Sensitivity of WRF model in simulating wind zonal and meridional components in Tehran

Document Type : Research Article

Authors

1 Ph.D. Student of Climatology, Faculty of Geography, University of Tehran, Tehran, Iran

2 Associate Professor of Climatology, Faculty of Geography, University of Tehran, Tehran, Iran

3 Professor of Climatology, Faculty of Geography, University of Tehran, Tehran, Iran

Abstract

The wind is one of the main factors in determining the weather condition and the daily air quality of urban spaces depends on the wind. Therefore, to achieve the dominant behavioral patterns of wind direction and intensity, various simulation models are used. This study considers the weather research and forecasting model (WRF). To simulate wind zonal and meridional components, the role of ECMWF (ERA5) and NCEP (FNL) boundary condition data with 7 physical schemas (Exp1 to Exp7) in two modes: 1) with default static data and 2) with high-resolution static data DEM (Aster satellite image with a spatial resolution of 30 m) instead of the default data with a spatial resolution of approximately 1 km, the land use/cover of Copernicus with a resolution of 100 m instead of the modis data with a spatial resolution of approximately 500 m to 1 km) for January, May, July, and October 2018 was evaluated as representative of seasons. Observational data of wind direction and speed with a 3-hour UTC scale in 2018 for 4 synoptic stations of Mehrabad, Chitgar, Geophysics, and Shemiran were obtained from the Meteorological Organization, and by applying 180 degrees to the weather direction, the vector orientation was obtained. Then, using the Rewind plugin in the developed R software, the zonal and meridian components of observational wind were calculated. By examining the correlation coefficient of wind zonal component in 4 selected stations in both default and high-resolution conditions with ERA5 and FNL boundary conditions, Shemiran station had the lowest correlation in January, May, and October, while Chitgar station on the suburban of the city had the weakest correlation in July; However, Shemiran station had the strongest correlation in July by a significant margin. Due to the location of the Shemiran station, the main reason for the weakness and strength of the simulations is the topography and height of this station. The results show that the simulation of the wind zonal component, except July in most schemas with the default static data and high resolution, is much better than the meridional component. Proper configuration of schemas, selection of ideal boundary conditions, determination of appropriate spatial resolution, and replacement of static data with a high resolution instead of default data can bring the model simulation much closer to the observation data. According to the results of the average output of 4 correlations, bias, mean square error, and mean absolute error statistics, considering the same weight coefficient of each statistic in Mehrabad, Geophysics, Shemiran, and Chitgar stations for each schema, FNL boundary conditions with default static data for the component Wind except July, and ERA5 boundary conditions were selected as the best boundary conditions for the meridional wind component except January with the best performance. Among the seven schemas tested for zonal and meridional components under the ERA5 and FNL boundary conditions with default and high-resolution data, Exp (1) (YSU) and Exp (6) (ACM2) generally yielded the best results for Tehran wind simulation.

Keywords

Main Subjects


Bahmanzadeh, F., Qader, S., HaghShenas, S.A., and Yazji, D., 2019, Investigating the performance of the WRF model to predict the field wind at 10 meters level and two meters level location using satellite data and synoptic stations in the region of Oman Sea and Arabian Sea. Earth and Space Physics, 45(2), 458-441.
Belmonte Rivas, M., Stoffelen, A., 2019, Characterizing ERA-Interim and ERA5 surface wind biases using ASCAT. Ocean Sci. 15 (3), 831–852.
Carvalho, D.; Rocha, A., Gómez-Gesteira, M., and Santos, C.A., 2012, Sensitivity study of the WRF model in wind simulation for an area of high wind energy: Environmental Modelling & Software. 33, 23–34.
Carvalho, D., Rocha, A., Gómez-Gesteira, M., Silva Santos, C., 2014, WRF wind simulation and wind energy production estimates forces by different reanalyses: Comparison with observed data for Portugal: Applied Energy, 117, 116–126.
Chou M.-D., and . Suarez, M.J. 1994, An efficient thermal infrared radiation parameterization for use in general circulation models: NASA Technical Memorandum, 104606, 3, 85pp.
Draxl, C., Hahmann, A.N., Peña, A., Giebel, G. 2014, Evaluating winds and vertical wind shear from weather research and forecasting model forecasts using seven planetary boundary layer schemes: Wind Energy  17, 39–56.
Dvorak, M.J., Archer, C.L., and Jacobson, M.Z. 2010, California offshore wind energy potential. Renew: Energy  35, 1244–1255.
Everitt, B.S., and Skrondal, A., 2010, The Cambridge Dictionary of Statistics, the fourth edition, Cambridge University Press. 55.
Fernandez-Gonz´alez, S., Martín, M.L., García-Ortega, E., Merino, A., Lorenzana, J., S´anchez, J.L., Valero, F., and Rodrigo, J.S., 2018, Sensitivity analysis of the WRF model: wind-resource assessment for complex terrain: Journal of Applied Meteorology and Climatology, 57 (3), 733–753.
Giannakopoulou, E.-M., and Nhili, R., 2014, WRF model methodology for offshore wind energy applications: Advances in Meteorology, 2014, 319819.
Graham, R.M., Hudson, S.R., and Maturilli, M., 2019, Improved performance of ERA5 in arctic gateway relative to four global atmospheric reanalyses: Geophysical Research Letters, 46 (11), 6138–6147.
Grell, G.A., and Freitas, S.R., 2014, A scale and aerosol aware stochastic convective parameterization for weather and air quality modeling: Atmospheric Chemistry and Physics, 14, 5233-5250.
Hersbach, H., Bell, B., Berrisford, P., Hirahara, S., Horanyi, ´ A., Munoz-Sabater, ˜ J., Nicolas, J., Peubey, C., Radu, R., Schepers, D., Simmons, A., Soci, C., Abdalla, S., Abellan, X., Balsamo, G., Bechtold, P., Biavati, G., Bidlot, J., Bonavita, M., De Chiara, G., Dahlgren, P., Dee, D., Diamantakis, M., Dragani, R., Flemming, J., Forbes, R., Fuentes, M., Geer, A., Haimberger, L., Healy, S., Hogan, R., Holm, ´ E., Janiskova, ´ M., Keeley, S., Laloyaux, P., Lopez, P., Lupu, C., Radnoti, G., de Rosnay, P., Rozum, I., Vamborg, F., Villaume, S., Th´epaut, J.-N., 2020, The ERA5 global reanalysis. Q. J. R. Meteorol. Soc. 146 (730), 1999–2049.
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. 
Hong, S.Y., and Lim, J.O., 2006, The WRF single–moment 6–class microphysics scheme (WSM6). J. Korean Meteor. Soc., 42, 129–151.
Hong, S.Y., Yign, N., and Dudhia, J., 2006, A new vertical diffusion package with an Explicit treatment of entrainment processes. Monthly Weather Review, 134, 2318–2341. 
Janjic, Z.I., 1994, The Step–Mountain Eta Coordinate Model: Further developments of the convection, viscous sublayer, and turbulence closure schemes. Mon. Wea. Rev., 122, 927–945.
 Kotz, S., 2006, Encyclopedia of Statistical Sciences: John Wiley & Sons, Inc. NewYork.
Men´endez, M., Tom´ as, A., Camus, P., García-Díez, M., Fita, L., Fern´ andez, J., M´endez, F. J., Losada, I. J., 2011, A methodology to evaluate regional-scale offshore wind energy resources. In: OCEANS 2011. IEEE, Spain, pp. 1–8.
Mlawer, E.J., Steven. J., Taubman, P., Brown, D., Iacono, M.J. and Clough, S.A., 1997, Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated–k model for the longwave: Journal of Geophysical Research, 102, 16663–16682. 
Mughal, M. O., Lynch, M., Yu, F., McGann, B., Jeanneret, F., Sutton, J., 2017, Wind modelling, validation and sensitivity study using Weather Research and Forecasting model in complex terrain. Environ. Model. Software 90, 107–125.
Muñoz‐Esparza, D., Lundquist, J. K., Sauer, J. A., Kosović, B. and Linn, R. R., 2017, Coupled mesoscale‐LES modeling of a diurnal cycle during the CWEX‐13 field campaign: From weather to boundary‐layer eddies. Journal of Advances in Modeling Earth Systems, 9(3), 1572-1594.
Nakanishi, M. and Niino, H., 2006, An improved Mellor–Yamada level 3 model: its numerical stability and application to a regional prediction of advecting fog. Bound. Layer Meteor. 119, 397–407. 
Nistani, A., and Qader, S., and Mohib al-Hajjah, A., 2017, Application of data mining in WRF model to simulate precipitation caused by a synoptic system in western Iran. Iranian Journal of Geophysics, 11, 101-123.
Nunalee, C.G., and Basu, S., 2013, Mesoscale modeling of coastal low-level jets: Implications for offshore wind resource estimation. Wind Energy,  17, 1199–1216.
Olauson, J., 2018, ERA5: the new champion of wind power modelling? renewable energy, 126, 322–331.
Penchah, M.M., Malakooti, H., and Satkin, M., 2017, Evaluation of planetary boundary layer simulations for wind resource study in East of Iran: renewable energy, 111, 1–10.
Pleim, J.E., 2007, A Combined Local and Nonlocal Closure Model for the Atmospheric Boundary Layer. Part I: Model Description and Testing: Journal of Applied Meteorology and Climatology, 46, 1383–1395.
Puliafito, S.E., Allende, D.G., Mulena, C.G., Cremades, P. and Lakkis, S.G., 2015, Evaluation of the WRF model configuration for Zonda wind events in a complex terrain: Atmospheric Research, 166, 24-32.
Ramon, J., Lledo, L., Torralba, V., Soret, A., and Doblas-Reyes, F.J., 2019, What global reanalysis best represents near-surface winds? Quarterly Journal of the Royal Meteorological Society, 145 (724), 3236–3251.
Salkind, N.J., 2010, Encyclopedia of research design (Vols. 1-0). Thousand Oaks, CA: SAGE Publications, Inc.
Santos-Alamillos, F. J., Pozo-Vázquez, D., Ruiz-Arias, J. A., Lara-Farego, V., Tovar-Pescador, J., 2013, Analysis of WRF model wind estimate sensitivity to physics parameterization choice and terrain representation in Andalusia (Southern Spain): Journal of Applied Meteorology and Climatology, 52, 1592–1609.
Solbakken, K., Birkelund, Y. and Samuelsen, E. M., 2021, Evaluation of surface wind using WRF in complex terrain: Atmospheric input data and grid spacing: Environmental Modelling & Software, 145, 105182.
Storm, B., and Basu, S., 2010, The WRF model forecast-derived low-level wind shear climatology over the United States Great Plains: Energies  3, 258–276.
Skamarock, W.C. 2004, Evaluating mesoscale NWP models using kinetic energy spectra: Monthly Weather Review, 132, 3019–3031.
Tewari, M., Chen, F., Wang, W., Dudhia, J., LeMone, M.A., Mitchell, K., Ek, M., Gayno, G., Wegiel, J. .and Cuenca, R.H., 2004, Implementation and verification of the unified NOAH land surface model in the WRF model: 20th conference on weather analysis and forecasting/16th conference on numerical weather prediction, pp. 11–15.
Valkonen, T., Stoll, P., Batrak, Y., Køltzow, M., Schneider, T.M., Stigter, E. E., Aashamar, O. B., Støylen, E., Jonassen, M.O. 2020, Evaluation of a sub-kilometre NWP system in an Arctic fjord-valley system in winter. Tellus Dyn. Meteorol. Oceanogr. 72 (1), 1–21.
Vishnu, S. and Francis, P. A., 2014, Evaluation of high-resolution WRF model simulations of surface wind over the west coast of India. Atmospheric and Oceanic Science Letters, 7(5), 458-463.
Wen, M., Yang, S., Vintzileos, A., Higgins, W. and Zhang, R., 2012, Impacts of model resolutions and initial conditions on predictions of the Asian summer monsoon by the NCEP Climate Forecast System. Weather Forecast. 27, 629–646.
Wang, J. and Wang, H.J., 2013, Forecasting of wind speed in Rudong, Jiangsu province by the WRF model, Climatic Environ. Res. (in Chinese), 18(2), 145–155.