Skip to main content

Advertisement

Log in

Atmospheric Flow over a Mountainous Region by a One-Way Coupled Approach Based on Reynolds-Averaged Turbulence Modelling

  • Article
  • Published:
Boundary-Layer Meteorology Aims and scope Submit manuscript

Abstract

The atmospheric flow over a mountainous region has been simulated using a model-chain approach, whereby the flow in a larger region was simulated using a mesoscale model with three nesting levels, down to a 3-km horizontal resolution, within which a fourth nesting level was set with a microscale flow solver and a domain with varying horizontal resolution, around 300 m at the site of interest. Two periods in the summer (July) and autumn (November–December) 2005, each with a duration of two weeks, were selected to test the present approach. Two sites were chosen, comprising a total of seven meteorological masts with wind vanes and anemometers at two heights. The microscale solver improved the wind-speed prediction of the mesoscale model in 10 of the 14 anemometers and replicated the high wind speeds, which were under-predicted in the mesoscale model. The wind conditions in summer varied with the daily cycle, related to regional-scale sea breezes and their interaction with local circulations induced by the topography. Regarding the turbulence intensity, the predicted decay with wind-speed increase was in agreement with the measurements. This study shows the need of both models: the microscale model captures the details of the boundary-layer physics, which would not be possible without the boundary conditions provided by the mesoscale model.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Abiven C (2013) Mesoscale & CFD coupling for wind energy yield prediction. In: Proceedings of the EWEA annual conference and exhibition, vol 2. European Wind Energy Association, Vienna, pp 1322–1325

  • Ayotte KW (1997) Optimization of upstream profiles in modelled flow over complex terrain. Boundary-Layer Meteorol 83:285–309. doi:10.1023/A:1000229609412

    Article  Google Scholar 

  • Baker D, Downs T, Ku M, Hao W, Sistia G, Kiss M, Johnson M, Brown D (2010) Sensitivity testing of WRF physics parameterizations for meteorological modeling and protocol in support of regional SIP air quality modeling in the OTR. Technical report. Ozone Transport Commission, Modeling Committee, Washington, DC, 73 pp

  • Bechmann A, Sørensen NN, Berg J, Mann J, Réthoré PE (2011) The Bolund experiment, part II: blind comparison of microscale flow models. Boundary-Layer Meteorol 141(2):245–271. doi:10.1007/s10546-011-9637-x

    Article  Google Scholar 

  • Beljaars ACM, Walmsley JL, Taylor PA (1987) A mixed spectral finite-difference model for neutrally stratified boundary-layer flow over roughness changes and topography. Boundary-Layer Meteorol 38:273–303. doi:10.1007/BF00122448

    Article  Google Scholar 

  • Castro FA, Palma JMLM, Silva Lopes A (2003) Simulation of the Askervein flow. Part 1: Reynolds averaged Navier–Stokes equations (\(k\)\(\epsilon \) turbulence model). Boundary-Layer Meteorol 107:501–530. doi:10.1023/A:1022818327584

    Article  Google Scholar 

  • Castro FA, Silva Santos CM, Palma JMLM (2008) Parallelisation of the CFD code of a CFD–NWP coupled system for the simulation of atmospheric flows over complex terrain. In: High performance computing for computational science—VECPAR 2008. Springer, pp 27–38. doi:10.1007/978-3-540-92859-1_4

  • Dahl N, Xue H, Hu X, Xue M (2015) Coupled fire-atmosphere modeling of wildland fire spread using DEVS-FIRE and ARPS. Nat Hazards 7(2):1013–1035. doi:10.1007/s11069-015-1640-y

    Article  Google Scholar 

  • Dudhia J (1989) Numerical study of convection observed during the winter monsoon experiment using a mesoscale two-dimensional model. J Atmos Sci 46(20):3077–3107. doi:10.1175/1520-0469(1989)046<3077:NSOCOD>2.0.CO;2

  • Durran DR, Klemp JB (1983) A compressible model for the simulation of moist mountain waves. Mon Weather Rev 111(12):2341–2361. doi:10.1175/1520-0493(1983)111<2341:ACMFTS>2.0.CO;2

  • Duynkerke PG (1987) A model for the turbulent structure of the stratocumulus-topped atmospheric boundary layer. J Atmos Sci 44:43–64. doi:10.1175/1520-0469(1987)044<0043:AMFTTS>2.0.CO;2

  • Duynkerke PG (1988) Application of the \(E-\epsilon \)2.0.CO;2. doi:10.1175/1520-0469(1988)045<0865:AOTTCM>2.0.CO;2

  • Dyer AJ (1974) A review of flux–profile relationships. Boundary-Layer Meteorol 7:363–372. doi:10.1007/BF00240838

    Article  Google Scholar 

  • Eidsvik KJ, Holstad A, Lie I, Utnes T (2004) A prediction system for local wind variations in mountainous terrain. Boundary-Layer Meteorol 112:557–586. doi:10.1023/B:BOUN.0000030561.25252.9e

    Article  Google Scholar 

  • Ely A, Stuart P, Zhu M, Palma JMLM, Veiga Rodrigues C, Chertovskih R (2012) Mesoscale and CFD coupling: an improved technique for predicting microscale wind. In: Proceedings of the EWEA annual conference and exhibition, vol 1. European Wind Energy Association, Copenhagen, pp 575–584

  • Fernando H (2010) Fluid dynamics of urban atmospheres in complex terrain. Annu Rev Fluid Mech 42(1):365–389. doi:10.1146/annurev-fluid-121108-145459

    Article  Google Scholar 

  • Fletcher CAJ (1991) Computational techniques for fluid dynamics. 1. Fundamental and general techniques, 2nd edn. Springer, New York, 409 pp

  • Franke J, Hellsten A, Schlünzen H, Carissimo B (2007) Best practice guideline for the CFD simulation of flows in the urban environment. In: COST Action 732, quality assurance and improvement of microscale meteorological models. Meteorological Institute, Centre for Marine and Atmospheric Sciences, University of Hamburg, 52 pp

  • Garratt JR (1992) The atmospheric boundary layer. Cambridge University Press, Cambridge, UK, 316 pp

  • Hu XM, Nielsen-Gammon JW, Zhang F (2010) Evaluation of three planetary boundary layer schemes in the WRF model. J Appl Meteorol Climatol 49:1831–1844. doi:10.1175/2010JAMC2432.1

    Article  Google Scholar 

  • IEC (2003) Wind turbine generator systems—part 1: design requirements, 3rd edn. International Standard IEC 61400-1:2003. International Electrotechnical Commission, Geneva

  • Jackson PS, Hunt JCR (1975) Turbulent wind flow over a low hill. Q J R Meteorol Soc 101:929–955. doi:10.1002/qj.49710143015

    Article  Google Scholar 

  • Jiménez PA, Dudhia J (2013) On the ability of the WRF model to reproduce the surface wind direction over complex terrain. J Appl Meteorol Climatol 52:1610–1617. doi:10.1175/JAMC-D-12-0266.1

    Article  Google Scholar 

  • Kristóf G, Rácz N, Balogh M (2009) Adaptation of pressure based CFD solvers for mesoscale atmospheric problems. Boundary-Layer Meteorol 131(1):85–103. doi:10.1007/s10546-008-9325-7

    Article  Google Scholar 

  • Kunz R, Khatib I, Moussiopoulos N (2000) Coupling of mesoscale and microscale models—an approach to simulate scale interaction. Environ Model Softw 15(6–7):597–602. doi:10.1016/S1364-8152(00)00055-4

    Article  Google Scholar 

  • Leonard BP (1979) A stable and accurate convection modelling procedure based on quadratic upstream interpolation. Commun Methods Appl Mech Eng 19(1):59–98. doi:10.1016/0045-7825(79)90034-3

    Article  Google Scholar 

  • Mahrt L (2011) Surface wind direction variability. J Appl Meteorol Climatol 50(1):144–152. doi:10.1175/2010JAMC2560.1

    Article  Google Scholar 

  • Muñoz-Esparza D, Kosović B, Mirocha J, van Beeck J (2014) Bridging the transition from mesoscale to microscale turbulence in numerical weather prediction models. Boundary-Layer Meteorol 153:409–440. doi:10.1007/s10546-014-9956-9

    Article  Google Scholar 

  • Mirocha J, Kirkil G, Bou-Zeid E, Chow FK, Kosović B (2013) Transition and equilibration of neutral atmospheric boundary layer flow in one-way nested large-eddy simulations using the Weather Research and Forecasting model. Mon Weather Rev 141(3):918–940. doi:10.1175/MWR-D-11-00263.1

    Article  Google Scholar 

  • Mlawer EJ, Taubman SJ, Brown PD, Iacono MJ, Clough SA (1997) Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J Geophys Res 102(D14):16,663–16,682. doi:10.1029/97JD00237

    Article  Google Scholar 

  • Murphy AH (1988) Skill scores based on the mean square error and their relationships to the correlation coefficient. Mon Weather Rev 116:2417–2424. doi:10.1175/1520-0493(1988)116<2417:SSBOTM>2.0.CO;2

  • NCEP FNL ds0832 (1999–2012) FNL operational model global tropospheric analyses, ID: ds083.2. National Centers for Environmental Prediction (NCEP). http://rda.ucar.edu/datasets/ds083.2/

  • NIMA (2000) Department of Defense World Geodetic System 1984: its definition and relationships with local geodetic systems. Technical Report. NIMA TR8350.2, National Imagery and Mapping Agency, St. Louis, Missouri, p 174

  • Otte TL (2008) The impact of nudging in the meteorological model for retrospective air quality simulations. Part I: evaluation against national observation networks. J Appl Meteor Climatol 47:1853–1867. doi:10.1175/2007JAMC1790.1

    Article  Google Scholar 

  • Palma JMLM, Castro FA, Ribeiro LF, Rodrigues AH, Pinto AP (2008) Linear and nonlinear models in wind resource assessment and wind turbine micro-siting in complex terrain. J Wind Eng Ind Aerodyn 96:2308–2326. doi:10.1016/j.jweia.2008.03.012

    Article  Google Scholar 

  • Patankar SV (1980) Numerical heat transfer and fluid flow. Hemisphere Publishing Corporation, Washington DC, 214 pp

  • Pleim JE (2007) A combined local and nonlocal closure model for the atmospheric boundary layer. Part I: model description and testing. J Appl Meteorol Climatol 46:1383–1395. doi:10.1175/JAM2539.1

    Article  Google Scholar 

  • Rhie CM, Chow WL (1983) Numerical study of the turbulent flow past an airfoil with trailing edge separation. AIAA J 21(11):1525–1532. doi:10.2514/3.8284

    Article  Google Scholar 

  • Richards PJ, Norris SE (2011) Appropriate boundary conditions for computational wind engineering models revisited. J Wind Eng Ind Aerodyn 99(4):257–266. doi:10.1016/j.jweia.2010.12.008

    Article  Google Scholar 

  • Shaw WJ, Lundquist JK, Schreck SJ (2009) Research needs for wind resource characterization. Bull Am Meteorol Soc 90:535–538. doi:10.1175/2008BAMS2729.1

    Article  Google Scholar 

  • Skamarock WC, Klemp JB, Dudhia J, Gill DO, Barker DM, Duda MG, Huang XY, Wang W, Powers JG (2008) A description of the advanced research WRF version 3. Technical Note NCAR/TN-475+STR. National Center for Atmospheric Research, Boulder, CO, 113 pp

  • Thiébaux J, Rogers E, Wang W, Katz B (2003) A new high-resolution blended real-time global sea surface temperature analysis. Bull Am Meteorol Soc 84:645–656

    Article  Google Scholar 

  • Troen I, Petersen EL (1989) European wind atlas. Risø National Laboratory, Roskilde, Denmark, 656 pp

  • van de Wiel BJH, Moene AF, Steeneveld GJ, Hartogensis OK, Holtslag AAM (2007) Predicting the collapse of turbulence in stably stratified boundary layers. Flow Turbul Combust 79:251–274. doi:10.1007/s10494-007-9094-2

    Article  Google Scholar 

  • Veiga Rodrigues C (2013) Computer modelling of atmospheric flows in wind energy applications. PhD thesis, Faculdade de Engenharia, Universidade do Porto, Portugal

  • Veiga Rodrigues C, Palma JMLM (2014) Estimation of turbulence intensity and shear factor for diurnal and nocturnal periods with an URANS flow solver coupled with WRF. J Phys Conf Ser 524(012115):10 pp. doi:10.1088/1742-6596/524/1/012115

  • Veiga Rodrigues C, Matos JC, Paiva LT, Palma JMLM (2010) Analysis of the similarity in turbulence intensity and wind shear as function of the wind velocity: field measurements and numerical results. In: Proceedings of the European wind energy conference and exhibition, vol 4. European Wind Energy Association, Warsaw, Poland, pp 2742–2751

  • Veiga Rodrigues C, Cardoso R, Soares P, Miranda PMA, Palma JMLM (2011) Downscaling flow conditions for wind farms using a dynamical coupling between a microscale CFD code and a regional mesoscale model. Presentation at the EERA Workshop on Wind Conditions, Porto, Portugal. http://indico.conferences.dtu.dk/contributionDisplay.py?sessionId=2&contribId=26&confId=70

  • Warner TT (2011) Numerical weather and climate prediction. Cambridge University Press, Cambridge, UK 550 pp

  • Wyngaard JC (2004) Toward numerical modeling in the “terra incognita”. J Atmos Sci 61(14):1816–1826. doi:10.1175/1520-0469(2004)061<1816:TNMITT>2.0.CO;2

  • Xie ZT (2011) Modelling street-scale flow and dispersion in realistic winds-towards coupling with mesoscale meteorological models. Boundary-Layer Meteorol 141(1):53–75. doi:10.1007/s10546-011-9629-x

    Article  Google Scholar 

  • Yamada T, Koike K (2011) Downscaling mesoscale meteorological models for computational wind engineering applications. J Wind Eng Ind Aerodyn 99:199–216. doi:10.1016/j.jweia.2011.01.024

    Article  Google Scholar 

  • Zhou B, Chow FK (2014) Nested large-eddy simulations of the intermittently turbulent stable atmospheric boundary layer over real terrain. J Atmos Sci 71(3):1021–1039. doi:10.1175/JAS-D-13-0168.1

    Article  Google Scholar 

Download references

Acknowledgments

The authors are grateful to the Institute of Science and Innovation in Mechanical and Industrial Engineering (INEGI) and EEVM, Empreendimentos Eólicos da Espiga e Ventominho S.A. for allowing the acess to the wind resource campaign data. This study benefited from the financial support of Fundação para a Ciência e Tecnologia (PEst-OE/EME/UI0679/2014 research project) and Programa Operacional Regional do Norte (ON.2 — O Novo Norte), under Quadro de Referência Estratégico Nacional (QREN), by Fundo Europeu de Desenvolvimento Regional (FEDER). This work was partly developed under research contract with Renewable Energy Systems Ltd and Natural Power Consultants Ltd.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to C. Veiga Rodrigues.

Appendices

Appendix 1: Surface-Layer Stability Functions and Flow Near Walls

The vertical profiles of velocity and temperature across the surface layer are functions of stability, \(\zeta \), defined as,

$$\begin{aligned} \zeta =\dfrac{z}{\mathcal {L}} =-\dfrac{g}{\varTheta }\dfrac{\kappa \, z\,q_{\mathrm {w}}}{\rho _{\mathrm {w}}\,c_\mathrm{p}\,u_*^3}, \end{aligned}$$
(22)

where z is the distance from the ground, \(\mathcal {L}\) is the Obukhov length, \(\rho _{\mathrm {w}}\) and \(q_{\mathrm {w}}\) are the fluid density and the surface heat flux, \(u_*\) is the friction velocity and \(\varTheta \) is a temperature representative of the surface layer, taken as the wall value of \(\theta _{\textsc {h}}\).

The stability functions are formulated as in Dyer (1974). Considering horizontal flow aligned with the longitudinal direction, we have

(23)
(24)

with the constants \(b_\mathrm{mu}=b_\mathrm{hu}=16\) and \(b_\mathrm{ms}=b_\mathrm{hs}=5\), \(\kappa =0.4\) and the turbulent Prandtl number at neutral stratification \(\sigma _{\textsc {n}}=1\).

For a surface layer characterized by constant fluxes of momentum and heat, the velocity and temperature, obtained by integration of \(\phi _\mathrm{m}\) and \(\phi _\mathrm{h}\), are

$$\begin{aligned} \overline{u}\,(z,\mathcal {L})&= \dfrac{u_*}{\kappa }\left[ \ln \left( \dfrac{z}{z_{\mathrm{m}0}}\right) -\psi _\mathrm{m}\left( \dfrac{z}{\mathcal {L}}\right) +\psi _\mathrm{m}\left( \dfrac{z_{\mathrm{m}0}}{\mathcal {L}}\right) \right] , \end{aligned}$$
(25)
$$\begin{aligned} \Delta \overline{\theta }\,(z,\mathcal {L})&= \dfrac{\sigma _{\textsc {n}}\,\theta _*}{\kappa }\left[ \ln \left( \dfrac{z}{z_{\mathrm{h}0}}\right) -\psi _\mathrm{h}\left( \dfrac{z}{\mathcal {L}}\right) +\psi _\mathrm{h}\left( \dfrac{z_{\mathrm{h}0}}{\mathcal {L}}\right) \right] , \end{aligned}$$
(26)

where \(z_{\mathrm{m}0}\) and \(z_{\mathrm{h}0}\) are roughness lengths for momentum and heat, with \(z_{\mathrm{h}0} \approx 0.1\,z_{\mathrm{m}0}\) (Garratt 1992). The potential temperature difference, \(\Delta \overline{\theta }\), is relative to the respective value at the wall and the turbulence temperature scale, \(\theta _*\), is defined such that \(q_{\mathrm {w}}=-\rho _{\mathrm {w}}\,c_\mathrm{p}\,u_*\,\theta _*\). Functions \(\psi _\mathrm{m}\) and \(\psi _\mathrm{h}\) are defined as,

$$\begin{aligned} \psi _\mathrm{m}\left( \zeta \right)= & {} {\left\{ \begin{array}{ll} \ln \left( \dfrac{[1+\eta ^2][1+\eta ]^2}{8}\right) -2\,\mathrm {atan}\left( \eta \right) +\dfrac{\pi }{2}, &{} \text {for}\; \zeta < 0,\\ -b_\mathrm{ms}\,\zeta ,&{} \text {for}\; \zeta \ge 0, \end{array}\right. } \end{aligned}$$
(27)
$$\begin{aligned} \psi _\mathrm{h}\left( \zeta \right)= & {} {\left\{ \begin{array}{ll} 2\,\ln \left( \dfrac{1+\chi }{2}\right) ,&{}\text {for}\; \zeta < 0,\\ -\dfrac{b_\mathrm{hs}}{\sigma _{\textsc {n}}}\,\zeta , &{}\text {for}\; \zeta \ge 0, \end{array}\right. } \end{aligned}$$
(28)

with and . From (25) and (26), the bulk transfer coefficients for momentum and heat, \(C_\mathrm{m}\) and \(C_\theta \), are defined as

$$\begin{aligned} C_\mathrm{m}\left( z,\mathcal {L}\right)= & {} \dfrac{u_*}{\overline{u}} =\kappa \,\left[ \ln \left( \dfrac{z}{z_{\mathrm{m}0}}\right) -\psi _\mathrm{m}\left( \dfrac{z}{\mathcal {L}}\right) +\psi _\mathrm{m}\left( \dfrac{z_{\mathrm{m}0}}{\mathcal {L}}\right) \right] ^{-1}, \end{aligned}$$
(29)
$$\begin{aligned} C_\theta \left( z,\mathcal {L}\right)= & {} \dfrac{\theta _*}{\overline{\Delta \theta }} =\dfrac{\kappa }{\sigma _{\textsc {n}}} \, \left[ \ln \left( \dfrac{z}{z_{\mathrm{h}0}}\right) -\psi _\mathrm{h}\left( \dfrac{z}{\mathcal {L}}\right) +\psi _\mathrm{h}\left( \dfrac{z_{\mathrm{h}0}}{\mathcal {L}}\right) \right] ^{-1}. \end{aligned}$$
(30)

1.1 Wall Laws from Known Velocity and Temperature Profiles

For microscale grid nodes below the first vertical mesoscale level, tri-linear interpolation from the \(\text {WRF}\) model results is identical to assume that velocity and temperature profiles obey a linear law. To use logarithmic laws instead, the Obukhov length \(\mathcal {L}\) should be consistent with (25) and (26) to compute both \(u_*\) and \(\theta _*\). From the definition of the bulk Richardson number and Eqs. 2229 and 30,

$$\begin{aligned} \dfrac{g}{\varTheta } \dfrac{z\,\Delta \overline{\theta }}{\left[ \overline{u}^2 + \overline{v}^2\right] } = \dfrac{z}{\mathcal {L}}\, \dfrac{C_\mathrm{m}^2}{\kappa \,\sigma _{\textsc {n}}\,C_\theta }, \end{aligned}$$
(31)

where the horizontal wind speed and temperature difference relative to the wall (\(\Delta \overline{\theta }\)) at height z are known, given by the mesoscale solution at the first vertical level above the wall. The value of \(\mathcal {L}\) is computed by numerically solving Eq. 31 using a bisection root-finding method. Equations 29 and 30 are then used to obtain \(u_*\) and \(\theta _*\).

1.2 Prescribed Heat-Flux Condition at the Bottom Surface

The mesoscale model provides fields for both surface temperature and sensible heat flux. Although a prescribed temperature is straightforward to implement, e.g. using (31), it may introduce severe discrepancies due to the sensitivity of the temperature with height. As the microscale and mesoscale grids have different elevations for the same location, the temperature variation between both heights may be large. Instead, as the heat flux within the surface layer is expected to be nearly constant with height, it is preferable to use it as a boundary condition for the microscale.

This boundary condition may yield two solutions of \(\mathcal {L}\) for the same surface heat flux, \(q_{\mathrm {w}}\), known as the duality problem (van de Wiel et al. 2007). As one of these solutions leads to a collapse of turbulence, such that both turbulent and laminar states are possible, a limitation was imposed on the maximum value of \(\zeta \) to force a continuous turbulent regime,

$$\begin{aligned} \max \left( \zeta \right) = \dfrac{z\,\ln \left( \left. { z}/{z_{\mathrm{m}0}}\right. \right) }{2\,b_\mathrm{ms}\left( z-z_{\mathrm{m}0}\right) }. \end{aligned}$$
(32)

This limit arises from the minimum value possible for \(q_{\mathrm {w}}\) to attain,

$$\begin{aligned} \min \left( q_{\mathrm {w}}\right) = \dfrac{-4\,\kappa ^2\,\rho _{\mathrm {w}}\,c_\mathrm{p}\,\vert \mathbf {u}_\parallel \vert ^3}{ 27\,\dfrac{g}{\varTheta }\,b_\mathrm{ms}\left( z-z_{\mathrm{m}0}\right) \, \ln \left( z/z_{\mathrm{m}0}\right) ^2}, \end{aligned}$$
(33)

where \(\vert \mathbf {u}_\parallel \vert \) is the magnitude of tangential velocity at the centre of the control volumes adjacent to the wall.

The stability parameter \(\zeta \) is computed from the known values of \(\vert \mathbf {u}_\parallel \vert \) and \(q_{\mathrm {w}}\). The velocity is related to \(u_*\) through the momentum bulk transfer coefficient (29),

$$\begin{aligned} u_* = \vert \mathbf {u}_\parallel \vert \, C_\mathrm{m}. \end{aligned}$$
(34)

A value for \(\zeta \) is computed by numerically solving the implicit relation obtained by combining (22) and (29). This was done at each surface grid point for each iteration of the segregated solver, using a fixed-point iteration method which was verified to converge in the range of \(\zeta \) for which the stability functions are valid, considering the limit imposed by Eq. 32.

Appendix 2: Error Measures

Considering the time series of a quantity X and \(X_n\) as its sample at instant n, for \(n=1,\,\ldots ,\,N\) records, the absolute error, \(\Delta X\), is defined as,

$$\begin{aligned} \Delta X_n = X^\textsc {f}_n - X^\textsc {m}_n, \end{aligned}$$
(35)

where the superscripts \(\textsc {f}\) and \(\textsc {m}\) refer to the forecasted and measured values. A positive value indicates that the forecast over-predicts the observation and otherwise for a negative value. If X represents the wind direction, \(\phi \), as this is a circular quantity in the interval \([0^{\circ },\,360^{\circ } [\), the azimuth difference between \(\phi ^\textsc {f}\) and \(\phi ^\textsc {m}\) was used to define the error, such that

$$\begin{aligned} \Delta X_n= \Delta \phi ^*_n -360\,{\text {sgn}}\big (\Delta \phi ^*_n\big )\, \max \Big (0,\,{\text {sgn}}\big (\vert \Delta \phi ^*_n\vert -180\big )\Big ), \end{aligned}$$
(36)

where \(\Delta \phi ^*_n = \phi ^\textsc {f}_n - \phi ^\textsc {m}_n\). The azimuthal error will only take values in \([-180^{\circ },\,180^{\circ } ]\).

From \(\Delta X\), the bias (B), mean-squared-error (\(\textit{MSE}\)) and its square-root (\(\textit{RMSE}\)) are defined as:

$$\begin{aligned} B(X^\textsc {f})= & {} \dfrac{1}{N}\sum ^{N}_{n=1}\Delta X_n, \end{aligned}$$
(37)
$$\begin{aligned} \textit{MSE}(X^\textsc {f})= & {} \dfrac{1}{N}\sum ^{N}_{n=1}\Delta X_n^2, \end{aligned}$$
(38)
$$\begin{aligned} \textit{RMSE}(X^\textsc {f})= & {} \sqrt{\textit{MSE}}. \end{aligned}$$
(39)

The linear correlation between \(X^\textsc {f}\) and \(X^\textsc {m}\) was estimated using the Pearson correlation coefficient,

$$\begin{aligned} r=\left[ \sum _{n=1}^{N} \left( X^\textsc {f}_n - \overline{X^\textsc {f}}\right) \, \left( X^\textsc {m}_n - \overline{X^\textsc {m}}\right) \right] \Bigg /\sqrt{ \sum _{n=1}^{N}\left( X^\textsc {f}_n-\overline{X^\textsc {f}}\right) ^2 \,\sum _{n=1}^{N}\left( X^\textsc {m}_n-\overline{X^\textsc {m}}\right) ^2},\nonumber \\ \end{aligned}$$
(40)

where the overbar denotes the sample mean.

If two forecasts for X exist, \(X^\textsc {f}\) and \(X^\textsc {r}\), the skill score (SS) may be used to estimate the improvement of one forecast over another (Murphy 1988). Choosing one of the forecasts as reference, \(X^\textsc {r}\), the SS is obtained by comparing the respective mean squared errors:

$$\begin{aligned} \textit{SS} = 1 - \textit{MSE}(X^\textsc {f})/\textit{MSE}(X^\textsc {r}), \end{aligned}$$
(41)

where \(\textit{MSE}(X^\textsc {f})\) and \(\textit{MSE}(X^\textsc {r})\) refer to the mean squared error of \(X^\textsc {f}\) and \(X^\textsc {r}\). The SS is dimensionless and varies between \(-\infty \) and 1, indicating that forecast \(X^F\) has lower error than \(X^R\) when \(SS>0\).

Appendix 3: Statistics for Unsteady RANS Results

The flow fields of an unsteady RANS model refer to mean and turbulent quantities specific to a timestep \(\Delta t\). The velocity field is regarded as either its ensemble or time average, \(\overline{\mathbf {u}}\), while \(k=\tfrac{1}{2}\sum _i\overline{u'_i\,u'_i}\). The overbar is here used to represent the mean field which is related to the simulation timestep \(\Delta t\), whereas the operator \(\langle \,\rangle \) represents an average over a larger integration time \(\Delta T\), where \(\Delta T \gg \Delta t\).

To compare unsteady RANS results with cup anemometer measurements, the former have to be resampled as as the latter refer to integrations over \(\Delta T = 10\,\mathrm {min}\). Albeit the simulation mean fields vary in time, it is assumed that the mean and turbulent quantities are quasi-stationary, i.e. the time scale of turbulent motions is much smaller than that of the mean flow and at the timestep scale \(\Delta t\), turbulent fluctuations become null when averaged. Breaking \(\Delta T\) into \(N\,\Delta t\) intervals, the average of quantity u over \(\Delta T\) becomes,

$$\begin{aligned} \langle u\rangle =\dfrac{1}{\Delta T}\int \limits _0^{\Delta T}u\,\text {d}t =\dfrac{1}{N\,\Delta t}\sum \limits _{n=1}^{N} \left[ \overset{n\,\Delta t}{\underset{(n-1)\,\Delta t}{\int u\,\text {d}t}} \right] =\dfrac{1}{N}\sum \limits _{n=1}^{N} \overline{u}|_n \equiv \langle \overline{u} \rangle , \end{aligned}$$
(42)

where \(\overline{u}|_n\) represents the mean field of u at timestep \(n\Delta t\). Assuming that turbulent fluctuations are uncorrelated with the motions of the mean flow, \(\langle \overline{u}\,u'\rangle \approx 0\), the variance of u over \(\Delta T\) is

$$\begin{aligned} \sigma ^2_\mathrm{u} = \langle u^2\rangle - \langle u\rangle ^2 =\dfrac{1}{N}\sum \limits _{n=1}^{N} \left[ \overline{u}\big |_n^2 +2\,\overline{\overline{u}|\,u'}\Big |_n +\overline{u'u'}\Big |_n \right] -\langle \overline{u} \rangle ^2 =\langle \overline{u}{^2} \rangle +\langle \overline{u'u'}\rangle -\langle \overline{u} \rangle ^2.\nonumber \\ \end{aligned}$$
(43)

If u is a velocity component, the variance \(\overline{u'u'}\) at timestep \(n\Delta t\) is obtained applying the eddy viscosity hypothesis. From the definition of the turbulence stress tensor in Eq. 4,

$$\begin{aligned} \overline{u'_i u'_i} = -\dfrac{\tau '_{ii}}{\rho _{\textsc {h}}} = -2\,\nu _\mathrm{t}\,\dfrac{\partial \,\overline{u}_i}{\partial \,x_i} +\dfrac{2}{3}\,k. \end{aligned}$$
(44)

Cup anemometers measure the horizontal magnitude of the wind velocity and its variance, integrated over \(\Delta T\), while the simulations return the average components of the velocity vector at time scale \(\Delta t\). Averaging \(\overline{u}\) and \(\overline{v}\) over \(\Delta T\) will not yield the same value as computing the average of the magnitude itself. Applying the average operator on the horizontal wind speed, s, then

$$\begin{aligned} s=\sqrt{u^2+v^2} \;\Rightarrow \; \left\{ \begin{array}{l} \langle {s}\rangle =\big \langle \sqrt{u^2+v^2}\,\big \rangle , \\ \langle {s^2}\rangle =\langle {u^2}\rangle +\langle {v^2}\rangle . \end{array}\right. \end{aligned}$$
(45)

Substituting in (45) the mean of the squares by the variance, i.e. \(\langle {\chi ^2}\rangle = \sigma ^2_\chi +\langle {\chi }\rangle ^2\),

$$\begin{aligned} \sigma ^2_\mathrm{s} + \langle {s}\rangle ^2 =\sigma ^2_\mathrm{u} + \langle {u}\rangle ^2 +\sigma ^2_\mathrm{v} + \langle {v}\rangle ^2. \end{aligned}$$
(46)

This identity is equally valid for a smaller timestep, \(\Delta t\), relating both \(\overline{s}\) and \(\overline{s's'}\) to \(\overline{u}\), \(\overline{u'u'}\), \(\overline{v}\) and \(\overline{v'v'}\). As the unsteady RANS model returns no information on \(\overline{s}\) (only its components \(\overline{u}\) and \(\overline{v}\)) the best assumption becomes

$$\begin{aligned} \overline{s}^2 \approx \overline{u}^2 + \overline{v}^2 \;\Rightarrow \; \overline{s's'} \approx \overline{u'u'} + \overline{v'v'}. \end{aligned}$$
(47)

Thus, for an integration time \(\Delta T\), the mean and variance of the horizontal wind speed are estimated as

$$\begin{aligned} \langle {s}\rangle= & {} \langle \overline{s}\rangle =\big \langle \sqrt{\overline{u}{^2} + \overline{v}{^2}}\,\big \rangle , \end{aligned}$$
(48)
$$\begin{aligned} \sigma ^2_\mathrm{s}= & {} \langle \overline{u'u'}\rangle + \langle \overline{u}^2\rangle + \langle \overline{v'v'}\rangle + \langle \overline{v}^2\rangle - \langle \overline{s}\rangle ^2. \end{aligned}$$
(49)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rodrigues, C.V., Palma, J.M.L.M. & Rodrigues, Á.H. Atmospheric Flow over a Mountainous Region by a One-Way Coupled Approach Based on Reynolds-Averaged Turbulence Modelling. Boundary-Layer Meteorol 159, 407–437 (2016). https://doi.org/10.1007/s10546-015-0116-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10546-015-0116-7

Keywords

Navigation