Skip to main content
Log in

Space–time calibration of wind speed forecasts from regional climate models

  • Published:
Environmental and Ecological Statistics Aims and scope Submit manuscript

Abstract

Numerical weather predictions (NWPs) are systematically subject to errors due to the deterministic solutions used by numerical models to simulate the atmosphere. Statistical postprocessing techniques are widely used nowadays for NWP calibration. However, time-varying bias is usually not accommodated by such models. The calibration performance is also sensitive to the temporal window used for training. This paper proposes space–time models that extend the main statistical postprocessing approaches to calibrate NWP model outputs. Trans-Gaussian random fields are considered to account for meteorological variables with asymmetric behavior. Data augmentation is used to account for the censoring of the response variable. The benefits of the proposed extensions are illustrated through the calibration of hourly 10-m height wind speed forecasts in Southeastern Brazil coming from the Eta 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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Ailliot P, Monbet V, Prevosto M (2006) An autoregressive model with time-varying coefficients for wind fields. Environmetrics 17(2):107–117

    Article  Google Scholar 

  • Ailliot P, Thompson C, Thomson P (2009) Space-time modelling of precipitation by using a hidden Markov model and censored Gaussian distributions. J R Stat Soc Ser C 58(3):405–426

    Article  Google Scholar 

  • Akaike H (1974) A new look at the statistical model identification. IEEE Trans Autom Control 19(6):716–723

    Article  Google Scholar 

  • Allard D, Bourotte M (2015) Disaggregating daily precipitations into hourly values with a transformed censored latent Gaussian process. Stochast Environ Res Risk Assess 29(2):453–462

    Article  Google Scholar 

  • Allcroft DJ, Glasbey CA (2003) A latent Gaussian Markov random-field model for spatiotemporal rainfall disaggregation. J R Stat Soc Ser C 52(4):487–498

    Article  Google Scholar 

  • Amarante OD, Silva F, Andrade P (2010) Atlas Eólico: Minas Gerais. CEMIG, Belo Horizonte

    Google Scholar 

  • Anderson JL (1996) A method for producing and evaluating probabilistic forecasts from ensemble model integrations. J Clim 9(7):1518–1530

    Article  Google Scholar 

  • Bardossy A, Plate EJ (1992) Space-time model for daily rainfall using atmospheric circulation patterns. Water Res Res 28(5):1247–1259

    Article  Google Scholar 

  • Berrocal VJ, Raftery AE, Gneiting T (2007) Combining spatial statistical and ensemble information in probabilistic weather forecasts. Month Weather Rev 135(4):1386–1402

    Article  Google Scholar 

  • Black TL (1994) The new NMC mesoscale Eta model: description and forecast examples. Weather Forecast 9(2):265–278

    Article  Google Scholar 

  • Box GE, Cox DR (1964) An analysis of transformations. J R Stat Soc Ser B 211–252

  • Carter CK, Kohn R (1994) On Gibbs sampling for state space models. Biometrika 81(3):541–553

    Article  Google Scholar 

  • Chou S, Souza Cd, Gomes JL, Evangelista E, Osório C, Cataldi M (2007) Refinamento estatístico das previsões horárias de temperatura a 2 m do modelo Eta em estações do Nordeste do Brasil. Rev Bras Meteorol 22(3):287–296

    Article  Google Scholar 

  • Cressie N (1993) Statistics for spatial data. Wiley, New York

    Book  Google Scholar 

  • De Oliveira V, Kedem B, Short DA (1997) Bayesian prediction of transformed Gaussian random fields. J Am Stat Assoc 92(440):1422–1433

    Google Scholar 

  • Diggle P, Ribeiro P, Geostatistics M-b (2007) Springer series in statistics

  • Eddelbuettel D, François R, Allaire J, Ushey K, Kou Q, Russel N, Chambers J, Bates D (2011) Rcpp: seamless R and C++ integration. J Stat Softw 40(8):1–18

    Article  Google Scholar 

  • Epstein ES (1969) Stochastic dynamic prediction. Tellus 21(6):739–759

    Google Scholar 

  • Feldmann K, Scheuerer M, Thorarinsdottir TL (2015) Spatial postprocessing of ensemble forecasts for temperature using nonhomogeneous Gaussian regression. Month Weather Rev 143(3):955–971

    Article  Google Scholar 

  • Frühwirth-Schnatter S (1994) Data augmentation and dynamic linear models. J Time Series Anal 15(2):183–202

    Article  Google Scholar 

  • Fuentes M, Chen L, Davis JM (2008) A class of nonseparable and nonstationary spatial temporal covariance functions. Environmetrics 19:487–507

    Article  CAS  Google Scholar 

  • Gel Y, Raftery AE, Gneiting T (2004) Calibrated probabilistic mesoscale weather field forecasting: the geostatistical output perturbation method. J Am Stat Assoc 99(467):575–583

    Article  Google Scholar 

  • Gelfand AE (1996) Model determination using sampling-based methods. Markov chain Monte Carlo in practice 145–161

  • Gelman A, Rubin DB et al (1992) Inference from iterative simulation using multiple sequences. Stat Sci 7(4):457–472

    Article  Google Scholar 

  • Genton MG, Zhang H (2012) Identifiability problems in some non-Gaussian spatial random fields. Chilean J Stat 3(2):171–179

    Google Scholar 

  • Glahn HR, Lowry DA (1972) The use of model output statistics (MOS) in objective weather forecasting. J Appl Meteorol 11(8):1203–1211

    Article  Google Scholar 

  • Gneiting T (2014) Calibration of medium-range weather forecasts. European Centre for Medium-Range Weather Forecasts, Reading

  • Gneiting T, Raftery AE (2007) Strictly proper scoring rules, prediction, and estimation. J Am Stat Assoc 102(477):359–378

    Article  CAS  Google Scholar 

  • Gneiting T, Raftery AE, Westveld AH III, Goldman T (2005) Calibrated probabilistic forecasting using ensemble model output statistics and minimum CRPS estimation. Month Weather Rev 133(5):1098–1118

    Article  Google Scholar 

  • Gneiting T, Schlather M (2004) Stochastic models that separate fractal dimension and the Hurst effect. SIAM Rev 46(2):269–282

    Article  Google Scholar 

  • Grimit EP, Mass CF (2007) Measuring the ensemble spread-error relationship with a probabilistic approach: stochastic ensemble results. Month Weather Rev 135(1):203–221

    Article  Google Scholar 

  • Guerrero VM (1993) Time-series analysis supported by power transformations. J Forecast 12(1):37–48

    Article  Google Scholar 

  • Hamill TM (2001) Interpretation of rank histograms for verifying ensemble forecasts. Month Weather Rev 129(3):550–560

    Article  Google Scholar 

  • Hansen FV (1993) Surface roughness lengths. Technical Report ARL-TR-61, U.S. Army Research Laboratory, p 45

  • Kalman RE (1960) A new approach to linear filtering and prediction problems. J Basic Eng 82(1):35–45

    Article  Google Scholar 

  • Krishnamurti T (1995) Numerical weather prediction. Ann Rev Fluid Mech 27(1):195–225

    Article  Google Scholar 

  • Lange M (2005) On the uncertainty of wind power predictions-analysis of the forecast accuracy and statistical distribution of errors. J Solar Energy Eng 127(2):177–184

    Article  Google Scholar 

  • Li W, Duan Q, Miao C, Ye A, Gong W, Di Z (2017) A review on statistical postprocessing methods for hydrometeorological ensemble forecasting. WIREs Water 4(e1246):1–24

    Google Scholar 

  • Mao Q, McNider RT, Mueller SF, Juang H-MH (1999) An optimal model output calibration algorithm suitable for objective temperature forecasting. Weather Forecast 14(2):190–202

    Article  Google Scholar 

  • Matérn B (1986) Spatial variation, 2nd edn. Springer, New York

    Book  Google Scholar 

  • Matheson JE, Winkler RL (1976) Scoring rules for continuous probability distributions. Manage Sci 22(10):1087–1096

    Article  Google Scholar 

  • Mesinger F, Janjić ZI, Ničković S, Gavrilov D, Deaven DG (1988) The step-mountain coordinate: model description and performance for cases of Alpine lee cyclogenesis and for a case of an Appalachian redevelopment. Month Weather Rev 116(7):1493–1518

    Article  Google Scholar 

  • Monin AS, Obukhov AM (1954) Basic laws of turbulent mixing in the surface layer of the atmosphere. Contrib Geophys Inst Acad Sci USSR 151(163):e187

    Google Scholar 

  • Piani C, Haerter J, Coppola E (2010) Statistical bias correction for daily precipitation in regional climate models over Europe. Theor Appl Climatol 99(1–2):187–192

    Article  Google Scholar 

  • R Core Team (2017) R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. URL http://www.R-project.org

  • Raftery A, Lewis S (1995) The number of iterations, convergence diagnostics and generic Metropolis algorithms. Practical Markov Chain Monte Carlo, 115–130

  • Raftery AE, Gneiting T, Balabdaoui F, Polakowski M (2005) Using Bayesian model averaging to calibrate forecast ensembles. Month Weather Rev 133(5):1155–1174

    Article  Google Scholar 

  • Saha S, Moorthi S, Wu X, Wang J, Nadiga S, Tripp P, Behringer D, Hou Y-T, Chuang H-Y, Iredell M et al (2014) The NCEP climate forecast system version 2. J Clim 27(6):2185–2208

    Article  Google Scholar 

  • Sanderson C, Curtin R (2016) Armadillo: a template-based C++ library for linear algebra. J Open Source Softw 1(2):26

  • Sansó B, Guenni L (1999) Venezuelan rainfall data analysed by using a Bayesian space-time model. J R Stat Soc Ser C 48(3):345–362

    Article  Google Scholar 

  • Sansó B, Guenni L (2004) A Bayesian approach to compare observed rainfall data to deterministic simulations. Environmetrics 15(6):597–612

    Article  Google Scholar 

  • Scheuerer M, Möller D et al (2015) Probabilistic wind speed forecasting on a grid based on ensemble model output statistics. Ann Appl Stat 9(3):1328–1349

    Article  Google Scholar 

  • Schuhen N, Thorarinsdottir TL, Gneiting T (2012) Ensemble model output statistics for wind vectors. Month Weather Rev 140(10):3204–3219

    Article  Google Scholar 

  • Sigrist F, Künsch HR, Stahel WA (2015) Stochastic partial differential equation based modelling of large space-time data sets. J R Stat Soc Ser B 77(1):3–33

    Article  Google Scholar 

  • Sigrist F, Künsch HR, Stahel WA et al (2012) A dynamic nonstationary spatio-temporal model for short term prediction of precipitation. Ann Appl Stat 6(4):1452–1477

    Article  Google Scholar 

  • Spiegelhalter DJ, Best NG, Carlin BP, Van Der Linde A (2002) Bayesian measures of model complexity and fit. J R Stat Soc Ser B 64(4):583–639

    Article  Google Scholar 

  • Tanner MA, Wong WH (1987) The calculation of posterior distributions by data augmentation. J Am stat Assoc 82(398):528–540

    Article  Google Scholar 

  • Thorarinsdottir TL, Gneiting T (2010) Probabilistic forecasts of wind speed: ensemble model output statistics by using heteroscedastic censored regression. J R Stat Soc Ser A 173(2):371–388

    Article  Google Scholar 

  • Thorarinsdottir TL, Johnson MS (2012) Probabilistic wind gust forecasting using nonhomogeneous Gaussian regression. Month Weather Rev 140(3):889–897

    Article  Google Scholar 

  • Vannitsem S, Wilks D, Messner J (2018) Statistical postprocessing of ensemble forecasts. Elsevier, Amsterdam

    Google Scholar 

  • Vihola M (2012) Robust adaptive metropolis algorithm with coerced acceptance rate. Stat Comput 22(5):997–1008

    Article  Google Scholar 

  • West M, Harrison P (1997) Bayesian forecasting and dynamic models, 2nd edn. Springer Verlag, New York

    Google Scholar 

  • Whitaker JS, Loughe AF (1998) The relationship between ensemble spread and ensemble mean skill. Month Weather Rev 126(12):3292–3302

    Article  Google Scholar 

  • Willmott CJ (1981) On the validation of models. Phys Geogr 2(2):184–194

    Article  Google Scholar 

  • Wilson LJ, Vallée M (2002) The Canadian updateable model output statistics (UMOS) system: design and development tests. Weather Forecast 17(2):206–222

    Article  Google Scholar 

  • Zhang H (2004) Inconsistent estimation and asymptotically equal interpolations in model-based geostatistics. J Am Stat Assoc 99(465):250–261

    Article  Google Scholar 

  • Zhang H, El-Shaarawi A (2010) On spatial skew-Gaussian processes and applications. Environmetrics 21(1):33–47

    CAS  Google Scholar 

  • Zhu X, Genton MG (2012) Short-term wind speed forecasting for power system operations. Int Stat Rev 80(1):2–23

    Article  Google Scholar 

Download references

Acknowledgements

We thank Dr. Chou Sin Chan (CPTEC/INPE) for providing the Eta model outputs used in this study and for her advice during the research. The first author also thanks the financial support from the partnership between Fundação de Amparo à Pesquisa do Estado de Minas Gerais (FAPEMIG) and Companhia Energética de Minas Gerais S.A. (CEMIG) under Project APQ-03813-12.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thaís C. O. Fonseca.

Additional information

Handling Editor: Bryan F. J. Manly.

Appendices

Appendix A: Calculation of the roughness length

According to Hansen (1993), the vegetation present on a surface influences the aerodynamic roughness characteristics encountered by the mean wind flow over that surface, affecting both the mean wind speed and direction predicted by numerical models and various other atmospheric parameters. Therefore, the surface roughness length, \(z_0\), defined as the height at which the wind speed equals zero, has an important role in the modeling of atmospheric processes.

The aerodynamic roughness length parameter, \(z_0\), used as a covariate in the application, was estimated for each calibration site from key atmospheric variables, following some principles of the Monin-Obukhov similarity theory (Monin and Obukhov 1954). Particularly, assuming a logarithmic wind profile, the averaged wind speed \(\overline{u_i}\) at height \(z_i\) (10 m), given by:

$$\begin{aligned} \overline{u_i}= \frac{u_{*}}{k} \left[ \ln \left( \frac{z_i}{z_{0}}\right) -\Psi (\zeta _i) \right] \end{aligned}$$

was used to derive the roughness length \(z_0\) as

$$\begin{aligned} z_{0}= z_{i}\exp \left( \frac{-\overline{u_{i}} k}{{u}_{*}} - \Psi (\zeta _{i})\right) , \end{aligned}$$
(12)

where k is the von Karman constant, \(u_{*}\) is the friction velocity, \(\Psi (\zeta _{i})\) is the stability correction function of the wind profile and \(\zeta _i=\frac{z_i}{L}\) is the dimensionless stability parameter given by the height above ground, \(z_i\), normalized by the Obukhov length, L.

Measurements of air temperature, air pressure, sensible heat flux, momentum flux, and wind stress can be used to derive the parameters \(u_{*}\), L and \(\Psi (\zeta _{i})\). In this application hourly air temperature and air pressure data were obtained from the available meteorological stations, while hourly reanalysis data from the CFSV2 model (Saha et al. 2014) were used as the source of heat and momentum fluxes, after interpolation from the CFSV2 regular grid to the calibration sites.

Hourly values of the roughness length, \(z_0\), were first obtained from (12) for the 2 years of available data. Then, median values of \(z_0\) by month and by hour within each month (288 values in total for each calibration site) were calculated, considering only those \(z_0\) values that were estimated during neutral conditions of atmospheric stability (such condition is achieved when \(|L|>500\)).

Appendix B: Robustness of prior distributions

In this section, we verify the robustness of prior distributions assigned to the parameters of the proposed postprocessing models applied in Sect. 4. Based on these prior distributions in simulated experiments, the posterior and prior distributions associated with the parameter vectors from DGOP and STEMOS are presented in Fig. 10. For clarity of exposition, only the results of a subset of common parameter \(\varvec{\theta }_0\) (vector with dimension r) are exhibited. Specifically, the intercept parameter at moment \(t=0\) represented by \(\theta _{0,0}\) is shown.

Even when using this specific set of prior distributions in a distinct application from that previously reported, Fig. 10 shows that the posterior distributions differ significantly from the prior distributions for all parameters. This is evidence that the gain of information is mostly provided by the data, ensuring that non-informative properties are preserved in general applications.

Fig. 10
figure 10

Posterior and prior distributions associated with parameters of a DGOP and b STEMOS in simulated experiments. The vertical line represents the actual value assigned for parameters

Appendix C: Sensitivity analysis of discount factors

This sensitivity analysis was performed to choose the best combination of discount factors (\(\tilde{\delta }_T\), \(\tilde{\delta }_S\), \(\tilde{\delta }_V)'\) for (\(\delta _{T}\), \(\delta _{S}\), \(\delta _{V})'\). In particular, we also consider a subset of 59 weather stations and a random selection of 5 days per season from the temporal range of the available dataset. Following the same model comparison criteria described in Sect. 4, Table 4 reports the sensitivity analysis of discount factors through average MAE, RMSE, d, IS, DIC, and LPML values for 24-h wind speed forecasts at 10-m height. The setups of the dynamic models differ only in the discount factor values. The results indicate greater sensitivity for \(\delta _T\), which significantly reduces the predictive performance when assuming lower values. In contrast, the variation of values for \(\delta _{S}\) and \(\delta _{V}\) only slightly alters the predictive performance. Thus, we determine (\(\tilde{\delta }_T\), \(\tilde{\delta }_S\), \(\tilde{\delta }_V)' = (0.99, 0.95, 0.99)'\).

Table 4 Average MAE, RMSE, d, IS, DIC, and LPML of 24-h deterministic forecasts for wind speed at 10-m height (sensitivity analysis of discount factors)

Appendix D: Model comparison criteria

In this section, we briefly describe the model comparison criteria used to compare the prediction of the fitted models in Sect. 4. The first three criteria (RMSE, MAE and index of agreement) are appropriate to compare numerical predictions from the Eta model, which provides only deterministic estimates, with the proposed postprocessing models. The probabilistic forecasts are evaluated through IS, which takes into account the amplitude and coverage of the prediction intervals in a parsimonious way.

1.1 D.1  Mean absolute error and root-mean-square error

Standard measures of goodness of fit were also entertained in this study for comparison purposes. The root-mean-square error (RMSE) and the mean absolute error (MAE) are given by:

$$\begin{aligned} \text{ RMSE }=\frac{1}{nT}\sum _{i=1}^{n}\sum _{t=1}^{T}(y_{t}(s_i)-\hat{y}_{t}(s_i))^2 \ \ \ \text{ and } \ \ \ \text{ MAE }=\frac{1}{nT}\sum _{i=1}^{n}\sum _{t=1}^{T}|y_{t}(s_i)-\hat{y}_{t}(s_i)|, \end{aligned}$$

where \(\hat{y}_{t}(s_i)\) is obtained through a Monte Carlo estimate of the posterior mean of the predictive distribution, that is, \(E\left[ y_{t}(\mathbf{s}_i) \mid \mathbf{y}\right] \), across N draws. Smaller values of RMSE and MAE indicate better model fit.

1.2 D.2  Index of agreement

Willmott (1981) introduced a standard measure for assessing the quality of forecasts. The index of agreement (d) ranges between 0 (absence of agreement) and 1 (perfect agreement), and is given by:

$$\begin{aligned} d = 1 - \frac{\sum _{i=1}^n \sum _{t=1}^{T}\left( y_{t}(s_i) - \hat{y}_t(s_i)\right) ^2}{\sum _{i=1}^n \sum _{t=1}^{T}\left( |\hat{y}_t(s_i) - \bar{y}|+|y_t(s_i) - \bar{y}|\right) ^2}, \end{aligned}$$

where \(\bar{y} = \frac{1}{n} \sum _{i=1}^n\sum _{t=1}^{T} y_t(s_i)\).

1.3 D.3  Interval score

The interval score (IS, Gneiting and Raftery 2007) is a scoring rule for interval predictions considering the symmetric prediction interval with level \((1-\alpha )\times \)100%. The score is rewarded by accurate intervals and penalized when there is no coverage of the forecast. If actual values are contained in the prediction interval, this measure is reduced to the range amplitude. The average IS is given by:

$$\begin{aligned} \begin{aligned} \text {IS} =&\quad \frac{1}{nT}\sum _{i=1}^n\sum _{i=1}^T (\hat{u}_t(s_i)-\hat{l}_t(s_i)) \\&+\frac{2}{\alpha }(\hat{l}_t(s_i)-y_t(s_i))\mathbb {1}\left\{ y_t(s_i)<\hat{l}_t(s_i) \right\} \\&+\frac{2}{\alpha }(y_t(s_i)-\hat{u}_t(s_i))\mathbb {1}\left\{ y_t(s_i)>\hat{u}_t(s_i) \right\} \end{aligned} \end{aligned}$$

where \(\hat{l}_t(s_i)\) and \(\hat{u}_t(s_i)\) are, respectively, the lower bound obtained by the \(\frac{\alpha }{2}\) quantile, and the upper bound, obtained by the \(1-\frac{\alpha }{2}\) quantile based on the predictive distribution. The indicator function is represented by \(\mathbb {1}\).

Smaller IS values indicate more efficient probabilistic forecasts.

1.4 D.4  DIC

Particularly useful in Bayesian model selection problems, the deviance information criterion (DIC, Spiegelhalter et al. 2002) is a hierarchical modeling generalization of the Akaike information criterion (AIC, Akaike 1974). This criterion is negatively oriented implying that models with smaller DIC should be preferred to models with larger DIC. The DIC is given by:

$$\begin{aligned} \text {DIC} = -2 \int \log \{p(\mathbf{y} |\Theta )\}p(\Theta |\mathbf{y} )d\Theta + \text {P}_\text {D}, \end{aligned}$$

where \(\mathbf{y} \) is a vector of observed values and \(\Theta \) is the parameter vector. Thus, \(p(\mathbf{y} |\Theta )\) represents the likelihood function and \(p(\Theta | \mathbf{y} )\), the posterior distribution. Defined as a Bayesian measure of model complexity, \(\text {P}_\text {D}\) is given by:

$$\begin{aligned} \text {P}_\text {D} = 2 \log \{p(\mathbf{y} |\tilde{\Theta })\} -2 \int \log \{p(\mathbf{y} |\Theta )\}p(\Theta |\mathbf{y} )d\Theta , \end{aligned}$$

with \(\tilde{\Theta }\) denoting the Bayes estimator.

A Monte Carlo approximation of DIC is given by:

$$\begin{aligned} \widehat{\text {DIC}} = -2 \frac{1}{M}\sum _{m = 1}^{M} \log \{p(\mathbf{y} |\Theta ^{(m)})\} + \hat{\text {P}}_\text {D}, \end{aligned}$$

where

$$\begin{aligned} \hat{\text {P}}_\text {D} = 2 \log \{p(\mathbf{y} |\tilde{\Theta })\} -2 \frac{1}{M}\sum _{m = 1}^{M}\log \{p(\mathbf{y} |\Theta ^{(m)})\}, \end{aligned}$$

with \({\Theta ^{(m)}}\) denoting the m-th MCMC sample of \(\Theta \) from posterior distribution \(p(\Theta |\mathbf{y} )\), \(m = 1, \dots , M\).

1.5 D.5  LPML

A component of the Bayes factor, the logarithm of the pseudo marginal likelihood (LPML, Gelfand 1996) is given by:

$$\begin{aligned} \text {LPML} = \frac{1}{n} \sum _{i=1}^{n}\log \{\text {CPO}_i\}, \end{aligned}$$

where \(\text {CPO}_i\) represents the conditional predictive ordinate (CPO) for location i and is given by:

$$\begin{aligned} \begin{aligned} \text {CPO}_i =&\quad p\big (\mathbf{y} (s_i)| \mathbf{y} (s_{-i})\big ) \\ =&\quad \bigg (\int \frac{1}{p(\mathbf{y} (s_i)|\Theta )}p(\Theta |\mathbf{y} )d\Theta \bigg )^{-1}, \end{aligned} \end{aligned}$$

with \(\mathbf{y} (s_{i})=\big ({y}_{1}(s_{i}), \ldots , {y}_{T}(s_{i})\big )'\) and \(\mathbf{y} (s_{-i})=\big (\mathbf{y} (s_{1}),\ldots ,\mathbf{y} (s_{i-1}), \mathbf{y} (s_{i+1}),\ldots , \mathbf{y} (s_{n})\big )'\), \(i = 1,\ldots ,n.\) Note that the CPO\(_i\) is based on the leave-one-out-cross-validation process and estimates the probability of \(\mathbf{y} (s_{i})\) given the observation of \(\mathbf{y} (s_{-i})\). In particular, a Monte Carlo approximation of CPO\(_i\) is given by:

$$\begin{aligned} \widehat{\text {CPO}}_i = \bigg (\frac{1}{M}\sum _{m = 1}^{M} \frac{1}{p(\mathbf{y} (s_i)|\Theta ^{(m)})}\bigg )^{-1}, \end{aligned}$$

with \({\Theta ^{(m)}}\) denoting the m-th MCMC sample of \(\Theta \) from posterior distribution \(p(\Theta |\mathbf{y} )\), \(m = 1, \ldots , M\).

Finally, the Monte Carlo approximation of LPML is given by:

$$\begin{aligned} \widehat{\text {LPML}} = \frac{1}{n} \sum _{i=1}^{n}\log \{\widehat{\text {CPO}}_i\}. \end{aligned}$$

The preferred model maximizes this criterion.

Appendix E: Supplementary results

See Figs. 11, 12, 13, 14.

Fig. 11
figure 11

Boxplots of average MAE of 24-h deterministic forecasts for wind speed at 10-m height during the seasons. The dense horizontal line and the \(\mathbf {\times }\) symbol in each boxplot represent respectively, the median and mean values

Fig. 12
figure 12

Boxplots of average RMSE of 24-h deterministic forecasts for wind speed at 10-m height during the seasons. The dense horizontal line and the \(\mathbf {\times }\) symbol in each boxplot represent respectively, the median and mean values

Fig. 13
figure 13

Boxplots of average index of agreement of 24-h deterministic forecasts for wind speed at 10-m height during the seasons. Absolute agreement between predictions and actual values occurs when \(d=1\). The dense horizontal line and the \(\mathbf {\times }\) symbol in each boxplot represent respectively, the median and mean values

Fig. 14
figure 14

Normal quantile-quantile plots for residuals of each fitted model during the seasons

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gomes, L.E.S., Fonseca, T.C.O., Gonçalves, K.C.M. et al. Space–time calibration of wind speed forecasts from regional climate models. Environ Ecol Stat 28, 631–665 (2021). https://doi.org/10.1007/s10651-021-00509-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10651-021-00509-0

Keywords

Navigation