the prediction even when the ECMWF forecast was accurate, thereby indicating the need to deal with the uncertainties associated with the NWP model. However, (Yamaguchi et al., 2012) used only a single NWP model, and they did not further reveal the source of the initial errors, or the source of the model errors.

    A new generation of the national operational NWP system, known as the Global/Regional Assimilation and Prediction System (GRAPES) Global Forecast System (GRAPES_ GFS), was developed by the China Meteorological Administration (CMA) to enhance meteorological services in China in the 21st century (Zhang and Shen, 2008). The GRAPES_GFS model has developed rapidly over recent years (Huang et al., 2013, 2014), and now generates reliable forecasts for the subtropical high (Chen et al., 2008). In addition, the anomaly correlation coefficients calculated based on the 500-hPa geopotential height between GRAPES_GFS hindcasts and NCEP analysis were found to be high (Zhang and Shen, 2008, Fig. 2), and it was shown that GRAPES_GFS has an effective forecast time of about 6 days. This indicates that GRAPES_GFS may be capable of forecasting the tracks of TCs well.

    In this paper, we examine the forecast ability of the GRAPES_GFS model (hereafter, GRAPES) for landfalling TCs, and identify the source of the forecast errors associated with these TCs. To isolate the two main factors that determine the forecast errors (i.e., the initial uncertainty and the model uncertainty), the present study is divided into two parts. In part I, we distinguish between the forecast errors attributable to the initial errors and those attributable to the model errors by using a similar approach to (Yamaguchi et al., 2012). Then, we focus on the TCs for which the forecast errors were caused mainly by the initial uncertainty, and determine the source of the initial errors. In part II (to be presented in a separate paper), we focus on those forecast errors that resulted mainly from model uncertainty, and we determine the source of the model errors.

    The remainder of this paper is organized as follows. Section 2 introduces the methodology, including the model designs and the approach used to distinguish the forecast errors attributable to the initial errors from those attributable to the model errors. Section 3 classifies the TC cases according to the contributions of the initial errors and model errors. Section 4 reveals the source of the initial errors by analyzing the initial uncertainties and the results of sensitivity experiments conducted to identify the important variables and areas in the initials for accurate forecasts of landfalling TCs using the GRAPES model. Finally, discussion and conclusions are presented in section 5.

2. Data and methodology
  • The GRAPES model has a horizontal resolution of 1°× 1° and 36 vertical levels. The physical parameterizations used in this analysis were as follows: the WSM6 scheme for microphysics (Hong and Lim, 2006); a modification of RRTM (Rapid Radiative Transfer Model) scheme named RRTMG for long- and shortwave radiation (Iacono et al., 2008), the common land model scheme for land surface processes (Dai et al., 2003); the MRF scheme for the PBL (Hong and Pan, 1996); and the simple Arakawa-Schubert scheme for cumulus clouds (Han and Pan, 2011). The default initials, and boundary conditions, were supplied by the NCEP FNL Operational Global Analysis (1°× 1°) interpolated into the GRAPES forecast system. Herein, the forecasts made by the GRAPES model with its default initials are denoted as GRAPES-FNL.

    According to (Yamaguchi et al., 2012), among nine forecast centers, the TC tracks over the period 2008-10 were forecast most accurately by the ECMWF for the forecast period of 1 to 5 days (see their Fig. 1). Consequently, we assumed that the model and initials used by ECMWF were both accurate, and therefore used the ECMWF initials as the candidate initials for GRAPES. The initials used by the ECMWF model can be found in the Year of Tropical Convection (YOTC) dataset. YOTC was a program launched in May 2008 by the WCRP and the World Weather Research Programme/THORPEX. Its purpose is to organize coordinated observations, modeling and forecasting of organized tropical convection, with the goal to better understand its multiscale structure and interactions (see http://www.wmo.int/pages/prog/arep/wwrp/new/yotc.html for further details of the project). The source of the ECMWF analysis data can be found from YOTC (Yamaguchi et al., 2012). We downloaded the YOTC data with a horizontal resolution of 1°, equivalent to that of GRAPES. Herein, the forecasts made by GRAPES using the ECMWF initials are denoted as GRAPES-EI. For comparison, the forecasts of the ECMWF model with its initials (denoted as ECMWF-EI) were also downloaded from the YOTC dataset with a resolution of 1°. The track and intensity observations used to test the accuracy of the model forecasts were derived from the TC best-track dataset of the Shanghai Typhoon Institute of the CMA and were downloaded from the typhoon website of the Chinese government (www.typhoon.gov.cn; Ying et al., 2014).

    Sixteen TCs that occurred in the Northwest Pacific and made landfall between 2008 and 2009 are studied. In each case the forecast period was 72 h ahead of landfall (Table 1). Some TCs were short-lived, and the initials used in these forecasts were for tropical depressions rather than tropical storm intensity.

3. Distinguishing between forecast errors caused by initial and model errors
  • Although the ECMWF model usually produces accurate forecasts, it is still necessary to examine its performance with respect to the storms under consideration here and at the chosen resolution. First, the forecasts made by GRAPES-FNL were compared with those made by ECMWF-EI, and the average forecast error of the 16 TCs obtained from ECMWF-EI was the smallest of all (Fig. 1). Figure 2 shows that the 72 h forecasts of the TC tracks from ECMWF-EI are better than those from GRAPES-FNL for all 16 TCs. In most cases, the accumulated track errors generated by ECMWF-EI are significantly smaller than those from GRAPES-FNL. Next, comparisons were made between the GRAPES-EI and GRAPES-FNL runs, which showed that the forecasts for 12 of the 16 TCs were improved in the GRAPES-EI experiment compared with GRAPES-FNL (Fig. 2). As both the ECMWF model and its initials are generally accurate, we concluded that the poor forecasts of the remaining four TCs by GRAPES were caused by model errors. However, additional analysis of the other 12 cases is required before a decision can be made on whether their poor forecasts in GRAPES-FNL were caused by initial uncertainties in the FNL data or by uncertainties within the GRAPES model. To do this, we defined an index to measure the improvement of the forecasts, as follows:

    Figure 1.  Average forecast errors (km) for the 16 TCs every 6 h. Blue line: forecast by GRAPES with default initials; red line: forecast by ECMWF with its initials; green line: forecast by GRAPES but with ECMWF initials.

    Figure 2.  Accumulated track errors (km; accumulated every 6 h over the analysis period of 72 h) for each TC from the GRAPES model with its default initials (GRAPES-FNL, blue bars) and with the ECMWF initials (GRAPES-EI, green bars), and from the ECMWF model with its initials (ECMWF-EI, red bars). The red box indicates the TCs in the much-improved group; the green box indicates the TCs in the unimproved group; and the remainder are the little-improved group.

    \begin{equation} \!{IM}\!=\!\!\dfrac{(\!E_{GRAPES\hbox{-}FNL}\!\!-\!E_{ECMWF\hbox{-}EI}\!)\!-\!(\!E_{GRAPES\hbox{-}EI}\!-\!E_{ECMWF\hbox{-}EI}\!)}{E_{GRAPES\hbox{-}FNL}\!\!-\!E_{ECMWF\hbox{-}EI}}, (1)\end{equation} where E represents the departure of the forecast from the observations (i.e., the 72 h accumulated track errors calculated every 6 hours) and the subscripts indicate the various forecast errors.

    If IM is greater than or equal to 0.5, the improvement of the initials is highly effective, as it would make the forecasts much closer to the observations. In this case, we would conclude that the initial uncertainties are the main source of the forecast error; and consequently, the initials should be improved first. If IM is less than 0.5, the improvement of the initials is less effective, and in this case the model uncertainty would be taken as the main source of the forecast uncertainty, and improvement of the model would be necessary.

    Table 1 shows that only four TCs had large IM values, whereas eight had small IM values. According to the above definition, the four TCs with large IMs are those for which forecast errors resulted from initial uncertainty, and we classified these as TCs for which the initials should be improved first (the initial-first group). The remaining eight TCs with small IMs were defined as TCs for which the model should be improved first (the model-first group). Therefore, there are a total of twelve TCs for which the model is unable to forecast the TC landfalls, and we should therefore improve the model first. It is noted that the IM value for Morakot (2009) is 79.9, which is two orders of magnitude larger than the other eleven IM values. This is due to the large improvement for Morakot (2009) when using ECMWF initials. The forecast of GRAPES-EI is now better than the forecast of ECMWF-EI in the Morakot (2009) case. As we focus on the impact of the initial uncertainty associated with the TC landfall forecasts made by the GRAPES model, we concentrate on the initial-first TCs in the remainder of this paper. However, the initials of the model-first TCs will also be analyzed for comparison. To facilitate these comparisons, the initial-first TCs were renamed as the "much-improved" group, and the model-first TCs were divided into two subgroups as follows: The eight TCs with small IMs were classified as the "little-improved" group, and the other four, which showed no improvements when using the ECMWF initials, were defined as the "unimproved" group.

    Figure 3.  Initial position errors when using the NCEP FNL dataset (the default initials of GRAPES) versus using the ECMWF YOTC dataset. The red box indicates the TCs in the much-improved group, and the remainder are the little-improved group.

    Figure 4.  Differences in initial values between the NCEP FNL dataset (the default initials of GRAPES) and the ECMWF YOTC dataset. The first row shows the differences in SLP (hPa); the second row shows the differences in geopotential height at 500 hPa (m); the third row shows the differences in the zonal wind (m s$^-1$) at 850 hPa; the fourth row shows the differences in specific humidity (g g$^-1$) at 850 hPa; and the fifth row shows the differences in temperature (K) at 700 hPa. The first column shows data for TC Hagupit (2008) (an example of the much-improved group); the second column shows data for TC Nuri (2008) (little-improved group); and the third column shows data for TC Fengshen (2008) (unimproved group).

4. Analysis of the initial uncertainty
  • In this section, the difference between the ECMWF initials and the default initials is examined for the three TC groups defined in section 3 as we try to answer the following questions: First, what will be improved in the initials when using ECMWF initials? Second, what are the differences among the initials in the three TC groups? In other words, what type of TC forecasts can be significantly improved by improving the initials? Analysis was carried out from the perspectives of the positions, intensity, and environmental features of the TCs.

    First, for those TCs whose forecasts have been improved when using ECMWF initials, we analyzed what has been improved in the initials. As it was pointed out by (Yamaguchi et al., 2012) that the initial positions of TCs may have an important impact on their track forecasts, we first examined the positions of the TCs at the initial time. Figure 3 shows that the initial position errors for three cases are larger using ECMWF initials than the default initials, while for the other cases they are comparable between the two types of initials. This indicates that the improvements in the TC track forecasts are not necessarily related to the improvement in the initial position of the TC.

    The initial intensity of the TC is another factor to which the track forecast is sensitive (Galarneau and Hamill, 2015). Thus, we focus next on the intensity of the TCs. The differences in SLP between the default initials and ECMWF initials are obvious at the TC center (Fig. 4), and the SLP from the default initials is higher than that from the ECMWF initials for all TCs, regardless of the degree of improvement. Consequently, the initial TCs simulated using the GRAPES default settings have lower intensities than those simulated by the ECMWF model. The average of the differences between the central pressures based on the default initials and those based on the ECMWF initials is positive for all three TC groups (Fig. 5). As the central pressure generated by the ECMWF initials is closer to the observations than that based on the default initials of GRAPES (Table 2), the initial intensity of the TCs simulated using GRAPES-FNL is lower than observed. Consequently, replacing the default initials with the ECMWF initials strengthens the TC intensity and improves the accuracy of TC intensity at the initial time. These conclusions are supported by our analysis of the geopotential height at 500 hPa and the zonal wind at 850 hPa (Fig. 4; second and third columns, respectively). In this case the differences between the default and the ECMWF scenarios is positive in the geopotential heights at the TC center. It is positive in the zonal wind on the north side and negative in the zonal wind on the south side of the cyclone for each TC.

    The environmental field is without doubt an important factor that influences the track of a TC. For all three TC groups, the geopotential height at 500 hPa with the GRAPES default initials was lower than that with the ECMWF initials around the subtropical high and monsoon trough, but was higher around the South Asia high (Fig. 4; second column). This indicates that, compared with those generated by the ECMWF initials, the subtropical high generated by the GRAPES initials has weaker intensities, while the South Asia high and the monsoon trough have stronger intensities. However, if we take an overall view, the difference in the TC center is the most noticeable. Although the differences in temperature and humidity between the default initials and ECMWF initials are not significant, we still find systematic differences between them. That is, the temperature generated by the GRAPES default initials was higher than that for the ECMWF initials in the ocean area, but lower than that for the ECMWF initials on land (Fig. 4; fifth column). Thus, in the TC center the temperature was slightly higher with the GRAPES default initials. However, the specific humidity was lower in the TC center but higher in the surroundings with the GRAPES default initials compared with the ECMWF initials (Fig. 4; fourth column).

    Figure 5.  (a) Average accumulated track forecast errors (km; accumulated every 6 h over the analysis period of 72 h) for each group of TCs from the GRAPES-FNL results. (b) Average initial differences between the central SLP (hPa) from the NCEP FNL data and the ECMWF YOTC data.

    Next we consider the nature of the improvements in the initials when using the ECMWF data. Generally, the intensity of the TC at the initial time is improved and strengthened when using the ECMWF dataset, but the initial positions of the TCs are not. The subtropical high is strengthened, but the South Asia high and monsoon trough are weakened when using the ECMWF data. Also, the temperature at the TC center is lower, whereas the specific humidity is higher, when using the ECMWF initials compared with the GRAPES default initials. The total effect of these factors improves the TC landfall forecasts. However, the question remains as to which factor is the most important. This is considered in section 4.2.

    It is important to consider which type of TC forecast can be most improved using the ECMWF initials. Figures 2 and 5 show that the much-improved group has relatively small forecast errors compared with the other two groups. Consequently, to achieve a significant improvement, the model should be able to predict the TC; i.e., the model error should not be too large. The differences of TC intensities between the default initials and the ECMWF initials for the much-improved group are larger than those for the other two groups (Fig. 5), and this indicates that the much-improved group has a sizeable uncertainty associated with the TC intensity. Another interesting point is that three of the four much-improved cases were in the intensifying phase and had a weaker initial intensity, whereas three of the four unimproved cases were decaying and were generally strong at the initial time (Fig. 6 and Table 2). This means that the model physics may catch up with the intensification phase of the cyclone, but fail to describe the decaying phase. Therefore, for intensifying TCs the forecasts are greatly improved when the initials are improved, but for decaying TCs the forecasts cannot be improved simply by improving the initials. An investigation of the environmental fields revealed little of interest in this context.

    In summary, given that TC forecasts made with the GRAPES model can be greatly improved by improving the initials, the TCs are expected to have a large uncertainty in initial intensity, and the TCs are in the intensifying phase with a relatively weak initial intensity.

  • In section 4.1 we demonstrated systematic differences between the GRAPES default initials and ECMWF initials, such as the intensities of the TCs, subtropical high, South Asia high, and monsoon trough. A complete replacement of the default initials with the ECMWF initials as the model input variables (wind, geopotential height, temperature, and relative humidity; note that the boundary conditions were kept the same as the default) improves the forecasts for the studied TCs. Next, we consider which variables should be replaced to yield the greatest improvement in forecasts, and which location is most effective (insofar as the replacement of initials there would obtain the greatest improvement). In other words, we assess which system should be improved at the initial time to generate the greatest improvement in TC forecasts. To address these questions we examined the impacts on the forecasts of replacing each variable, as well as the impact of the replacements in different locations. For this purpose, the four cases of the initial-first group are studied.

    Figure 6.  Observed SLP (hPa) for the (a) much-improved and (b) unimproved TCs during the study period.

    Figure 7.  Difference in geopotential height (m) at the 500-hPa level between the NCEP FNL and ECMWF YOTC datasets. The boxes indicate areas where the data are replaced. The yellow boxes indicate the area in and around the TC center. The four TCs (indicated by the titles above each panel) are from the much-improved group. (a) 2008 Hagupit; (b) 2008 Higos; (c) 2009 Ketsana; (d) 2009 Morakot.

    Figure 8.  Accumulated forecast errors (km) for the typhoon tracks predicted by the GRAPES model with its default initials (NCEP FNL data) and the GRAPES model with the ECMWF initials partly replacing the default initials. The labels h-TC, uv-TC, rh-TC, and T-TC indicate that the geopotential height, wind field, relative humidity, and temperature, respectively, of the default initials have been replaced by the ECMWF initials around the TC central area. The labels huv-TC, huv-subhigh, huv-westTC, huv-southTC, and huv-plateau indicate that the geopotential height and wind field were replaced at the center, northeast, west, south, or southwest of the TC, and the plateau area, respectively. The four TCs (indicated by the titles above each panel) are from the much-improved group.

    As described in section 4.1, the wind, geopotential height, temperature, and relative humidity at the TC center generated by the GRAPES default initials show systematic departures from those generated by the ECMWF initials. First, we respectively replaced the four variables in the default initials with the corresponding variables from the ECMWF initials in and around the TC center (yellow rectangles in Fig. 7), and ran GRAPES with the newly produced initials that combined both the default and ECMWF data. The track forecast errors of the TCs were then compared to determine which variable around the TC center is the most important in terms of improving the track forecast.

    Figure 8 shows that the track forecast errors were significantly reduced by improving the geopotential height and wind fields surrounding the TC. However, the improvements in the relative humidity and temperature surrounding the TC failed to obviously reduce the forecast errors (Fig. 8). This result further demonstrates that an accurate estimation of TC intensity at the initial time is important to improve the forecasts. The importance of the geopotential height seems comparable to that of the wind fields, and both of these factors are reflections of the TC intensity. Therefore, it is necessary to consider whether this is the main factor that drives the improvement in the forecasts.

    As indicated previously, the intensities of the subtropical high, South Asia high, and monsoon trough were changed when using the ECMWF initials. Next, we replaced the geopotential height and wind fields of the default initials in the areas where the above systems are located, with those of the ECMWF data in the corresponding areas, and examined whether this replacement improved the forecasts, and to what extent. To this end, another four equal-sized areas were selected for each TC for comparison (Fig. 7). The first location is on the Tibetan Plateau where the South Asia high is located; we refer to this as the plateau area. There are always large differences between the default initials of GRAPES and the ECMWF initials in this region, and other studies have reported that the South Asia high influences TC movements (Chen, 1965; Ding and Wright, 1983). The second area is located in the west side of the subtropical high and to the northeast of the TC. This is also an important region in terms of TC tracks, and with a relatively large difference between the default initials of GRAPES and the ECMWF initials. As the subtropical high influences TCs in this area, we denote this as the subtropical high area. The third area is located on the Indochina Peninsula and to the west of the TCs. Figure 7 shows obvious differences between the GRAPES default initials and the ECMWF initials in this region. This area is usually influenced by the monsoon trough. The fourth area is also influenced by the monsoon and is located to the south or southwest of the TCs, in the path of southwesterly winds. This region is marked by obvious differences between the GRAPES default initials and the ECMWF initials.

    Sensitivity experiments were then carried out by respectively replacing the default GRAPES initials with the ECMWF initials in the plateau area, the center of the TCs, and the areas to the northeast, west, and south of the TCs. As mentioned above, the greatest forecast improvement is achieved when both the geopotential height and wind fields are replaced, so these replacements were made simultaneously while keeping the other variables the same as in the default scenario. The results for all cases were similar; that is, the replacements at the centers of the TCs generated the largest forecast improvements (Fig. 8). However, replacing only the default initials with the ECMWF initials in the other locations had little or no effect (Fig. 4). The sensitivity experiments in these areas further confirmed that the improvement in TC intensity at the initial time is important for improvements in the TC track forecasts. Of note is that when the geopotential height and wind fields in and around the TCs are improved, not only the intensity of the TC is improved, but also the steering flows and the structures of the TCs. Many studies have pointed out that steering flows and TC structure have an important influence on TC motion (McTaggart-Cowan et al., 2006; Brennan and Majumdar, 2011; Galarneau and Davis, 2013). This may explain why an improvement in and around the TC center results in an improvement in track forecasts.

    Generally, all of the experiments showed that a realistic estimation of TC intensity at the beginning of the forecast is important if the GRAPES model is to generate accurate forecasts of TC landfall tracks.

5. Discussion and conclusions
  • This paper (part I) and the following paper (part II) focus on determining the source of the forecast errors for TC landfalls made with the GRAPES model. First, the method proposed by (Yamaguchi et al., 2012) was used to distinguish the forecast errors that are attributable to initial errors from those attributable to model errors. Next, attention was paid to cases in which the forecast errors were attributable to the initial errors. An assessment of the differences in the initials between the default forecasts and the improved forecasts revealed systematic differences in the initials. Further comparisons were made between the initials of the TCs whose forecast errors were attributable to the initial errors, and those whose forecast errors were attributable to the model errors, in order to determine the best way to significantly improve the forecasts by improving only the initials. Then, sensitivity experiments were carried out to identify the most important variables and areas at the beginning of the forecasts.

    Of the sixteen TCs studied here, all of which made landfall during 2008 and 2009 in the Northwest Pacific, four were identified as TCs whose forecast errors were attributable to the initial errors, and twelve as TCs whose forecast errors were attributable to the model errors. Further examination showed obvious systematic differences between the default initials of the GRAPES model and the ECMWF initials for all the cases. That is, both the TCs and the subtropical high based on the default initials were weaker than the observations and weaker than those based on the ECMWF initials. However, the intensity of the South Asia high and the monsoon trough based on the default initials were stronger than those based on the ECMWF initials. Compared with the ECMWF initials, the default initials in GRAPES had a higher temperature and lower specific humidity at the TC center. The TCs whose forecast errors were attributable to the initial errors had larger initial uncertainties in intensity compared with those attributable to the model errors, and they were usually in the intensifying phase with a relatively weak intensity at the initial time. Sensitivity experiments demonstrated that the accurate estimation of TC intensity at the beginning of the forecast is most important if the GRAPES model is to generate reliable forecasts of the tracks of landfalling TCs.

    This study has also confirmed previous findings that the initial values of the wind and pressure fields associated with a TC are important for its track forecast, and not only the subtropical high and the midlatitude trough. In addition, we have demonstrated that when the initial uncertainty (comparing the initials among several NWP centers) is large, the improvement in the initial values may be more beneficial to the forecast than when the initial uncertainty is small. This result indicates that data assimilation may be more useful when the initial uncertainty is large. Generally, the above results indicate when and where the data assimilation is needed and what should be assimilated for good forecasts of landfalling TCs made using the GRAPES model.

    Finally, as noted above, twelve cases were identified as TCs whose forecast errors were attributable to model errors. Thus, the GRAPES model has defects in simulating the motion of TCs. In our companion paper (part II), we seek to determine the source of the model errors and suggest how future versions of the model may be improved.

    The GRAPES model used in this paper has a very coarse resolution of 1°× 1°. As reported by (Fiorino, 2009), based on the ECMWF IFS, simply increasing the model resolution does not lead to significant improvement in the forecast accuracy; rather, large improvement is related more to upgrading the model, including its physics. Thus, we hypothesize that the current conclusion that the model itself plays an important role in improving the forecasts will remain valid even with a high-resolution model of GRAPES. In addition, it is important to note that TC intensity forecasting will continue to be a big challenge for TC research and forecasting communities, even if it is possible to use high-resolution NWP systems with a horizontal resolution of around 2 km.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return