Next Article in Journal
Design of 3D Carbon Nanotube Monoliths for Potential-Controlled Adsorption
Next Article in Special Issue
Research Needs for Realization of Zero-Carbon Power Grids with Selected Case Studies
Previous Article in Journal
3D Numerical Simulation and Performance Analysis of CO2 Vortex Tubes
Previous Article in Special Issue
Economic Assessment and Control Strategy of Combined Heat and Power Employed in Centralized Domestic Hot Water Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Combined Forecasting System Based on Modified Multi-Objective Optimization for Short-Term Wind Speed and Wind Power Forecasting

School of Information Science and Engineering, Lanzhou University, Lanzhou 730000, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(20), 9383; https://doi.org/10.3390/app11209383
Submission received: 20 July 2021 / Revised: 18 September 2021 / Accepted: 23 September 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Grid Innovation in the Era of Smart Grids)

Abstract

:
Wind speed and wind power are two important indexes for wind farms. Accurate wind speed and power forecasting can help to improve wind farm management and increase the contribution of wind power to the grid. However, nonlinear and non-stationary wind speed and wind power can influence the forecasting performance of different models. To improve forecasting accuracy and overcome the influence of the original time series on the model, a forecasting system that can effectively forecast wind speed and wind power based on a data pre-processing strategy, a modified multi-objective optimization algorithm, a multiple single forecasting model, and a combined model is developed in this study. A data pre-processing strategy was implemented to determine the wind speed and wind power time series trends and to reduce interference from noise. Multiple artificial neural network forecasting models were used to forecast wind speed and wind power and construct a combined model. To obtain accurate and stable forecasting results, the multi-objective optimization algorithm was employed to optimize the weight of the combined model. As a case study, the developed forecasting system was used to forecast the wind speed and wind power over 10 min from four different sites. The point forecasting and interval forecasting results revealed that the developed forecasting system exceeds all other models with respect to forecasting precision and stability. Thus, the developed system is extremely useful for enhancing forecasting precision and is a reasonable and valid tool for use in intelligent grid programming.

1. Introduction

With the rise of globalization, renewable and alternative energy sources, which address security issues associated with conventional energy sources, are increasingly being favored to provide power for a wide range of social and economic activities. Wind energy, a promising technology utilized in many renewable energy systems, has attracted an increasing amount of attention in recent decades due to the drive to meet the rapidly growing electricity demand across the globe without emitting environmental pollutants, such as CO2. According to the GLOBAL WIND REPORT 2021 [1], China possesses 39% (278324MW) of the world’s total onshore wind power capacity, and accounted for 56% of new onshore installations by the end of 2020, making it the world leader. As a result, the accurate forecasting of wind speed—the determining factor in wind energy electricity generation—has increasingly become a focus of public conversation, especially on the short-term horizon. Unfortunately, it is difficult to obtain excellent forecasting results because of the stochastic and nonlinear oscillations caused by uneven atmosphere stratification and complex topography [2].
In recent years, the capacity of onshore wind power sets has been rapidly increasing, which requires high reliability and maintainability in wind turbines with poor natural conditions. Various methods have been developed to analyze wind turbine rotor blades. To identify the best-performing design for blade geometry, aerodynamic analyses are used to specify the desired radial distributions of the angle of attack, α (or sectional lift coefficient, cl) and axial induction factor along the blade and iterate the blade geometry (determined by the radial distributions of the blade chord, c, and twist, θ) until the required specifications are met. In aerodynamic evaluated processing, Selig and Tangler [3] use the blade element momentum (BEM) theory as for a method for aerodynamic analysis. Lee [4] used the vortex line method (VLM) and Moghadassian and Sharma [5] resolved Reynolds-averaged Navier–Stokes (RANS) equations using an actuator disk model (ADM) to emulate the rotor as a body force. These extensions permit analysis of unconventional rotor- and blade geometries, e.g., multi-rotor turbine configurations and rotor blades with dihedral and/or sweep. Meanwhile, the blade design of the wind turbine also affects the turbine output power coefficient (CP). For example, Tahani et al. [6] restricted the chord and twist distributions to linear profiles to increase the manufacturability of their design, and designed the rotor blade geometry (chord and twist) of a 1 MW wind turbine to maximize the CP. Liu et al. [7] investigated the blade design of a fixed-pitch, fixed-speed (FPFS) horizontal-axis wind turbine to maximize its annual energy production (AEP) for a prescribed wind speed Weibull distribution. They held the blade geometry fixed at the blade tip and assumed linear variations to reduce the design variables of the chord and twist values at the blade root.
Motivated by the need to urgently develop more precise measurement methods, a variety of studies have been conducted on wind speed forecasting over the past few decades. These studies can be broadly divided into two types: causal analyses and time series-based analyses [8]. Causal analyses use the established historical causality between explanatory and interpreted variables to forecast changes in future dependent variables [9]; however, such methods are prone to multiple collinear negative effects. Time series-based analyses can be subdivided into four approaches: physical [10], statistical [11], artificial intelligence [12], and hybrid [13] or combined models [14].
Advantages and shortcomings of those models include the following:
(1)
For physical methods, to attain effective forecasting results, information on a range of physical factors is needed, rather than information on just a single factor such as the wind speed time series; thus, these meteorological models cannot generate forecasts simply [15]. In addition, physical models are not proficient in dealing with short-term series and involve a complex calculation process and high expansion costs, which all contribute to significant forecasting errors [16]. Physical models necessitate the gathering of numerical physical variables, such as the horizontal pressure gradient, geostrophic force, and fractionate to undergo wind speed forecasting [17]. However, the complex calculation and polytrophic processes involved are not only time consuming but create the risk of forecasting error [18].
(2)
With sufficient accessible spatial and temporal information from multiple wind farms, a few spatio-temporal prediction methodologies have been able to be investigated in recent studies. The spatio-temporal characteristics of wind speed are extracted by an undirected graph of wind farms [19]. A hybrid support vector machine forecasting model is proposed, which is based on the spatio-temporal and grey wolf optimization, to forecast wind power for multiple wind farms [20]. Ref. [21] uses copula theory and Bayesian theory to simulate spatio-temporal correlations between wind farms and deduce a conditional distribution of aggregated wind power. A probabilistic wind speed prediction approach was presented in Ref. [22] based on a spatio-temporal neural network (STNN) and variational Bayesian inference. With feeding of both the spatial and temporal information into the forecasting model, these methods have achieved a better forecasting performance. Nevertheless, most of these forecasting models often collected the wind speed information from different wind farms indiscriminately, and to some extent, the implicit spatial correlations cannot be fully exploited in the original wind speed data. Also, the thorny multi-dimensional computing problem caused by the large amount of wind speed data from multiple wind farms needs to be solved in an effective way.
(3)
Typical statistical methods can yield excellent forecasting results under the assumption that the input series was recorded under normal conditions [23]; however, nonlinearity, noise, instability, fluctuations, and other features within the raw time series are always difficult to control, resulting in a lack of modeling information. Therefore, such methods often result in bad short-term wind speed forecasting performance, especially in multistep-ahead forecasting [24]. In addition, historical data are utilized for statistical modeling methods, and only potential linear correlations between the variables and future forecasts are revealed; such models are unable to obtain a good forecasting performance within the required limits [25]. Statistical models, including typical autoregressive moving average family models (e.g., AR [26], MA, ARMA [27], autoregressive integrated moving average models (ARIMA) [28], SARIMA, etc.), exponential smoothing [29], Kalman filtering [30], vector autoregression structures for very short-term wind power forecasting [31], and autoregressive conditional heteroskedastic family models (e.g., ARCH, GARCH, and EARCH) [32] utilize significant amounts of historical data for wind speed forecasting with no consideration of other potential influencing factors to support the stochastic process. Meanwhile, a spatial-temporal forecasting method based on the vector autoregression framework has been proposed for renewable forecasting [33]. Generally, statistical methods work well for approaching linear features; however, they tend to fail when it comes to nonlinear problems due to the linear assumptions of the models [34].
(4)
Fortunately, the timely emergence of artificial intelligence (AI) arithmetic, subsuming artificial neural networks (ANNs) [35], support vector machines (SVMs) [36], deep neural networks [37], and fuzzy logical methods (FLMs) [38] have efficiently remedied the flaws in the wind speed forecasting territory in recent years [39]. However, because of the inherent disadvantages of each model and the boom in the integration of wind power into the grid system, a variety of hybrid and combined models with promising forecasting potentials have been created [40]. Generally, artificial intelligence methods can achieve greater forecasting accuracy than physical or statistical models [41]; however, they also possess insurmountable drawbacks. ANNs have been extensively studied and applied to explore the complexity of wind speed forecasting; however, their performance mostly relies on training sets, which can result in a focus on local optima, over-fitting, and a reduction in the convergence rate [42].
(5)
Differently to conventional or single models, hybrid models can reduce the current shortcomings associated with the forecasting of irregular, fluctuant, and nonstationary time series with noise or unpredictable components. In this regard, significant hybrid models have recently been launched [43]. Hybrid methods integrate different single algorithms to achieve a superior forecasting performance, and can overcome defects in AI models (e.g., falling into local minima, over-fitting, etc.)—greatly improving the accuracy of continuous fluctuant wind speed forecasting and providing better validity and stability than a single model [44]. Recently, the use of hybrid methods in the wind speed forecasting field has been widespread. Jiang et al. [45] proposed a hybrid model consisting of a grey correlation analysis, cuckoo search algorithm, and v-SVM (v-support vector machine). Dong et al. [46] proposed a hybrid preprocessing strategy coupled with an optimized local linear fuzzy neural network for wind power forecasting. It has been proven that this is a effective approach for predicting wind power. In 2017, Hu et al. [47] proposed a novel approach based on the Gaussian process with a t-observation model for short-term wind speed forecasting. Based on a spatio-temporal method, in [48], the performance of predictive clustering trees with a new feature space for wind power forecasting was investigated. The results showed that the proposed model achieved a satisfactory level of point forecasting accuracy and interval forecasting performance. A forecasting framework has been proposed in 2021 [49], that is multi-layer stacked bidirectional long/short-term memory (LSTM)-based for short-term time series forecasting. After being studied extensively, it is clear that no arithmetic method is omnipotent across all data cases. In future research, individual statistical models and artificial intelligence algorithms will be integrated to improve the precision of wind speed forecasting—these are referred to as hybrid models [50].
As mentioned previously, conventional or individual methods always show inherent weaknesses when approaching complex and actual wind time series with noise, which results in poor forecasting performances.
In accordance with previous studies, we designed three procedures for wind speed and wind power forecasting. First, data preprocessing was employed to identify features and eliminate useless information from the original time series. Then, the preprocessed time series was used to train and test the multi-artificial neural network. The models were ranked based on the test set accuracy for each model. Finally, the top five models were used as sub-models of the combined model. To obtain stabilized and accurate forecasting results, the weight of the combined model was optimized by the multi-objective optimization algorithm, which improved the stabilization and accuracy of the combined model.
The key findings of this study and comparisons with relative research in the field of wind speed forecasting are as follows:
(1)
To achieve accurate and stable forecasting of short-term wind speed and wind power, a robust, novel, combined system based on three modules was developed in this study. This novel hybrid system for forecasting short-term wind speed and wind power includes a data preprocessing module, a forecast optimization module, and an evaluation module. The excellent performances of these algorithms are combined to provide accurate and stable results for multi-step wind speed forecasting.
(2)
To effectively eliminate fluctuations in the original time series and avoid the limitations of a single algorithm, a new data preprocessing step was proposed. This was shown to decrease uncertainty and irregularity in the wind speed times series. SSA-EEMD, a powerful secondary denoising algorithm, was used to decompose and further denoise the actual wind speed time series. These steps were found to successfully overcome the limitations of single SSA algorithms.
(3)
To overcome the disadvantages of individual models, multi models were used to forecast the wind speed and wind power. To consider the nonlinear characteristics of wind speed and wind power time series, seven artificial neural networks (ANNs) were employed to forecast two types of time series, and five optimal hybrid models were selected based on the accuracy of data testing by hybrid models to form a combined model and act as sub-models.
(4)
To further improve forecasting accuracy and stability, the multi-objective dragonfly algorithm was used to determine the optimal weight of the combined model. In the optimization process, the optimized parameters were found to not only have good accuracy, but they also ensured that the output results had a high level of stability. Therefore, this paper used the multi-objective optimization algorithm to optimize the weight of the combined model.
(5)
A more scientific and comprehensive forecasting evaluation method was conducted to estimate the forecasting performance of the developed forecasting system in the model evolution module. Interval forecasting was used to assess the uncertainty of the combined model, and this indicated that the forecasting results of the proposed combined model were accurate and stabilized in an all-around manner. Additionally, the Diebold–Mariano test and Wilcoxon rank-sum test were implemented to further analyze the forecasting accuracy of each model.
The remaining sections of this paper are as follows: The methods involved in the proposed model are introduced in Section 2. Section 3 demonstrates the experiment preparations and the four numerical experiments used, as well as presenting the forecast results. Deeper discussion about the forecasting performance of our developed model is presented in Section 4. Section 5 provides the conclusions.

2. Flow of the Proposed Combined Model

In the 1960s, J.M. Bates et.al. discovered combined forecasting methods, which use more than two different forecasting models to address the same problem. Multiple models can be combined by the combination of quantitative methods. The main purpose of this combination is to make full use of the information provided by various models to improve forecasting accuracy [51]. In our study, a combined forecasting model was developed to forecast the nonlinear characteristics of wind speed and wind power. The forecasting processes used are shown in Figure 1, and the details of the forecasting procedure are described below.
  • Procedure 1: Data Pretreatment
This stage included two integral parts. First, SSA was used to decompose the raw wind speed series into a (r) low-frequency components group [52] and a (d-r) high-frequency components group, according to KPAC. Second, EEMD [53] was utilized to further denoise the (d-r) high-frequency components group on the premise of SSA.
  • Procedure 2: Prediction of Hybrid Models
The forecasting accuracy of a single model cannot be optimal when a change in training data occurs during the forecasting process. Thus, single models do not always give optimal forecasting results when training data changes. To avoid a poor forecasting performance, combined models were developed. In this study, seven different models (ANFIS, BPNN, ELM, ENN, GRNN, LSTM, RNN, WNN and SVM) were employed to forecast wind speed and wind power, and five optimal forecasting models were selected as sub-models of the combined model. Meanwhile, data pre-processing was used to eliminate noise and useless information in the original time series. In the process of forecasting and modeling, the first 9 days (1296 data point) were used as the training set for the model, and days 10–14 (720 data points) were used as the testing set of the model. Each model produced 144 forecasting values.
  • Procedure 3: Establishment of the Proposed Combined System
After completing the modeling and forecasting process for the single hybrid model, the accuracy of each model testing set was calculated, and the five forecasting models with the highest accuracy levels were selected as sub-models of the combined model. Then, the testing data (720 data points) of the sub-models were used to build the combined model. The modified multi-objective dragonfly algorithm (the detail of MODA shown in Appendix A) was used to determine the optimal weight of the combined model. Subsequently, the forecasting results obtained from every individual model were integrated using the obtained weight coefficients, and wind speed and wind power forecasting was conducted.
  • Procedure 4: Wind Speed and Wind Power Forecasting
Based on realistic data, the developed combined model was used to carry out day-ahead forecasting. The theory behind day-ahead forecasting is as follows: First, the forecast origin and forecast horizon are set and denoted by a time point h and a positive whole number l, respectively. Then, if we want to forecast y ^ h + l at time point h where l ≥ 1, we now define y ^ h ( l ) as the prediction data of y h + l ; therefore, y ^ h ( l ) can represent the l-step day-ahead forecasting of yt at prediction point h. Once l = 1, y ^ h ( l ) is used to carry out one-step day-ahead forecasting [54].
  • Procedure 5: Model Evaluation
The forecasting precision and stability of the models during point forecasting were evaluated by four metrics. Three metrics were used for the uncertainty analysis.

3. Experiment and Results

In this section, the wind speed series data selection and experiment settings used for our study are illustrated systematically.

3.1. Data Acquisition

In this paper, the original wind speed and wind power time series used for forecasting were acquired in 2018 from four observation sites in Shandong province in China. These were used to establish and test the forecasting performance of each model.
Statistical descriptions for the datasets obtained for the two sites (mean, standard deviation, skewness, kurtosis, minimum, maximum, and median) are provided in Table 1. The mean values characterize the central tendency of the historical observations. The standard deviation values (the wind speed hovered around 2 m/s, and the wind power was around 200 kw) clearly reflect the fluctuations in wind speed and wind power. The skewness values of the wind speed and wind time series were greater than zero, which indicates that both types of time series had right-skewed distributions. Most wind time series had peak values greater than 3, indicating a fat tail distribution. Most kurtosis values of the wind power time series were greater than 3, indicating a fat tail distribution. The kurtosis values of the wind speed time series were less than 3, indicating a slight tail distribution.

3.2. Experimental Setup

In this study, three experiments were conducted to verify the validity, superiority, and generalizability of the proposed combinatorial optimization model. Experiment I was designed to compare the forecasting performances of combined models with different numbers of sub-models for wind speed and wind power forecasting in the first season. Experiment II contrasted the forecasting performances of the combined models. The models were optimized by different optimization algorithms for wind speed and wind power forecasting in the second season. Experiment III verified the performance of the combined model for wind speed and wind energy forecasting in the third and fourth seasons. The details of the experiments are as follows:
Experiment I was designed to compare the combined model with different numbers of forecasting sub-models to determine the optimal number of sub models required by the combined model. Wind speed and wind power data collected from four sites in the first season were used to determine the forecasting capacity of the proposed model. Furthermore, the forecasting step was classified as day-ahead and used to assess the conducted models.
Experiment II was conducted to compare the performances of different optimization algorithms to optimize the weight of the combined model. Wind speed and wind power series from Site 1 to Site 4 in the second season were collected at 10 min intervals to establish four combined models based on different optimization algorithms.
Experiment III aimed to verify the applicability of the combined model based on five sub-models and optimized by MODA. Wind speed and wind power time series data collected at 10 min intervals in the third and fourth seasons were employed to verify the prediction performance of the combined model for day-ahead forecasting.
Table 2 shows the forecasting precision and stability of the models during point forecasting.

3.3. Performance Metrics and Benchmark Model

Five fitting error indices and two main benchmark models were introduced to ensure the predictability of the proposed model. These indicators were MAE, RMSE, MAPE, STDAPE, DA, U1, U2, and R2 (the coefficient of determination), as well as seven ANN models. MAPE was a significant focus.

3.4. Parameter Setting

Parameter settings and initializations important to the methods applied included the following:
Input dimension. For the input vector, we set varying values of 1–8 for the input vector based on the ten-minute wind speed data and values of 1–6 for the input vector based on the 10-min wind power and wind speed data. The results of our empirical study indicated that the forecasting accuracy was at its best when the input vector dimension was 6.
SSA decomposition and EEMD denoising. For the previously mentioned SSA decomposition stage, the window length L and the principal components were the two most vital factors. On the basis of the trial-and-error method, L was set to 24 for ten-minute intervals to assess the wind speed transition and wind power from observation Sites 1 to Site 4 due to the homogeneity of the data structure among intervals. Thirteen principal components were selected for all sites. Table 3 shows the eigenvalue contributions to the various low- and high-frequency data groups in the time sequences. Taking the Site 1 wind speed series as an example, the principal components related to the first 13 eigenvalues accounted for 97.7643% of the variability, representing the main trend in the series. Further denoising was carried out for the high-frequency components. By parity of reasoning, the wind power time series data from the four sites were also analyzed by SSA using the trial-and-error and principal component analysis methods. The details are given in Table 3. For the EEMD denoising stage, the number of IMFs was acquired by the formula log2N − 1, where N is the length of the series. The IMFs ranged from high-frequency to low-frequency.
For MODA, the population size of the dragonfly was 40, the archive size was 500, and the maximum number of iterations was 200.
For each artificial neural network, the building and training of the network were nearly parallel. There were 5 input layer neurons, 20 initial hidden layer neurons, and 1 output layer neuron.

3.5. Experimental Results and Analysis

This section presents the results from the three experiments on the forecasting performance of the proposed novel model.

3.5.1. Experiment I: The Forecasting Performance of the Combined Models Optimized by Different Optimization Algorithms

Based on the historical wind speed and wind power time series, four experiments were designed to analyze and contrast the developed combined model with the sub-model forecasting models. Experiment I contrasted combined models with different numbers of sub-models in terms of their forecasting performances and analyzed the ability of the proposed model with five sub-models to forecast wind speed and wind power. The wind speed and wind power forecasting performance were evaluated by eight metrics.
For Site 1, the proposed combined model had the best forecasting performance for both wind speed forecasting and wind power forecasting. Specifically, the mean absolute percentage error (MAPE) value was 6.51% for wind speed forecasting and 13.21% for wind power forecasting, smaller than those of the other combined models and sub-models. The relevant characteristics of the models used for Site 1 are shown in Figure 2.
For Site 2, the best mean absolute error (MAE), root-mean-square error (RMSE), mean absolute percentage error (MAPE), and standard deviation of absolute percentage error (STDAPE) were all obtained with the developed combined method for wind speed forecasting, with values of 0.2608, 0.3783, 6.90%, and 6.41%, respectively. For wind power forecasting, the best forecasting metrics were obtained with the proposed combined model shown in Table 4. This confirms the good performance of the proposed combined model.
For Site 3, the wind speed forecasting accuracy of the combined model with five sub-models was significantly better than that of the combined models with less than five sub-models. This means that increasing the number of sub-models can improve the forecasting accuracy of the combined model.
The forecasting results of the combined model for Site 4 were similar to those obtained for previous sites; the proposed combined model provided better forecasting results in the first season. Taking wind power forecasting as an example, the lowest MAPE value, obtained with the advanced model, was 13.66%. The combined model with two sub-models had the worst forecasting result, having a MAPE value of 18.48% (the forecasting results of the sub-models are shown in Table A1). Table 4 also shows that the goodness of fit values for different combined models were over 0.9, indicating that combined models have a reduced forecasting error.
Remark 1.
A general survey of the forecasting results for wind speed and wind power in the first season showed that the proposed combined model hds an optimal wind speed forecasting capacity. MAPE values at one day ahead were 6.51%, 6.41%, 7.00%, and 6.16% for Sites 1 to 4, respectively. Meanwhile, for wind power forecasting, the proposed combined model had the lowest forecasting error among all involved models. Moreover, for the forecasting results of wind speed and wind power, the MAPE value of wind power was twice that of wind speed with similar goodness-of-fit values.

3.5.2. Experiment II: The Forecasting Performance of Combined Models Optimized by Different Optimization Algorithms

This experiment was designed to compare combined models optimized by different optimization algorithms, including MOGWO, MODA, MOMVO, and MODE. This experiment aimed to assess wind speed and wind energy forecasting in the second season. The evaluation metrics obtained for each model (MAE, RMSE, STDAPE, DA, U1, U2, MAPE and R2) are shown in Table 5 and Figure 3.
For Site 1, the combined model optimized by MODA for wind speed forecasting and the combined model optimized by MODA for wind power forecasting produced the most accurate forecasting results in the most efficient manner. The MAPE values of the optimal combined model were 8.965% and 19.506%, respectively. In comparison, the forecasting performances of the combined models optimized by four algorithms showed no significant differences; for example, the RMSE values of the four combined models were 0.4647, 0.4649, 0.4644, and 0.4648 for wind speed forecasting. The wind power forecasting results for each model are listed in Table 5.
For wind speed forecasting at Site 2, the combined model optimized by MODA obtained the best assessment metrics for wind speed forecasting. Specifically, for wind speed forecasting, the MAPE value obtained by the combined model optimized by MOGWO was 10.409%, better than the values obtained by the combined models optimized by MOGWO, MOMVO, and MODE by 0.0019%, 0.0019%, and 0.0048%. These models were ranked from second to last in terms of their forecasting accuracy, respectively. Figure 3 shows the wind power forecasting results for the combined models optimized by four optimization algorithms.
For Site 3, all of the combined models optimized by optimization algorithms provided satisfactory wind speed and wind power forecasting results, as proven by their better goodness-of-fit values compared with those of the sub-models (the forecasting results of sub-models shows in Table A2).
For Site 4, the combined models optimized by four different optimization algorithms showed outstanding forecasting potential. For wind power forecasting, the combined model optimized by MODA obtained the lowest MAE, RMSE, STDAPE, and MAPE values of 45.4039, 81.9297, 24.57%, and 14.226%, respectively. There were no obvious differences in MAPE among the other combined models.
Remark 2.
The evaluation index values obtained in Experiment II reveal that regardless of the optimization algorithm applied to optimize the combined model and the site used for forecasting, there is no optimal method for wind speed and wind power forecasting. For the second season of wind speed and wind power forecasting, the forecasting evaluation metrics of the combined models optimized by each algorithm had no significant differences.

3.5.3. Experiment III: Verification of the Performance of the Combined Model Based on Five Sub-Models and Optimized by MODA

In this experiment, 10 min intervals of wind speed and wind power forecasts in the third and fourth seasons were analyzed. The forecasting results for wind speed and wind power at four observation sites are shown in Figure 4 and Figure 5, and the forecasting performance of each model, as evaluated by eight metrics, is shown in Table 6 and Table 7. The results illustrate the following:
(1) The combined model with five sub-models produced the smallest MAE, RMSE, STDAPE, U1, U2, and MAPE values and the largest DA and R2 values at all sites. This indicates that the combined model has the best forecasting performance and stability.
(2) For wind power forecasting, the results of the combined model with five sub-models (CM5) for four sites, as evaluated by eight metrics, were better than those of the other combined models. For example, the RMSE values of the CM5 model at the four sites were 40.1238, 32.9801, 56.0705, and 69.8088. Compared with the other combined models, the forecasting accuracy of the CM5 was greater. The main reason for this is that the CM5 can effectively use the advantages of each sub-model and the weight of the combined model optimized by the optimization algorithm. By comparing the CM5 with the CM2, CM3 and CM4 models, we determined that the number of sub-models included influences the forecasting performance of the combined model. The MAPE value of the CM5 was 12.00% lower than those of the other combined models for Site 1 in the third season.
(3) For wind speed forecasting, the combined model obtained satisfactory forecasting results. The forecasting results of the combined models were better than those of the sub-models. Thus, the use of combined models can improve forecasting accuracy. For wind speed forecasting at Site 1, the MAPE values of the five sub-models were 14.61%, 16.31%, 18.46%, 21.04%, and 24.38%, respectively. In contrast, the MAPE value of the combined model (CM5) was 6.29%, an improvement of 8.32%, 10.02%, 12.17%, 14.75%, and 18.09%, respectively, compared with the five sub-models.
(4) The models’ goodness of fit results are shown in Table 6. The minimum value was 0.9665, which proves that the forecasting series obtained by combined model (CM5) was consistent with the actual series.
Remark 3.
The CM5 combined model had the smallest MAE, RMSE, STDAPE, U1, U2, and MAPE values and the biggest R2 and DA values. This indicates that combined models can identify changes in wind speed and wind power.
To verify the applicability of the combined model with five sub-models optimized by MODA for determining the wind speed and wind power series, the forecasting results for the fourth season are shown in Table 7 and Figure 5.
For Site 1, in the fourth season, the combined model with five sub-models (CM5) obtained the best forecasting accuracy. For wind speed and wind power forecasting, the evaluation metrics of the CM5 were also better than those of combined models with less than five sub-models and for the sub-models individually. This means that the CM5 is more suitable for wind speed and wind power forecasting.
For Site 2, the sub-models and combined models showed analogical forecasting abilities in terms of the values obtained for the forecasting evaluation metrics. To be more specific, for wind speed forecasting, the MAPE values of the five sub-models were 16.30%, 18.64%, 20.90%, 23.90% and 27.26%, respectively. In contrast, the MAPE value of the combined model (CM5) was 6.77%—an improvement of 9.53%, 11.87%, 14.13%, 17.13%, and 20.49%, respectively, compared with the five sub-models (shown in Table A4).
As for Site 3, the combined model with five sub-models showed a superior performance to the other combined models based on the eight employed evaluation criteria with MAE, RMSE, STDAPE, DA, U1, U2, MAPE, and R2 values for wind power forecasting of 26.4136, 48.6081, 31.34%, 81.83%, 0.0692, 0.9668, 14.28% and 0.9774, respectively. Among the remaining combined models, the ranking of methods in terms of their forecasting accuracy, from good to bad, was CM2, CM3, and CM4, with MAPE values of 19.27%, 17.61%, and 15.93%, respectively.
For Site 4, the MAPE values of the combined model (CM5) for wind speed and wind power forecasting were 7.86% and 33.05%, respectively. Compared with the other three combined models with the highest MAPE values, the combined model with five sub-models improved wind speed and wind power forecasting by 2.79% and 10.76%, respectively. Comparing the combined models and sub-models, the forecasting accuracy of the combined models was better. A comparison of the wind speed and wind power forecasting results for Site 4 is presented in Figure 5.
Remark 4.
The differences in the forecasting results obtained by the developed model and those obtained by the other individual models were significant. The evaluation indicator values obtained with the proposed model were more satisfactory, as they were lower than those computed with the contrasting models, regardless of the forecasting step. Hence, we conclude that the advanced combined model has a superior capacity relative to conventional individual models for short-term wind speed forecasting.

4. Discussion

In this subsection, a comprehensive discussion related to the proposed model is provided. This includes two parts: a significance test for forecasting values and forecasting errors and a forecasting uncertainty analysis.

4.1. Significance Test between Forecasting Values and Actual Data

To evaluate the significance level of forecasting errors between the proposed combined model and the other models, a classical hypothesis test method, the Diebold–Mariano Test [55], was used to measure the significance of prediction errors for different models. Theoretical support for this model is as follows.
Given a certain probability of coming to a wrong conclusion α, the original hypothesis H0 will not be rejected as long as the forecasting capacity of the proposed model possesses no evident distinction when contrasted with the comparison model; if the inverse is true, H0 will be rejected and H1 accepted. The hypothetical form is:
H 0 : E [ L ( e r r o r 1 ) ] = E [ L ( e r r o r 2 ) ] H 1 : E [ L ( e r r o r 1 ) ] E [ L ( e r r o r 2 ) ]
where L represents the loss function for forecasting errors, and error1 and error2 are the forecasting errors of the combined model with five sub-models and those of other combined models and single models, respectively.
Further, the DM statistical magnitude can be expressed by:
DM = i = 1 n ( L ( e r r o r 1 ) L ( e r r o r 2 ) ) / n S 2 / n s 2
In the above equation, S2 is the estimated variance of d = L ( e r r o r 1 ) L ( e r r o r 2 ) .
After obtaining the calculated DM value, a comparison between the DM statistics and a critical value Z α / 2 is conducted. Once the DM statistic is greater than Z α / 2 or less than Z α / 2 , the original hypothesis will be rejected. Moreover, it may be concluded that there is an evident distinction between our developed model and the compared model.
The Wilcoxon rank-sum test [56,57] is a nonparametric test that can be used to determine whether two independent samples have been selected from populations with the same distribution.
If h = 1, the null hypothesis that there is no difference between two samples at the 5% significance level is rejected.
If h = 0, there is a failure to reject the null hypothesis at the 5% significance level.
The results of the Diebold–Mariano test and Wilcoxon rank-sum test for wind speed and wind power forecasting are shown in Table 8. The following conclusions were made: First, the test results of DM test for wind speed showed little variation in the forecasting error among the different combined models, and the Wilcoxon rank-sum test results showed no difference between the forecasting results of each model and the actual wind speed at the 5% significance level. When the combined models were compared, the combined model with five sub-models was found to be evidently different from the other models at the 1% significance level for wind power forecasting. However, the DM test between the combined model and the sub-model revealed that some sub-models obtained DM values that were higher than the threshold at a 5% significance level. For example, in the first season, the SSAWD-Elman sub-model had a DM value of 1.6448 for wind power forecasting. The test results for the other sub-models are shown in Table A5. The Wilcoxon rank-sum test result of each combined model showed no differences between the actual wind power values and the forecasting results of each combined model. Meanwhile, the Wilcoxon rank-sum test results presented in Table A5 show that some sub-models rejected the null hypothesis at the 5% significance level, which indicates that the forecasting results of some sub-models differed from the actual data.
For the wind speed and wind energy forecasting error test results, there was little variation among the combined models at the 1% significance level, while there were greater differences between the combined models and the sub-models. Comparing the actual data with those forecast by each model, it was found that the wind speed data forecast by each sub-model and the real data represented the same sample at the 5% significance level. However, when the wind power forecasting values obtained with each model and the actual data were compared with the Wilcoxon rank-sum test result, they were found to represent two different samples at the 5% significance level. Based on these results, the wind speed forecasting accuracy was deemed to be higher than the wind power forecasting accuracy.

4.2. Uncertainty Analysis

Uncertainties in wind speed and wind power forecasting result in an imbalance between projected and actual time series, and this influences the operation reliability of wind farms [58]. Based on the forecasting results of a combined model for wind speed and wind power, interval forecasting was employed to analyze the level of uncertainty in wind speed and wind power forecasting and to assess possible threats for decision planners in terms of power dispatching.
More specifically, the uncertainty analysis used interval forecasting to analyze wind speed and wind power data sets from four seasons, each with a span of 10 min. By comparing the forecasting performances of different combined models and sub-models, the forecasting error of each model was analyzed in four seasons, and the differences in the prediction error were compared with the level of error in the proposed models, namely, the sub-model based on the ANN and the MODA-based combined model with different numbers of sub-models. Sub-models and combined models were used as contrasting models to verify the superiority of the proposed combined model in terms of interval forecasting. If the evaluation results of the proposed combined model were better than those of all comparison models under the same conditions, this would verify that the proposed model is better than combined models with less than five sub-models or sub-models alone in terms of uncertainty forecasting and stability forecasting [59].
To better identify the characteristics of the wind speed and wind power forecasting values from each model, a diverse probability density function including the t location-scale, stable distribution, logistic, and normal distribution functions was employed to fit the forecasting wind speed series and wind power series via maximum likelihood estimation (MLE) [60]. Referring to the results of the assessment index R2 shown in Table 9 and Figure 6, the t-location scale and stable function was selected as the most suitable probabilistic distribution for further interval forecasting.
The evaluation indices FICP, FINAW, and AWD were adopted to assess the interval forecasting results and to analyze the forecasting uncertainty of the four selected models [61]. Notably, the larger FICP values gave better analysis results, whereas larger FINAW and AWD values gave worse analysis results. The above three metrics and the lower and upper confines of the wind prediction values [62] are defined in Table 9, and the results of the uncertainty analysis are presented in Table 10. Furthermore, the expectation probability, defined as ( 1 α ) × 100 % , was set at 95%, 90%, and 85% to assess the forecasting uncertainty of the four selected models in the uncertainty analysis.
Table 11 shows that the combined model based on MODA and the five sub-models obtained satisfactory uncertainty analysis assessment results, which proves that the proposed model is superior to the other combined models. Using the wind power results for Site 1 as examples at the 15% significance level, the FICP values obtained from each combined model with different numbers of sub-models (MODA-CM2, MODA-CM3, MODA-CM4 and MODA-CM5) were found to be 77.58%, 79.71%, 81.97% and 82.96%, respectively, whereas the FINAW values were 0.0852, 0.0787, 0.0726 and 0.0694, respectively. Furthermore, the proposed model resulted in reductions in AWD values of 0.1119, 0.0645, and 0.0214, respectively, compared with the combined model with five sub-models.

5. Conclusions

As an important source of clean energy, the use of wind energy has undergone rapid development, becoming widespread in recent years. However, the irregularity and instability of wind speed series data greatly restricts the development of wind power generation. There is an urgent need to successfully and accurately forecast wind speed and wind power, solve the dispatch problem, and further improve the operation efficiency of the power market. In this study, a combined model for wind speed and wind power day-ahead forecasting was developed. First, the original wind speed and wind energy time series were preprocessed using the secondary denoising method. Then, nine different ANN and machine learning forecasting models were established to forecast the wind speed and wind energy time series after denoising.
By assessing the accuracy of the model validation set, the optimal sub-model was selected for the combined model. Finally, the combined weight of the combined model was optimized by the multi-objective optimization algorithm. Considering the wind speed forecasting results, the MAPE of the optimal combined model was between 5.86% and 11.92%, the R2 was over 0.94, and the RMSE was between 0.3379 and 0.7595. Furthermore, the corresponding values for wind power forecasting were in the range of 11.10% to 33.05%, over 0.97, and between 0.10102 and 0.33641, respectively. All of the experimental results indicate that the combined forecasting system has high levels of accuracy and stability for wind speed and wind power forecasting.

Author Contributions

The experiment, data analysis, and paper writing were conducted by Q.L.; the experiment and data analysis were completed by Q.L. and Q.Z.; supervision, paper writing, and editing were conducted by Q.Z. and G.Z.; validation, methodology, paper editing, and supervision were handled by Q.L. and G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by State Grid Corporation of China Science and Technology Project: SGGSKY00WYJS2000062.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Modified Multi-Objective Dragonfly Algorithm

In order to solve the problems associated with the MODA of easily falling into local optimal solutions and having slower convergence speeds, a steps-based strategy based on an exponential function and an elite opposition learning strategy were used for modification. First, the elite opposition learning strategy was used to generate a broader search scope and diversify the population identified as the next generation and to improve the global search capacity and search accuracy of the MODA [62]. Then, a steps-based strategy based on an exponential function was applied to enhance the convergence speed of the algorithm in the later stages. The specific implementation scheme is as follows:
(1)
Elite Opposition Learning Strategy
To improve the global search capacity and search accuracy of the MODA, the elite opposition learning strategy was applied to produce various elite opposition individuals and further develop an unplanned opposition population that can hunt in neighborhood space and enhance the local mining capacity. During the running of the MODA, the elite opposition learning strategy can define the elite dragonfly as the top dragonfly with the best fitness value. This is expressed by e x m t = [ e x m , 1 t , e x m , 2 t , , e x m , D t ] , m = 1 , 2 , , E N , where e x i , j t indicates the elite solutions corresponding to individuals x i , j t , t is the present iteration, and D represents the dimension of the algorithm space.
Moreover, an elite opposition solution e x ˜ i t = [ e x i , 1 t , e x i , 2 t , , e x i , D t ] with respect to a certain dragonfly in the current solution x i t = [ x i , 1 t , x i , 2 t , , x i , D t ] can be mathematically modeled by:
e x i , j t = k ( e a j t + e b j t ) x i , j t .
e a j t = m i n ( e x m , j t ) , e b j t = m a x ( e x m , j t ) .
e x i , j t = r a n d ( e b j t e a j t ) + e a j t ,   if   e x i , j t < L b i .
where i = 1 , 2 , , S N ,   j = 1 , 2 , , E N ,   k = r a n d ( 0 , 1 ) , SN represents the population size of the dragonflies, and, generally, EN represents the selected number of elite individuals with the value of SN*0.1. k represents a generalized argument that is subject to a uniform distribution. The elite opposition learning strategy is used in the basic MODA to effectively expand the search area, diversify the population, and enhance the optimization performance and global search capability.
(2)
Steps-based Strategy based on an Exponential Function
In a basic MODA, there are several parameters ( s , a , c , f , e ,   and   w ) , as described by Δ x t + 1 = ( s S i + a A i + c C i + f F i + e E i ) + w x t , which are adaptively adjusted at random; hence, the dragonfly individuals update themselves according to a stochastic linear step size in the course of iteration. This means that the position update of the dragonfly completely relies on the site of the present individual and the stochastic linear step size, which can be determined, respectively. Despite the above strategy being conducive to finding a globally optimal solution to some degree, it cannot ensure that the solution is optimal; thus, this situation causes a slow convergence rate [51]. To expedite the convergence rate of the MODA, the exponential step strategy is applied to replace the original linear step strategy. This means that an exponential function is added to the original step and generates an exponential function steps-based strategy. It should be noted that a vital parameter μ is adopted to renew the step, so that the local and global hunt capacities can be improved, and the convergence rate can be further accelerated. Hence, it is pivotal to set a suitable μ .
In our study, the steps were updated based on the following exponential function:
μ = ( r a n d 0.5 ) 2 r a n d
The enhanced step-size rule used is:
Δ = μ Δ x t + 1 = ( r a n d 0.5 ) 2 r a n d Δ x t + 1
Here, r a n d [ 0 , 1 ] is a stochastic constant, and Δ x t is the step size in the tth iteration.
The equation for updating the position vector of dragonflies is expressed by:
x t + 1 = x t + μ Δ x t + 1 = x t + ( r a n d 0.5 ) 2 r a n d Δ x t + 1
In the above, t is the current iteration, and x t is the position vector in the tth iteration.
Apparently, the updating speed of the step size accelerates as the number of iterations increases. At the beginning of an iteration, the advanced step can accomplish local hunting and discover a more optimal hunt space. During the medium and later periods in the course of iteration, the advanced step will accelerate the convergence rate, avoiding the local optimum and finding the optimal solution in the global hunt process.
Table A1. The forecasting performance of five optimal sub-models for wind speed and wind power in the first season.
Table A1. The forecasting performance of five optimal sub-models for wind speed and wind power in the first season.
SiteMetricWind Speed Forecasting ResultsWind Power Forecasting Results
LSTMWNNANFISRNNELMLSTMElmanBPNNRBFNNSVM
Site 1MAE0.71280.80470.93060.99521.122472.757884.719294.2429100.1841123.008
RMSE0.99691.04981.2131.3161.4909123.6352138.1371150.2004159.6466191.7045
STDAPE14.22%15.16%17.65%18.85%20.90%80.87%136.27%153.49%172.28%260.26%
DA53.23%47.27%44.39%47.47%44.59%50.05%44.49%45.08%43.40%40.81%
U10.09020.09510.10930.11820.13290.11440.12810.13890.14670.1749
U20.88810.8711.04131.0851.46720.49160.52180.58370.63571.7005
MAPE14.91%17.05%20.06%21.49%24.03%28.53%35.87%37.62%43.22%55.27%
R20.8960.88310.84510.81820.77710.94110.92570.91210.90220.8603
SiteMetricLSTMWNNELMRNNANFISLSTMElmanBPNNSVMRBFNN
Site 2MAE0.65920.75820.82990.96091.11355.250164.642368.544981.251892.9089
RMSE0.91571.01991.11011.28781.464994.3334101.4713107.9959129.0841146.0998
STDAPE15.65%18.03%18.53%20.95%26.19%105.16%234.08%255.50%348.94%427.78%
DA52.33%46.57%47.57%43.00%42.60%50.35%42.30%44.69%41.31%39.82%
U10.09260.10280.11180.12910.14590.10920.11690.12330.14760.1652
U21.06021.13151.28361.51891.89580.8270.79930.75070.74230.6758
MAPE15.81%18.63%20.35%23.44%27.72%31.61%45.31%52.43%62.54%75.86%
R20.90780.88570.86580.82210.77380.95750.94990.94320.9180.8969
SiteMetricLSTMRNNELMWNNANFISElmanLSTMBPNNRBFNNANFIS
Site 3MAE0.74010.85720.98911.14291.281786.0772102.83110.7717124.926141.6915
RMSE1.02651.12341.29261.47261.6495143.5346167.6258174.9203191.9277213.2843
STDAPE16.57%17.86%22.35%23.63%27.92%184.14%241.75%270.58%331.04%381.07%
DA52.04%46.38%43.50%40.32%42.01%52.14%44.89%44.79%45.38%41.81%
U10.09210.10090.11550.13090.14530.13670.15920.1650.17770.1965
U20.90281.01131.12261.21381.430.78370.81370.78460.74710.7233
MAPE16.43%18.98%22.27%25.56%29.17%38.51%48.10%55.41%64.25%75.54%
R20.89320.8720.83080.78420.7450.92160.8930.88340.86610.8373
SiteMetricANFISWNNELMRNNNARNNANFISBPNNRBFNNElmanGRNN
Site 4MAE0.7830.89311.03451.181.316690.0454107.0139117.4349132.8439154.0333
RMSE1.14621.22931.43231.62541.803152.6258173.0959180.8868213.9421228.7902
STDAPE18.73%21.71%25.73%27.44%28.79%64.24%79.57%123.60%153.34%149.54%
DA51.14%48.36%43.89%42.40%39.52%48.56%44.19%46.08%43.59%41.51%
U10.09140.0980.11360.12860.14220.12350.140.14510.16790.1794
U21.17561.02181.22421.42791.56870.95250.97080.9980.62640.7754
MAPE15.00%17.86%20.74%23.37%25.91%29.34%36.77%45.06%53.03%61.99%
R20.88210.86310.81560.77340.72720.92270.89950.89090.85980.8351
Table A2. The forecasting performance of five optimal sub-models for wind speed and wind power in the second season.
Table A2. The forecasting performance of five optimal sub-models for wind speed and wind power in the second season.
SiteMetricWind Speed Forecasting ResultsWind Power Forecasting Results
LSTMWNNRNNANFISELMLSTMElmanBPNNRBFNNSVM
Site 1MAE0.71280.80470.93060.99521.122472.757884.719294.2429100.1841123.008
RMSE0.99691.04981.2131.3161.4909123.6352138.1371150.2004159.6466191.7045
STDAPE14.22%15.16%17.65%18.85%20.90%80.87%136.27%153.49%172.28%260.26%
DA53.23%47.27%44.39%47.47%44.59%50.05%44.49%45.08%43.40%40.81%
U10.09020.09510.10930.11820.13290.11440.12810.13890.14670.1749
U20.88810.8711.04131.0851.46720.49160.52180.58370.63571.7005
MAPE14.91%17.05%20.06%21.49%24.03%28.53%35.87%37.62%43.22%55.27%
R20.8960.88310.84510.81820.77710.94110.92570.91210.90220.8603
SiteMetricLSTMELMRNNWNNANFISLSTMElmanBPNNSVMRBFNN
Site 2MAE0.65920.75820.82990.96091.11355.250164.642368.544981.251892.9089
RMSE0.91571.01991.11011.28781.464994.3334101.4713107.9959129.0841146.0998
STDAPE15.65%18.03%18.53%20.95%26.19%105.16%234.08%255.50%348.94%427.78%
DA52.33%46.57%47.57%43.00%42.60%50.35%42.30%44.69%41.31%39.82%
U10.09260.10280.11180.12910.14590.10920.11690.12330.14760.1652
U21.06021.13151.28361.51891.89580.8270.79930.75070.74230.6758
MAPE15.81%18.63%20.35%23.44%27.72%31.61%45.31%52.43%62.54%75.86%
R20.90780.88570.86580.82210.77380.95750.94990.94320.9180.8969
SiteMetricLSTMWNNELMRNNANFISLSTMElmanBPNNANFISSVM
Site 3MAE0.74010.85720.98911.14291.281786.0772102.83110.7717124.926141.6915
RMSE1.02651.12341.29261.47261.6495143.5346167.6258174.9203191.9277213.2843
STDAPE16.57%17.86%22.35%23.63%27.92%184.14%241.75%270.58%331.04%381.07%
DA52.04%46.38%43.50%40.32%42.01%52.14%44.89%44.79%45.38%41.81%
U10.09210.10090.11550.13090.14530.13670.15920.1650.17770.1965
U20.90281.01131.12261.21381.430.78370.81370.78460.74710.7233
MAPE16.43%18.98%22.27%25.56%29.17%38.51%48.10%55.41%64.25%75.54%
R20.89320.8720.83080.78420.7450.92160.8930.88340.86610.8373
SiteMetricANFISWNNELMRNNNARNNBPNNElmanLSTMANFISRBFNN
Site 4MAE0.7830.89311.03451.181.316690.0454107.0139117.4349132.8439154.0333
RMSE1.14621.22931.43231.62541.803152.6258173.0959180.8868213.9421228.7902
STDAPE18.73%21.71%25.73%27.44%28.79%64.24%79.57%123.60%153.34%149.54%
DA51.14%48.36%43.89%42.40%39.52%48.56%44.19%46.08%43.59%41.51%
U10.09140.0980.11360.12860.14220.12350.140.14510.16790.1794
U21.17561.02181.22421.42791.56870.95250.97080.9980.62640.7754
MAPE15.00%17.86%20.74%23.37%25.91%29.34%36.77%45.06%53.03%61.99%
R20.88210.86310.81560.77340.72720.92270.89950.89090.85980.8351
Table A3. The forecasting performance of five optimal sub-models for wind speed and wind power in the third season.
Table A3. The forecasting performance of five optimal sub-models for wind speed and wind power in the third season.
SiteMetricWind Speed Forecasting ResultsWind Power Forecasting Results
LSTMWNNANFISRNNELMLSTMElmanBPNNRBFNNSVM
Site 1MAE0.71280.80470.93060.99521.122472.757884.719294.2429100.1841123.008
RMSE0.99691.04981.2131.3161.4909123.6352138.1371150.2004159.6466191.7045
STDAPE14.22%15.16%17.65%18.85%20.90%80.87%136.27%153.49%172.28%260.26%
DA53.23%47.27%44.39%47.47%44.59%50.05%44.49%45.08%43.40%40.81%
U10.09020.09510.10930.11820.13290.11440.12810.13890.14670.1749
U20.88810.8711.04131.0851.46720.49160.52180.58370.63571.7005
MAPE14.91%17.05%20.06%21.49%24.03%28.53%35.87%37.62%43.22%55.27%
R20.8960.88310.84510.81820.77710.94110.92570.91210.90220.8603
SiteMetricLSTMELMRNNWNNANFISLSTMElmanBPNNRBFNNSVM
Site 2MAE0.65920.75820.82990.96091.11355.250164.642368.544981.251892.9089
RMSE0.91571.01991.11011.28781.464994.3334101.4713107.9959129.0841146.0998
STDAPE15.65%18.03%18.53%20.95%26.19%105.16%234.08%255.50%348.94%427.78%
DA52.33%46.57%47.57%43.00%42.60%50.35%42.30%44.69%41.31%39.82%
U10.09260.10280.11180.12910.14590.10920.11690.12330.14760.1652
U21.06021.13151.28361.51891.89580.8270.79930.75070.74230.6758
MAPE15.81%18.63%20.35%23.44%27.72%31.61%45.31%52.43%62.54%75.86%
R20.90780.88570.86580.82210.77380.95750.94990.94320.9180.8969
SiteMetricLSTMRNNWNNELMANFISLSTMElmanBPNNRBFNNSVM
Site 3MAE0.74010.85720.98911.14291.281786.0772102.83110.7717124.926141.6915
RMSE1.02651.12341.29261.47261.6495143.5346167.6258174.9203191.9277213.2843
STDAPE16.57%17.86%22.35%23.63%27.92%184.14%241.75%270.58%331.04%381.07%
DA52.04%46.38%43.50%40.32%42.01%52.14%44.89%44.79%45.38%41.81%
U10.09210.10090.11550.13090.14530.13670.15920.1650.17770.1965
U20.90281.01131.12261.21381.430.78370.81370.78460.74710.7233
MAPE16.43%18.98%22.27%25.56%29.17%38.51%48.10%55.41%64.25%75.54%
R20.89320.8720.83080.78420.7450.92160.8930.88340.86610.8373
SiteMetricANFISRNNWNNELMNARNNBPNNANFISElmanRBFNNGRNN
Site 4MAE0.7830.89311.03451.181.316690.0454107.0139117.4349132.8439154.0333
RMSE1.14621.22931.43231.62541.803152.6258173.0959180.8868213.9421228.7902
STDAPE18.73%21.71%25.73%27.44%28.79%64.24%79.57%123.60%153.34%149.54%
DA51.14%48.36%43.89%42.40%39.52%48.56%44.19%46.08%43.59%41.51%
U10.09140.0980.11360.12860.14220.12350.140.14510.16790.1794
U21.17561.02181.22421.42791.56870.95250.97080.9980.62640.7754
MAPE15.00%17.86%20.74%23.37%25.91%29.34%36.77%45.06%53.03%61.99%
R20.88210.86310.81560.77340.72720.92270.89950.89090.85980.8351
Table A4. The forecasting performance of five optimal sub-models for wind speed and wind power in the fourth season.
Table A4. The forecasting performance of five optimal sub-models for wind speed and wind power in the fourth season.
SiteMetricWind Speed Forecasting ResultsWind Power Forecasting Results
LSTMANFISRNNELMWNNLSTMElmanBPNNRBFNNSVM
Site 1MAE0.71280.80470.93060.99521.122472.757884.719294.2429100.1841123.008
RMSE0.99691.04981.2131.3161.4909123.6352138.1371150.2004159.6466191.7045
STDAPE14.22%15.16%17.65%18.85%20.90%80.87%136.27%153.49%172.28%260.26%
DA53.23%47.27%44.39%47.47%44.59%50.05%44.49%45.08%43.40%40.81%
U10.09020.09510.10930.11820.13290.11440.12810.13890.14670.1749
U20.88810.8711.04131.0851.46720.49160.52180.58370.63571.7005
MAPE14.91%17.05%20.06%21.49%24.03%28.53%35.87%37.62%43.22%55.27%
R20.8960.88310.84510.81820.77710.94110.92570.91210.90220.8603
SiteMetricLSTMRNNWNNELMANFISLSTMElmanBPNNSVMRBFNN
Site 2MAE0.65920.75820.82990.96091.11355.250164.642368.544981.251892.9089
RMSE0.91571.01991.11011.28781.464994.3334101.4713107.9959129.0841146.0998
STDAPE15.65%18.03%18.53%20.95%26.19%105.16%234.08%255.50%348.94%427.78%
DA52.33%46.57%47.57%43.00%42.60%50.35%42.30%44.69%41.31%39.82%
U10.09260.10280.11180.12910.14590.10920.11690.12330.14760.1652
U21.06021.13151.28361.51891.89580.8270.79930.75070.74230.6758
MAPE15.81%18.63%20.35%23.44%27.72%31.61%45.31%52.43%62.54%75.86%
R20.90780.88570.86580.82210.77380.95750.94990.94320.9180.8969
SiteMetricLSTMWNNRNNELMANFISElmanLSTMBPNNSVMRBFNN
Site 3MAE0.74010.85720.98911.14291.281786.0772102.83110.7717124.926141.6915
RMSE1.02651.12341.29261.47261.6495143.5346167.6258174.9203191.9277213.2843
STDAPE16.57%17.86%22.35%23.63%27.92%184.14%241.75%270.58%331.04%381.07%
DA52.04%46.38%43.50%40.32%42.01%52.14%44.89%44.79%45.38%41.81%
U10.09210.10090.11550.13090.14530.13670.15920.1650.17770.1965
U20.90281.01131.12261.21381.430.78370.81370.78460.74710.7233
MAPE16.43%18.98%22.27%25.56%29.17%38.51%48.10%55.41%64.25%75.54%
R20.89320.8720.83080.78420.7450.92160.8930.88340.86610.8373
SiteMetricANFISWNNELMRNNNARNNElmanBPNNRBFNNANFISGRNN
Site 4MAE0.7830.89311.03451.181.316690.0454107.0139117.4349132.8439154.0333
RMSE1.14621.22931.43231.62541.803152.6258173.0959180.8868213.9421228.7902
STDAPE18.73%21.71%25.73%27.44%28.79%64.24%79.57%123.60%153.34%149.54%
DA51.14%48.36%43.89%42.40%39.52%48.56%44.19%46.08%43.59%41.51%
U10.09140.0980.11360.12860.14220.12350.140.14510.16790.1794
U21.17561.02181.22421.42791.56870.95250.97080.9980.62640.7754
MAPE15.00%17.86%20.74%23.37%25.91%29.34%36.77%45.06%53.03%61.99%
R20.88210.86310.81560.77340.72720.92270.89950.89090.85980.8351
Table A5. The test result of two types of time series by each sub-model.
Table A5. The test result of two types of time series by each sub-model.
TypePeriod1st Season
SiteSite 1SiteSite 2SiteSite 3SiteSite 4
ModelDM TestWRS TestModelDM TestWRS TestModelDM TestWRS TestModelDM TestWRS Test
Wind SpeedLSTM2.0153 *0.3999 (1)LSTM2.2311 *0.5271 (1)LSTM2.1936 *0.5866 (1)ANFIS2.0345 *0.7851 (1)
WNN2.0911 *0.5008 (1)WNN2.3116 *0.4679 (1)RNN2.3051 *0.4623 (1)WNN2.0683 *0.8657 (1)
ANFIS2.0791 *0.6388 (1)ELM2.2490 *0.4506 (1)ELM2.2691 *0.9545 (1)ELM2.0176 *0.8652 (1)
RNN2.0237 *0.8169 (1)RNN2.2329 *0.5504 (1)WNN2.2762 *0.7556 (1)RNN2.0892 *0.8939 (1)
ELM2.0347 *0.9344 (1)ANFIS2.2511 *0.9087 (1)ANFIS2.2555 *0.4812 (1)NARNN2.0554 *0.9906 (1)
Wind PowerLSTM1.6448 **0.5148 (1)LSTM1.8413 **0.1658 (1)Elman2.1714 *0.469 (1)ANFIS1.8767 **0.5515 (1)
Elman1.8584 **0.6926 (1)Elman2.0113 *0.3029 (1)LSTM2.1891 *0.428 (1)BPNN1.8933 **0.7224 (1)
BPNN1.8545 **0.7135 (1)BPNN1.9819 *0.3315 (1)BPNN2.1510 *0.8385 (1)RBFNN1.9811 *0.9577 (1)
RBFNN1.8092 **0.6182 (1)SVM1.9397 **0.4709 (1)RBFNN2.1536 *0.883 (1)Elman1.8210 **0.8973 (1)
SVM1.6948 **0.9221 (1)RBFNN1.9759 *0.6006 (1)ANFIS2.1529 *0.8701 (1)GRNN1.9347 **0.6005 (1)
TypePeriod2nd Season
Wind SpeedLSTM2.1945 *0.6758 (1)LSTM2.2025 *0.4777 (1)LSTM2.1799 *0.5881 (1)ANFIS2.1549 *0.9343 (1)
WNN2.0491 *0.7707 (1)ELM2.1907 *0.3819 (1)WNN2.1802 *0.6535 (1)WNN2.1582 *0.8213 (1)
RNN2.1712 *0.9856 (1)RNN2.1570 *0.4754 (1)ELM2.1139 *0.7227 (1)ELM2.1066 *0.4826 (1)
ANFIS2.2617 *0.9507 (1)WNN2.1710 *0.6418 (1)RNN2.1974 *0.8559 (1)RNN2.1212 *0.3952 (1)
ELM2.2046 *0.7191 (1)ANFIS2.1974 *0.6615 (1)ANFIS2.2154 *0.8648 (1)NARNN2.1212 *0.5641 (1)
Wind PowerLSTM1.8129 **0.6877 (1)LSTM1.9293 *0.2834 (1)LSTM1.8850 **0.191 (1)BPNN1.9729 *0.8058 (1)
Elman1.7532 **0.5882 (1)Elman1.8266 **0.5608 (1)Elman1.7569 **0.652 (1)Elman1.9529 **0.7731 (1)
BPNN1.7520 **0.6515 (1)BPNN1.7419 **0.8147 (1)BPNN1.7442 **0.6615 (1)LSTM1.9488 **0.8387 (1)
RBFNN1.7478 **0.8189 (1)SVM1.9149 **0.9142 (1)ANFIS1.7942 **0.8653 (1)ANFIS1.9671 *0.796 (1)
SVM1.7491 **0.8162 (1)RBFNN1.7790 *0.9952 (1)SVM1.8647 **0.8221 (1)RBFNN1.9611 *0.7553 (1)
TypePeriod3rd Season
Wind SpeedLSTM2.1481 *0.3747 (1)LSTM2.1868 *0.403 (1)LSTM2.2618 *0.3518 (1)ANFIS2.0569 *0.5237 (1)
WNN2.2647 *0.6465 (1)ELM2.1580 *0.4889 (1)RNN2.2334 *0.5056 (1)RNN2.0278 *0.814 (1)
ANFIS2.2909 *0.7413 (1)RNN2.0821 *0.5868 (1)WNN2.2084 *0.9067 (1)WNN2.0900 *0.6827 (1)
RNN2.2743 *0.7856 (1)WNN2.1309 *0.979 (1)ELM2.1865 *0.3958 (1)ELM2.0578 *0.8486 (1)
ELM2.2512 *0.6406 (1)ANFIS2.1179 *0.8809 (1)ANFIS2.1745 *0.9007 (1)NARNN2.0572 *0.7602 (1)
Wind PowerLSTM1.8928 **0.4611 (1)LSTM1.6638 **0.3403 (1)LSTM2.0381 *0.0666 (1)BPNN1.7334 **0.2531 (1)
Elman1.8541 **0.5106 (1)Elman1.6214 ***0.185 (1)Elman2.1068 *0.1145 (1)ANFIS1.8031 **0.5048 (1)
BPNN1.8038 **0.4511 (1)BPNN1.7245 **0.7108 (1)BPNN2.0731 *0.5217 (1)Elman1.8146 **0.9873 (1)
RBFNN1.9330 **0.8213 (1)RBFNN1.6877 **0.7228 (1)RBFNN2.0916 *0.9811 (1)RBFNN1.9499 **0.7887 (1)
SVM1.8470 **0.7011 (1)SVM1.7908 **0.8306 (1)SVM2.1666 *0.6872 (1)GRNN1.8161 **0.9878 (1)
TypePeriod4th Season
Wind SpeedLSTM2.2046 *0.8297 (1)LSTM2.1795 *0.5349 (1)LSTM2.1252 *0.3748 (1)ANFIS2.1443 *0.2156 (1)
ANFIS2.2263 *0.9765 (1)RNN2.1626 *0.6195 (1)WNN2.1970 *0.4513 (1)WNN2.1105 *0.2088 (1)
RNN2.2232 *0.9149 (1)WNN2.1002 *0.8759 (1)RNN2.1791 *0.9334 (1)ELM2.0785 *0.6151 (1)
ELM2.1759 *0.953 (1)ELM2.1286 *0.9552 (1)ELM2.1603 *0.7556 (1)RNN2.1600 *0.8181 (1)
WNN2.2367 *0.8119 (1)ANFIS2.1282 *0.571 (1)ANFIS2.1505 *0.369 (1)NARNN2.2729 *0.4898 (1)
Wind PowerLSTM2.1095 *0.617 (1)LSTM2.1545 *0.2666 (1)Elman1.8865 **0.1524 (1)Elman2.0134 *0.3397 (1)
Elman2.0571 *0.7218 (1)Elman2.0155 *0.2679 (1)LSTM1.8103 **0.2156 (1)BPNN2.1494 *0.4345 (1)
BPNN2.1123 *0.8302 (1)BPNN2.0564 *0.442 (1)BPNN1.8667 **0.6435 (1)RBFNN2.0748 *0.8048 (1)
RBFNN2.0746 *0.9939 (1)SVM1.9941 **0.7566 (1)SVM1.8281 **0.9013 (1)ANFIS2.0411 *0.8842 (1)
SVM2.0468 *0.9744 (1)RBFNN1.9628 *0.992 (1)RBFNN1.8860 **0.9017 (1)GRNN2.0735 *0.6124 (1)
Note: * is at the 10% significance level. ** is at the 5% significance level. *** is at the 1% significance level. (0) there is difference between two sample at the 5% significance level. (1) there is no difference between two sample at the 5% significance level.

References

  1. GWEC|GLOBAL WIND REPORT 2021. 2021. Available online: https://gwec.net/wp-content/uploads/2021/03/GWEC-Global-Wind-Report-2021.pdf (accessed on 5 May 2021).
  2. Liu, H.; Tian, H.-Q.; Li, Y.-F. Four wind speed multi-step forecasting models using extreme learning machines and signal de-composing algorithms. Energy Convers. Manag. 2015, 100, 16–22. [Google Scholar] [CrossRef]
  3. Selig, M.S.; Tangler, J.L. Development of a multipoint inverse design method for horizontal axis wind turbines. Wind. Eng. 1995, 19, 91–105. [Google Scholar]
  4. Lee, S. Inverse design of horizontal axis wind turbine blades using a vortex line method. Wind Energy 2015, 18, 253–266. [Google Scholar] [CrossRef] [Green Version]
  5. Moghadassian, B.; Sharma, A. Inverse Design of Single-A and Multi-Rotor Horizontal Axis Wind Turbine Blades Using Com-Putational Fluid Dynamics. J. Sol. Energy Eng. 2017, 140, 021003. [Google Scholar] [CrossRef] [Green Version]
  6. Tahani, M.; Kavari, G.; Masdari, M.; Mirhosseini, M. Aerodynamic design of horizontal axis wind turbine with innovative local linearization of chord and twist distributions. Energy 2017, 131, 78–91. [Google Scholar] [CrossRef]
  7. Liu, X.; Wang, L.; Tang, X. Optimized linearization of chord and twist angle profiles for fixed-pitch fixed-speed wind turbine blades. Renew. Energy 2013, 57, 111–119. [Google Scholar] [CrossRef]
  8. Soman, S.S.; Zareipour, H.; Malik, O.; Mandal, P. A review of wind power and wind speed forecasting methods with different timehorizons. In Proceedings of the North-American Power Symposium 2010, Arlington, TX, USA, 26–28 September 2010; pp. 1–7. [Google Scholar] [CrossRef]
  9. Zhang, S.; Chen, Y.; Xiao, J.; Zhang, W.; Feng, R. Hybrid wind speed forecasting model based on multivariate data secondary decomposition approach and deep learning algorithm with attention mechanism. Renew. Energy 2021, 174, 688–704. [Google Scholar] [CrossRef]
  10. Jacondino, W.D.; Nascimento, A.L.D.S.; Calvetti, L.; Fisch, G.; Beneti, C.A.A.; da Paz, S.R. Hourly day-ahead wind power forecasting at two wind farms in northeast Brazil using WRF model. Energy 2021, 230, 120841. [Google Scholar] [CrossRef]
  11. Pearre, N.S.; Swan, L.G. Statistical approach for improved wind speed forecasting for wind power production. Sustain. Energy Technol. Assess. 2018, 27, 180–191. [Google Scholar] [CrossRef]
  12. Jahangir, H.; Golkar, M.A.; Alhameli, F.; Mazouz, A.; Ahmadian, A.; Elkamel, A. Short-term wind speed forecasting framework based on stacked denoising auto-encoders with rough ANN. Sustain. Energy Technol. Assess. 2020, 38, 100601. [Google Scholar] [CrossRef]
  13. Dong, Y.; Zhang, L.; Liu, Z.; Wang, J. Integrated Forecasting Method for Wind Energy Management: A Case Study in China. Processes 2019, 8, 35. [Google Scholar] [CrossRef] [Green Version]
  14. Jiang, P.; Liu, Z.; Niu, X.; Zhang, L. A combined forecasting system based on statistical method, artificial neural networks, and deep learning methods for short-term wind speed forecasting. Energy 2021, 217, 119361. [Google Scholar] [CrossRef]
  15. Brabec, M.; Craciun, A.; Dumitrescu, A. Hybrid numerical models for wind speed forecasting. J. Atmos. Sol. Terr. Phys. 2021, 220, 105669. [Google Scholar] [CrossRef]
  16. Zhao, J.; Guo, Z.; Guo, Y.; Lin, W.; Zhu, W. A self-organizing forecast of day-ahead wind speed: Selective ensemble strategy based on numerical weather predictions. Energy 2021, 218, 119509. [Google Scholar] [CrossRef]
  17. Al-Yahyai, S.; Charabi, Y.; Gastli, A. Review of the use of Numerical Weather Prediction (NWP) Models for wind energy as-sessment. Renew. Sustain. Energy Rev. 2010, 14, 3192–3198. [Google Scholar] [CrossRef]
  18. Mu, M.U.; Chen, B. Methods and Uncertainties of Meteorological Forecast. Meteorol. Mon. 2011, 37, 1–13. [Google Scholar] [CrossRef]
  19. Khodayar, M.; Wang, J. Spatio-Temporal Graph Deep Neural Network for Short-Term Wind Speed Forecasting. IEEE Trans. Sustain. Energy 2018, 10, 670–681. [Google Scholar] [CrossRef]
  20. Lu, P.; Ye, L.; Zhong, W.; Qu, Y.; Zhai, B.; Tang, Y.; Zhao, Y. A novel spatio-temporal wind power forecasting framework based on multi-output support vector machine and optimization strategy. J. Clean. Prod. 2020, 254, 119993. [Google Scholar] [CrossRef]
  21. Sun, M.; Feng, C.; Zhang, J. Conditional aggregated probabilistic wind power forecasting based on spatio-temporal correlation. Appl. Energy 2019, 256, 113842. [Google Scholar] [CrossRef]
  22. Liu, Y.; Qin, H.; Zhang, Z.; Pei, S.; Jiang, Z.; Feng, Z.; Zhou, J. Probabilistic spatiotemporal wind speed forecasting based on a variational Bayesian deep learning model. Appl. Energy 2020, 260, 114259. [Google Scholar] [CrossRef]
  23. Ma, X.; Jin, Y.; Dong, Q. A generalized dynamic fuzzy neural network based on singular spectrum analysis optimized by brain storm optimization for short-term wind speed forecasting. Appl. Soft Comput. 2017, 54, 296–312. [Google Scholar] [CrossRef]
  24. Niu, T.; Wang, J.; Zhang, K.; Du, P. Multi-step-ahead wind speed forecasting based on optimal feature selection and a modified bat algorithm with the cognition strategy. Renew. Energy 2018, 118, 213–229. [Google Scholar] [CrossRef]
  25. Iversen, E.B.; Morales, J.; Møller, J.K.; Madsen, H. Short-term probabilistic forecasting of wind speed using stochastic differential equations. Int. J. Forecast. 2016, 32, 981–990. [Google Scholar] [CrossRef]
  26. Torres, J.; García, A.; de Blas, M.; Francisco, A. Forecast of hourly average wind speed with ARMA models in Navarre (Spain). Sol. Energy 2005, 79, 65–77. [Google Scholar] [CrossRef]
  27. Erdem, E.; Shi, J. ARMA based approaches for forecasting the tuple of wind speed and direction. Appl. Energy 2011, 88, 1405–1414. [Google Scholar] [CrossRef]
  28. Shukur, O.B.; Lee, M. Daily wind speed forecasting through hybrid KF-ANN model based on ARIMA. Renew. Energy 2015, 76, 637–647. [Google Scholar] [CrossRef]
  29. Yang, D.; Sharma, V.; Ye, Z.; Lim, L.I.; Zhao, L.; Aryaputera, A.W. Forecasting of global horizontal irradiance by exponential smoothing, using decompositions. Energy 2015, 81, 111–119. [Google Scholar] [CrossRef] [Green Version]
  30. Zuluaga, C.; Álvarez, M.A.; Giraldo, E. Short-term wind speed prediction based on robust Kalman filtering: An experimental comparison. Appl. Energy 2015, 156, 321–330. [Google Scholar] [CrossRef]
  31. Cavalcante, L.; Bessa, R.J.; Reis, M.; Browell, J. LASSO vector autoregression structures for very short-term wind power forecasting. Wind Energy 2017, 20, 657–675. [Google Scholar] [CrossRef] [Green Version]
  32. Ziel, F.; Croonenbroeck, C.; Ambach, D. Forecasting wind power—Modeling periodic and non-linear effects under conditional heteroscedasticity. Appl. Energy 2016, 177, 285–297. [Google Scholar] [CrossRef] [Green Version]
  33. Bessa, R.; Trindade, A.; Miranda, V. Spatial-Temporal Solar Power Forecasting for Smart Grids. IEEE Trans. Ind. Informatics 2015, 11, 232–241. [Google Scholar] [CrossRef]
  34. Wang, J.; Heng, J.; Xiao, L.; Wang, C. Research and application of a combined model based on multi-objective optimization for multi-step ahead wind speed forecasting. Energy 2017, 125, 591–613. [Google Scholar] [CrossRef]
  35. Li, X.; Yang, D.; Yang, J.; Zheng, G.; Han, G.; Nan, Y.; Li, W. Analysis of coastal wind speed retrieval from CYGNSS mission using artificial neural network. Remote Sens. Environ. 2021, 260, 112454. [Google Scholar] [CrossRef]
  36. Prasad, S.; Nguyen-Huy, T.; Deo, R. Support vector machine model for multistep wind speed forecasting. Predict. Model. Energy Manag. Power Syst. Eng. 2020, 335–389. [Google Scholar] [CrossRef]
  37. Kalajdjieski, J.; Zdravevski, E.; Corizzo, R.; Lameski, P.; Kalajdziski, S.; Pires, I.; Garcia, N.; Trajkovik, V. Remote Sensing Air Pollution Prediction with Multi-Modal Data and Deep Neural Networks. Remote Sens. 2020, 12, 4142. [Google Scholar] [CrossRef]
  38. Jiang, P.; Yang, H.; Heng, J. A hybrid forecasting system based on fuzzy time series and multi-objective optimization for wind speed forecasting. Appl. Energy 2019, 235, 786–801. [Google Scholar] [CrossRef]
  39. Zhao, J.; Guo, Z.-H.; Su, Z.-Y.; Zhao, Z.-Y.; Xiao, X.; Liu, F. An improved multi-step forecasting model based on WRF ensembles and creative fuzzy systems for wind speed. Appl. Energy 2016, 162, 808–826. [Google Scholar] [CrossRef]
  40. Naik, J.; Dash, S.; Dash, P.; Bisoi, R. Short term wind power forecasting using hybrid variational mode decomposition and multi-kernel regularized pseudo inverse neural network. Renew. Energy 2018, 118, 180–212. [Google Scholar] [CrossRef]
  41. Liu, Z.; Hara, R.; Kita, H. Hybrid forecasting system based on data area division and deep learning neural network for short-term wind speed forecasting. Energy Convers. Manag. 2021, 238, 114136. [Google Scholar] [CrossRef]
  42. Aly, H.H. A proposed intelligent short-term load forecasting hybrid models of ANN, WNN and KF based on clustering techniques for smart grid. Electr. Power Syst. Res. 2020, 182, 106191. [Google Scholar] [CrossRef]
  43. Wang, D.; Luo, H.; Grunder, O.; Lin, Y. Multi-step ahead wind speed forecasting using an improved wavelet neural network combining variational mode decomposition and phase space reconstruction. Renew. Energy 2017, 113, 1345–1358. [Google Scholar] [CrossRef]
  44. Huang, X.; Wang, J.; Huang, B. Two novel hybrid linear and nonlinear models for wind speed forecasting. Energy Convers. Manag. 2021, 238, 114162. [Google Scholar] [CrossRef]
  45. Du, P.; Wang, J.; Guo, Z.; Yang, W. Research and application of a novel hybrid forecasting system based on multi-objective optimization for wind speed forecasting. Energy Convers. Manag. 2017, 150, 90–107. [Google Scholar] [CrossRef]
  46. Jiang, P.; Wang, Y.; Wang, J. Short-term wind speed forecasting using a hybrid model. Energy 2017, 119, 561–577. [Google Scholar] [CrossRef]
  47. Hu, J.; Wang, J.; Xiao, L. A hybrid approach based on the Gaussian process with t-observation model for short-term wind speed forecasts. Renew. Energy 2017, 114, 670–685. [Google Scholar] [CrossRef]
  48. Corizzo, R.; Ceci, M.; Fanaee-T, H.; Gama, J. Multi-aspect renewable energy forecasting. Inf. Sci. 2021, 546, 701–722. [Google Scholar] [CrossRef]
  49. Cai, C.; Tao, Y.; Zhu, T.; Deng, Z. Short-Term Load Forecasting Based on Deep Learning Bidirectional LSTM Neural Network. Appl. Sci. 2021, 11, 8129. [Google Scholar] [CrossRef]
  50. Han, Q.; Meng, F.; Hu, T.; Chu, F. Non-parametric hybrid models for wind speed forecasting. Energy Convers. Manag. 2017, 148, 554–568. [Google Scholar] [CrossRef]
  51. Zhou, Q.; Wang, C.; Zhang, G. A combined forecasting system based on modified multi-objective optimization and sub-model selection strategy for short-term wind speed. Appl. Soft Comput. 2020, 94, 106463. [Google Scholar] [CrossRef]
  52. Wang, C.; Zhang, H.; Ma, P. Wind power forecasting based on singular spectrum analysis and a new hybrid Laguerre neural network. Appl. Energy 2020, 259, 114139. [Google Scholar] [CrossRef]
  53. Wang, S.; Zhang, N.; Wu, L.; Wang, Y. Wind speed forecasting based on the hybrid ensemble empirical mode decomposition and GA-BP neural network method. Renew. Energy 2016, 94, 629–636. [Google Scholar] [CrossRef]
  54. Song, J.; Wang, J.; Lu, H.Y. A novel combined model based on advanced optimization algorithm for short-term wind speed forecasting. Appl. Energy 2018, 215, 643–658. [Google Scholar] [CrossRef]
  55. Wang, C.; Zhang, S.; Xiao, L.; Fu, T. Wind speed forecasting based on multi-objective grey wolf optimisation algorithm, weighted information criterion, and wind energy conversion system: A case study in Eastern China. Energy Convers. Manag. 2021, 243, 114402. [Google Scholar] [CrossRef]
  56. Haynes, W. Wilcoxon Rank Sum Test. In Encyclopedia of Systems Biology; Dubitzky, W., Wolkenhauer, O., Cho, K.-H., Yokota, H., Eds.; Springer: New York, NY, USA, 2013; pp. 2354–2355. [Google Scholar]
  57. Jiang, P.; Li, R.; Li, H. Multi-objective algorithm for the design of prediction intervals for wind power forecasting model. Appl. Math. Model. 2019, 67, 101–122. [Google Scholar] [CrossRef]
  58. Yang, W.; Wang, J.; Lu, H.; Niu, T.; Du, P. Hybrid wind energy forecasting and analysis system based on divide and conquer scheme: A case study in China. J. Clean. Prod. 2019, 222, 942–959. [Google Scholar] [CrossRef] [Green Version]
  59. Jiang, P.; Li, C.; Li, R.; Yang, H. An innovative hybrid air pollution early-warning system based on pollutants forecasting and Extenics evaluation. Knowledge-Based Syst. 2019, 164, 174–192. [Google Scholar] [CrossRef]
  60. Wang, J.; Niu, T.; Lu, H.; Yang, W.; Du, P. A Novel Framework of Reservoir Computing for Deterministic and Probabilistic Wind Power Forecasting. IEEE Trans. Sustain. Energy 2019, 11, 337–349. [Google Scholar] [CrossRef]
  61. Tian, C.; Hao, Y. Point and interval forecasting for carbon price based on an improved analysis-forecast system. Appl. Math. Model. 2020, 79, 126–144. [Google Scholar] [CrossRef]
  62. Liu, Z.; Jiang, P.; Zhang, L.; Niu, X. A combined forecasting model for time series: Application to short-term wind speed fore-casting. Appl. Energy 2019, 259, 114137. [Google Scholar] [CrossRef]
Figure 1. Flow Chart of the Proposed Model.
Figure 1. Flow Chart of the Proposed Model.
Applsci 11 09383 g001
Figure 2. Wind speed and wind power forecasting result in first season.
Figure 2. Wind speed and wind power forecasting result in first season.
Applsci 11 09383 g002
Figure 3. The optimal combined model for wind power and wind speed forecasting.
Figure 3. The optimal combined model for wind power and wind speed forecasting.
Applsci 11 09383 g003
Figure 4. Wind speed and wind power forecasting result for each site in the third season.
Figure 4. Wind speed and wind power forecasting result for each site in the third season.
Applsci 11 09383 g004
Figure 5. Wind speed and wind power forecasting result for each site by the CM5 model.
Figure 5. Wind speed and wind power forecasting result for each site by the CM5 model.
Applsci 11 09383 g005
Figure 6. The optimal distribution of forecasting error in CM5.
Figure 6. The optimal distribution of forecasting error in CM5.
Applsci 11 09383 g006
Table 1. Statistical descriptions for wind speed and wind power.
Table 1. Statistical descriptions for wind speed and wind power.
TypePeriodSiteMeanStandard DeviationSkewnessKurtosisMinimumMaximumMedian
Wind Power1st seasonSite 1411.7689363.52441.09213.53631.80001531.5000306.4000
Site 2305.9613321.89971.68865.52760.30001513.3000183.2500
Site 3389.7726365.30301.45644.34961.30001536.0000258.8000
Site 4492.9215393.05590.77322.56403.30001523.5000385.3000
2nd seasonSite 1489.9329410.85070.71602.39363.30001526.0000385.2500
Site 2265.8611283.62762.05747.60691.80001525.5000158.1500
Site 3434.1101399.52051.23233.49042.50001540.5000295.0500
Site 4564.8915480.97060.55641.802110.50001531.5000411.0000
3rd seasonSite 1424.9078381.97281.01053.05953.30001526.0000291.0500
Site 2276.0506292.74751.86526.58991.80001525.5000153.9000
Site 3592.1482446.71480.60882.09075.30001542.5000456.8000
Site 4488.0642390.17390.90793.02997.80001514.5000398.3000
4th seasonSite 1380.6880324.98170.81022.71742.50001462.3000277.6500
Site 2283.2654280.04991.87466.51630.30001517.8000179.5000
Site 3268.0734230.09431.58415.96101.50001346.5000195.9000
Site 4421.4685345.85561.07793.44310.50001526.0000326.9000
Wind Speed1st seasonSite 15.14812.21850.45632.46001.200011.60004.9000
Site 24.53492.16640.98983.58581.400012.60004.0000
Site 35.14882.25170.69682.85101.200012.00004.7000
Site 45.85182.41140.34642.33571.000012.40005.6000
2nd seasonSite 14.93072.28210.31522.02101.100011.00004.7000
Site 23.92321.76901.08343.93261.100010.50003.5000
Site 34.85792.21170.65512.91341.000012.70004.5500
Site 45.66592.65780.36301.93761.300011.80005.2500
3rd seasonSite 14.59101.92140.41642.32721.300010.30004.4000
Site 24.07301.50740.79483.53701.300010.70003.8000
Site 35.04991.83180.08512.51250.800010.40005.1000
Site 46.05692.41370.04212.12701.000011.90006.0000
4th seasonSite 14.50071.89740.39442.13001.20009.80004.2000
Site 24.10781.76221.23934.91651.000012.10003.7000
Site 34.11161.59630.59482.79391.20009.70003.8000
Site 44.89701.96370.42112.57760.800010.90004.7000
Note: Bold represents kurtosis values less than 3, indicating a slight tail distribution in the time series.
Table 2. The forecasting metric for each model.
Table 2. The forecasting metric for each model.
MetricDefinitionEquation
MAEThe mean absolute error of N forecasting results MAE = 1 N n = 1 N | y n y ^ n |
RMSEThe root mean square error of N forecasting results RMSE = 1 N n = 1 N ( y n y ^ n ) 2
MAPEThe mean absolute percentage error of N forecasting results MAPE = 1 N n = 1 N | y n y ^ n | y n
STD of APEThe standard deviation of absolute percentage error of N forecasting results STDAPE = Var ( | y n y ^ n | y n )
R2The goodness-of-forecasting fit R 2 = n = 1 N ( y n y ¯ ) n = 1 N ( y n y ^ n ) n = 1 N ( y n y ¯ )
DADirections or turning points between actual and forecasting values DA = 100 N 1 n = 1 N 1 m t
U1U-Statistic of 1-order U 1 = 1 N n = 1 N ( y n y ^ n ) 2 1 N n = 1 N ( y n ) 2 + 1 N n = 1 N ( y ^ n ) 2
U2U-Statistic of 2-order U 2 = 1 N n = 1 N ( y n + 1 y ^ n + 1 y n ) 2 1 N n = 1 N ( y n + 1 y ^ n y n ) 2
Table 3. Results for four sites after SSA.
Table 3. Results for four sites after SSA.
Data SetsFrequencySite 1Site 2
Eigenvalue i = 1 r λ i / i = 1 d λ i Eigenvalue i = 1 r λ i / i = 1 d λ i
Wind SpeedLow1–1397.76431–1395.5416
High14–242.235714–244.4584
Wind PowerLow1–1396.28041–1395.5364
High14–243.719614–244.4636
Data SetsFrequencySite 3Site 4
Eigenvalue i = 1 r λ i / i = 1 d λ i Eigenvalue i = 1 r λ i / i = 1 d λ i
Wind SpeedLow1–1398.09891–1399.1851
High14–241.901114–240.8149
Wind PowerLow1–1394.39941–1392.8841
High14–245.600614–247.1159
Table 4. Wind speed forecasting results for each site by different models for the first season.
Table 4. Wind speed forecasting results for each site by different models for the first season.
MetricWind Speed Forecasting ResultWind Power Forecasting Result
MODA-CM2MODA-CM3MODA-CM4MODA-CM5MODA-CM2MODA-CM3MODA-CM4MODA-CM5
MAE0.40620.37190.33540.300140.05636.56133.068629.6338
RMSE0.57210.52510.47240.423572.793366.398359.912453.9345
STDAPE9.23%8.45%7.62%6.83%51.54%46.59%42.60%38.78%
DA77.36%79.44%81.63%83.32%75.97%78.15%79.94%81.93%
U10.0510.04680.04210.03780.06640.06050.05460.0492
U20.94160.93130.92280.91660.79760.80110.80820.8168
MAPE8.81%8.06%7.27%6.51%17.78%16.21%14.68%13.21%
R20.96630.97160.97710.98160.97980.98320.98630.9879
MAE0.35240.32210.29170.260827.961125.444523.08220.5847
RMSE0.50980.46690.42340.378347.948243.526739.626735.2723
STDAPE9.39%8.57%7.77%6.90%93.10%85.07%82.63%71.73%
DA78.05%79.74%82.03%85.00%75.47%77.66%79.05%81.13%
U10.05080.04660.04220.03770.05420.04920.04480.0399
U20.98730.97190.95960.95020.92340.93010.93450.9415
MAPE8.67%7.93%7.17%6.41%19.52%17.81%16.31%14.49%
R20.97190.97650.98070.98460.98880.99080.99240.994
MAE0.42580.38810.35210.314542.918939.250935.489531.7399
RMSE0.60750.55410.50360.450476.010669.603962.679656.1495
STDAPE10.77%9.83%8.91%7.97%42.11%40.08%34.50%32.30%
DA78.15%79.94%81.63%83.12%77.56%79.74%81.63%82.72%
U10.05410.04940.04490.04010.0710.06510.05860.0525
U20.87350.86450.86540.86880.9630.96660.96910.9719
MAPE9.47%8.63%7.82%7.00%17.80%16.32%14.69%13.21%
R20.96290.96920.97470.97980.97830.98180.98520.9882
MAE0.41950.38250.34660.311146.502842.445938.412434.4107
RMSE0.63430.57840.52460.471682.438475.222468.099661.1102
STDAPE12.47%11.30%10.26%9.22%66.74%62.73%55.86%49.06%
DA72.00%72.79%74.58%75.87%72.59%73.09%74.08%75.07%
U10.05030.04590.04160.03740.06530.05960.0540.0485
U20.9520.93540.92690.91910.97730.97680.97580.9755
MAPE8.31%7.57%6.86%6.16%18.48%16.96%15.26%13.66%
R20.96480.97080.97610.98080.97790.98160.98490.9879
Table 5. Wind speed forecasting result for each site by different models in the second season.
Table 5. Wind speed forecasting result for each site by different models in the second season.
SiteDataModelMAERMSESTDAPEDAU1U2MAPER2
Site 1Wind speedMOGWO-CM50.35330.46479.07%78.55%0.04570.78018.97%0.9765
MODA-CM50.35310.46449.06%78.55%0.04560.77988.97%0.9765
MOMVO-CM50.35330.46499.08%78.45%0.04570.78028.97%0.9765
MODE-CM50.35330.46489.08%78.35%0.04570.77998.97%0.9765
Wind powerMOGWO-CM536.812458.383461.35%70.51%0.05100.587419.52%0.9883
MODA-CM536.808358.381961.15%70.51%0.05110.586119.51%0.9884
MOMVO-CM536.808758.382261.58%70.41%0.05120.586219.54%0.9883
MODE-CM536.828258.392961.58%70.31%0.05110.586819.54%0.9883
Site 2Wind speedMOGWO-CM50.37020.514711.57%77.36%0.05910.782510.41%0.9599
MODA-CM50.37010.514611.57%77.36%0.05910.782810.41%0.9599
MOMVO-CM50.37010.514711.58%77.16%0.05910.782610.41%0.9599
MODE-CM50.37020.514811.59%77.46%0.05920.782810.41%0.9599
Wind powerMOGWO-CM528.56653.286437.05%71.40%0.06630.794217.77%0.9833
MODA-CM528.552153.249636.91%71.40%0.06620.794517.75%0.9834
MOMVO-CM528.580953.347137.05%71.30%0.06630.794917.77%0.9833
MODE-CM528.579153.358537.09%71.40%0.06640.794417.77%0.9833
Site 3Wind speedMOGWO-CM50.57190.759513.09%74.68%0.0630.792111.93%0.9438
MODA-CM50.57190.759513.08%74.78%0.0630.792511.92%0.9438
MOMVO-CM50.5720.759813.10%74.88%0.0630.793111.93%0.9437
MODE-CM50.57230.760113.10%74.58%0.0630.792311.93%0.9437
Wind powerMOGWO-CM567.6516102.831555.22%70.11%0.06960.878821.67%0.9731
MODA-CM567.6172102.737154.96%70.31%0.06950.878121.65%0.9732
MOMVO-CM567.6892102.940355.08%70.11%0.06960.877621.68%0.9731
MODE-CM567.6626102.844255.40%70.21%0.06960.879321.68%0.9731
Site 4Wind speedMOGWO-CM50.41620.606310.78%60.68%0.05341.13428.88%0.9623
MODA-CM50.41620.606110.77%60.58%0.05341.1348.88%0.9624
MOMVO-CM50.41640.606410.77%60.68%0.05341.13388.89%0.9623
MODE-CM50.41630.606310.77%60.58%0.05341.13388.89%0.9623
Wind powerMOGWO-CM545.427581.993624.61%61.87%0.06561.082214.23%0.9777
MODA-CM545.403981.929724.57%61.77%0.06561.082414.23%0.9778
MOMVO-CM545.450382.103524.63%61.87%0.06571.082414.23%0.9777
MODE-CM545.469182.094524.64%61.77%0.06571.082414.24%0.9777
Table 6. The forecasting result of wind speed and wind power in the third season.
Table 6. The forecasting result of wind speed and wind power in the third season.
SiteModelWind Speed Forecasting Result
MAERMSESTDAPEDAU1U2MAPER2
Site 1MODA-CM20.36160.52429.37%77.66%0.05270.86088.50%0.9622
MODA-CM30.32990.47898.64%80.24%0.04820.85467.76%0.9685
MODA-CM40.29770.43017.79%81.73%0.04330.85647.01%0.9746
MODA-CM50.26760.38746.88%82.82%0.0390.85776.29%0.9795
Site 2MODA-CM20.32040.45728.92%78.85%0.05270.95458.45%0.9531
MODA-CM30.29260.4178.16%81.43%0.04810.947.72%0.961
MODA-CM40.26520.37917.48%83.52%0.04370.93127.00%0.9678
MODA-CM50.23650.33796.63%86.00%0.03890.92426.24%0.9745
Site 3MODA-CM20.45220.637912.53%77.56%0.05940.862710.26%0.9376
MODA-CM30.41150.579111.40%79.64%0.05390.85839.33%0.9488
MODA-CM40.37160.522810.32%81.13%0.04870.85618.43%0.9585
MODA-CM50.33360.47069.27%83.22%0.04380.85847.57%0.9665
Site 4MODA-CM20.38830.640513.61%66.93%0.04921.04777.92%0.9643
MODA-CM30.35470.585112.41%67.73%0.0451.02247.22%0.9702
MODA-CM40.32050.530111.18%69.02%0.04071.00126.53%0.9756
MODA-CM50.28710.475210.18%70.51%0.03650.99055.86%0.9804
SiteModelWind Power Forecasting Result
Site 1MODA-CM232.489553.787136.59%77.86%0.05250.842716.20%0.9874
MODA-CM329.688749.350233.95%79.25%0.04820.846214.83%0.9894
MODA-CM426.851144.480830.78%80.93%0.04350.85413.46%0.9913
MODA-CM524.105640.123827.06%82.72%0.03920.865412.00%0.9929
Site 2MODA-CM224.245844.495443.91%76.76%0.06521.014414.99%0.9802
MODA-CM322.159940.703640.24%79.15%0.05960.983413.72%0.9834
MODA-CM420.05536.984436.29%81.13%0.05420.962412.38%0.9863
MODA-CM517.960132.980131.82%82.82%0.04830.942111.10%0.9891
Site 3MODA-CM248.872276.0633129.59%77.56%0.07080.61223.75%0.9717
MODA-CM344.542869.2187115.33%79.44%0.06440.648221.44%0.9766
MODA-CM440.352362.7414101.21%81.53%0.05840.687619.35%0.9808
MODA-CM536.084856.070594.94%83.71%0.05220.703617.50%0.9847
Site 4MODA-CM250.366394.395651.41%66.34%0.06340.896517.36%0.9760
MODA-CM345.965586.172846.87%67.73%0.05790.90115.87%0.9788
MODA-CM441.659878.400743.20%68.82%0.05270.913114.42%0.9835
MODA-CM537.273869.808838.74%70.51%0.04690.917912.92%0.9865
Table 7. The forecasting results of each combined models for different sites in the fourth season.
Table 7. The forecasting results of each combined models for different sites in the fourth season.
SiteModelWind Speed Forecasting Result
MAERMSESTDAPEDAU1U2MAPER2
Site 1MODA-CM20.33360.48268.47%77.86%0.04950.888.04%0.9671
MODA-CM30.30380.43887.71%79.54%0.0450.87617.33%0.9729
MODA-CM40.27470.39696.98%82.42%0.04070.87366.63%0.9779
MODA-CM50.24550.35516.24%84.71%0.03640.87355.92%0.9823
Site 2MODA-CM20.33470.486611.57%77.66%0.05460.9359.16%0.9612
MODA-CM30.30530.44410.84%79.74%0.04980.9228.38%0.9677
MODA-CM40.27610.40019.56%82.03%0.04490.90777.57%0.9739
MODA-CM50.24680.35498.65%84.31%0.04030.90076.77%0.9789
Site 3MODA-CM20.37340.537311.29%79.34%0.06090.91299.84%0.9419
MODA-CM30.34080.490710.26%80.34%0.05560.90398.98%0.9517
MODA-CM40.30860.44419.27%81.83%0.05040.89558.13%0.9606
MODA-CM50.27620.39788.44%83.22%0.04510.89517.29%0.9685
Site 4MODA-CM20.45090.633612.34%77.66%0.06040.829310.65%0.9465
MODA-CM30.41070.577411.26%80.04%0.0550.82939.70%0.9558
MODA-CM40.37280.524510.22%81.73%0.050.83298.79%0.9636
MODA-CM50.33340.46939.09%83.61%0.04470.83657.86%0.9711
SiteModelWind Power Forecasting Result
Site 1MODA-CM230.450751.838943.30%76.86%0.05190.974616.11%0.9872
MODA-CM327.744147.071938.51%79.15%0.04720.969114.64%0.9895
MODA-CM425.121742.792335.70%80.64%0.04290.966413.30%0.9913
MODA-CM522.471238.238531.23%81.83%0.03830.964811.87%0.9931
Site 2MODA-CM222.063835.0959538.95%78.25%0.0440.518533.41%0.9922
MODA-CM320.114831.8624469.76%80.54%0.03990.570829.73%0.9935
MODA-CM418.178928.8482451.01%82.42%0.03620.585727.77%0.9947
MODA-CM516.263125.6994387.20%83.71%0.03220.636824.37%0.9958
Site 3MODA-CM235.749765.975442.12%74.48%0.09390.955519.27%0.9581
MODA-CM332.687360.097338.53%77.76%0.08560.958917.61%0.9653
MODA-CM429.662454.920334.63%79.54%0.07820.961515.93%0.9711
MODA-CM526.413648.608131.34%81.83%0.06920.966814.28%0.9774
Site 4MODA-CM250.033383.8969749.01%76.17%0.07720.717943.81%0.9702
MODA-CM345.682176.5383716.34%77.46%0.07040.687141.01%0.9752
MODA-CM441.155868.8028626.17%79.15%0.06330.62336.33%0.9789
MODA-CM536.994762.0098574.43%80.93%0.05710.601633.05%0.9838
Table 8. The results of the DM test and WRS test for wind speed and wind power by each combined model.
Table 8. The results of the DM test and WRS test for wind speed and wind power by each combined model.
PeriodSiteTestWind SpeedWind Power
MODA-CM2MODA-CM3MODA-CM4MODA-CM5MODA-CM2MODA-CM3MODA-CM4MODA-CM5
First
Season
Site 1DM Test15.5146 *14.8602 *12.4159 *-8.5860 *8.6528 *6.6730 *-
WRS Test0.696 (1)0.6998 (1)0.7299 (1)0.7249 (1)0.6832 (1)0.6988 (1)0.7095 (1)0.7203 (1)
Site 2DM Test13.1674 *11.5685 *10.0139 *-10.0496 *9.5269 *7.9863 *-
WRS Test0.8018 (1)0.7791 (1)0.7916 (1)0.7953 (1)0.8747 (1)0.8722 (1)0.8702 (1)0.8683 (1)
Site 3DM Test14.0858 *13.7322 *13.5449 *-8.2746 *8.6871 *9.6449 *-
WRS Test0.6724 (1)0.6695 (1)0.6935 (1)0.7111 (1)0.6688 (1)0.6717 (1)0.6756 (1)0.6817 (1)
Site 4DM Test12.2121 *12.2736 *10.0069 *-9.9215 *9.8950 *8.2886 *-
WRS Test0.9969 (1)0.9902 (1)0.974 (1)0.9904 (1)0.8323 (1)0.8339 (1)0.8323 (1)0.8319 (1)
Seconds
Season
Site 1DM Test18.1848 *16.5849 *15.0379 *-11.2257 *11.2726 *10.0365 *-
WRS Test0.8078 (1)0.8047 (1)0.8094 (1)0.8068 (1)0.8876 (1)0.878 (1)0.8706 (1)0.8809 (1)
Site 2DM Test13.8823 *15.1337 *12.4642 *-7.0921 *6.4662 *7.3349 *-
WRS Test0.7429 (1)0.7476 (1)0.759 (1)0.7608 (1)0.7528 (1)0.7378 (1)0.7216 (1)0.7325 (1)
Site 3DM Test17.6655 *16.8552 *14.6886 *-11.4051 *12.2241 *9.1368 *-
WRS Test0.5999 (1)0.6083 (1)0.613 (1)0.6217 (1)0.9473 (1)0.942 (1)0.9438 (1)0.9398 (1)
Site 4DM Test12.3603 *13.1989 *10.3810 *-8.9026 *8.8885 *8.0601 *-
WRS Test0.7159 (1)0.706 (1)0.6962 (1)0.7048 (1)0.831 (1)0.8319 (1)0.8204 (1)0.8272 (1)
Third
Season
Site 1DM Test10.2836 *10.0742 *8.1850 *-10.4397 *10.0330 *7.9945 *-
WRS Test0.8255 (1)0.8297 (1)0.8435 (1)0.8549 (1)0.9071 (1)0.8986 (1)0.8983 (1)0.8945 (1)
Site 2DM Test12.8066 *14.1692 *9.9269 *-5.8617 *5.7684 *7.1332 *-
WRS Test0.8133 (1)0.8018 (1)0.7958 (1)0.8084 (1)0.9597 (1)0.9449 (1)0.9404 (1)0.9437 (1)
Site 3DM Test14.2602 *14.8542 *12.3259 *-12.5005 *12.1284 *10.7953 *-
WRS Test0.6149 (1)0.6449 (1)0.647 (1)0.666 (1)0.6954 (1)0.7037 (1)0.707 (1)0.7183 (1)
Site 4DM Test9.8375 *9.9908 *9.1350 *-9.0853 *8.9344 *7.4904 *-
WRS Test0.9427 (1)0.966 (1)0.9616 (1)0.9752 (1)0.9713 (1)0.9665 (1)0.9513 (1)0.9471 (1)
Four
Season
Site 1DM Test14.1478 *13.6951 *12.2685 *-8.9806 *9.8720 *7.0639 *-
WRS Test0.9448 (1)0.914 (1)0.9242 (1)0.9258 (1)0.8697 (1)0.8692 (1)0.8744 (1)0.8766 (1)
Site 2DM Test11.9130 *11.7688 *9.2459 *-8.5189 *8.9237 *6.2209 *-
WRS Test0.8767 (1)0.8775 (1)0.8527 (1)0.8533 (1)0.905 (1)0.9045 (1)0.8972 (1)0.8991 (1)
Site 3DM Test13.8255 *13.7087 *11.2639 *-7.2845 *7.2283 *7.4506 *-
WRS Test0.4918 (1)0.5037 (1)0.5219 (1)0.5325 (1)0.7759 (1)0.7637 (1)0.7568 (1)0.7548 (1)
Site 4DM Test15.4436 *15.1429 *12.2494 *-10.0481 *10.1940 *9.0470 *-
WRS Test0.947 (1)0.9459 (1)0.9562 (1)0.9734 (1)0.9623 (1)0.9382 (1)0.9193 (1)0.9152 (1)
Note: * is the 5% significance level. (0) there is difference between two sample at the 5% significance level. (1) there is no dif-ference between two sample at the 5% significance level.
Table 9. R2 Values of Each Distribution Fitting.
Table 9. R2 Values of Each Distribution Fitting.
ModelDistributionR2 for Wind SpeedR2 for Wind Power
Site 1Site 2Site 3Site 4Site 1Site 2Site 3Site 4
MODA-CM2Normal0.88990.87700.87310.75680.65710.68000.60080.5499
Logistic0.95550.94870.94320.85490.81050.85220.74960.7045
Stable0.97600.97290.96710.94070.97950.99110.96980.9684
t-Location Scale0.98460.98250.97680.94660.97840.99260.96340.9573
MODA-CM3Normal0.89180.87520.87760.76700.65980.68130.60050.5572
Logistic0.95640.94680.94700.86500.81310.85290.74980.7132
Stable0.97610.97100.97040.94870.98310.99170.96940.9749
t-Location Scale0.98390.98120.97960.95440.98200.99330.96320.9642
MODA-CM4Normal0.89380.87660.87780.77550.66390.68220.60420.5599
Logistic0.95790.94900.94730.87390.81750.85630.75450.7160
Stable0.97720.97380.97080.95680.98480.99430.97260.9776
t-Location Scale0.98510.98390.97990.96240.98380.99530.96660.9682
MODA-CM5Normal0.89310.87660.87920.77670.66610.68500.60690.5648
Logistic0.95750.94860.94820.87520.82110.85820.75620.7225
Stable0.97710.97300.97120.95750.98760.99420.97310.9824
t-Location Scale0.98520.98300.98060.96290.98660.99520.96710.9725
Table 10. Evaluation Index Regulations for Uncertainty Forecasting Performance Comparison.
Table 10. Evaluation Index Regulations for Uncertainty Forecasting Performance Comparison.
MetricDefinitionEquation
Upper BoundUpper bounds of the wind speed forecasting value U ( i ) = F ( i ) + K 1 0.5 α × σ N
Lower BoundLower bounds of the wind speed forecasting value L ( i ) = F ( i ) K 1 0.5 α × σ N
FICPForecast interval coverage probability of testing dataset FICP = 1 N i = 1 N c i × 100 %
FINAWForecast interval normalized average width of testing dataset FINAW = 1 N R i = 1 N ( U i L i )
AWDiAccumulated width deviation of testing sample i A M D i = { ( L i A u i ) / ( U i L i ) , 0 , ( A u i U i ) / ( U i L i ) , A u i < L i A u i [ L i , U i ] A u i > U i
AWDAccumulated width deviation of testing dataset A M D = 1 N R i = 1 N A M D i
Note: F(i) is the corresponding point prediction value at point i. K and σ are the quantile and scale parameters of the logistic DF. N represents the forecasting length, and NR is the difference between the maximum and minimum forecasting values. If the actual value is A u i [ L i , U i ] ,   c i = 1 ; otherwise, ci = 0.
Table 11. Uncertainty forecasting Performance Comparison Table of the Proposed Model and Several Competitive Models from Experiment I-III.
Table 11. Uncertainty forecasting Performance Comparison Table of the Proposed Model and Several Competitive Models from Experiment I-III.
SiteAlphaMetricUncertainty Analysis for Wind PowerUncertainty Analysis for Wind Speed
MODA-CM2MODA-CM3MODA-CM4MODA-CM5MODA-CM2MODA-CM3MODA-CM4MODA-CM5
Site 15%FICP92.53%93.55%94.64%95.06%90.45%92.26%93.82%94.59%
FINAW0.18960.1750.16130.15440.23370.21880.2010.1921
AWD0.05970.04820.0380.03338.5996.67184.91764.0492
10%FICP84.40%86.28%88.05%88.79%82.71%85.14%87.67%88.99%
FINAW0.11550.10670.09840.09410.17340.16250.14920.1425
AWD0.18490.15570.12990.11720.569216.726212.961311.1074
15%FICP77.58%79.71%81.97%82.96%76.17%78.97%82.02%83.53%
FINAW0.08520.07870.07260.06940.1420.13310.12220.1168
AWD0.33380.28640.24330.221934.082328.373822.598119.7348
Site 25%FICP92.68%93.85%94.64%95.14%90.72%92.36%94.07%94.82%
FINAW0.13350.12280.11260.10730.20250.18710.17130.1634
AWD0.07350.06060.04940.04378.01956.19514.5783.8903
10%FICP84.97%86.61%88.19%88.91%83.11%85.39%87.50%89.01%
FINAW0.08280.07620.06990.06660.14870.13740.12580.12
AWD0.20250.17180.14390.129619.291815.535712.140610.5021
15%FICP78.08%80.26%82.54%83.71%75.92%79.32%82.39%83.73%
FINAW0.06160.05680.0520.04960.12120.11190.10250.0978
AWD0.35130.30210.25660.233632.071826.335821.11518.5686
Site 35%FICP93.15%94.05%95.14%95.68%90.20%92.24%93.70%94.44%
FINAW0.26190.24550.22770.21890.23850.21910.19950.1903
AWD0.04910.03990.03120.0277.34835.48473.97853.3008
10%FICP84.13%86.19%87.95%88.59%83.04%85.32%87.75%88.74%
FINAW0.15250.14290.13270.12760.17670.16240.14810.1411
AWD0.17410.14880.12370.111817.718714.036810.85559.3221
15%FICP76.74%78.77%80.51%81.65%76.26%79.19%82.07%83.73%
FINAW0.10980.10290.09560.09190.14470.13290.12130.1156
AWD0.33340.29070.24740.226229.408523.879818.958516.5917
Site 45%FICP93.63%94.44%95.31%95.76%91.52%93.15%94.47%95.16%
FINAW0.24220.22450.20990.20230.23750.22140.20480.1956
AWD0.04260.03420.02670.02336.89285.37314.02683.3892
10%FICP84.80%85.94%87.57%88.42%82.66%84.67%87.10%88.19%
FINAW0.13370.12380.11580.11170.16210.15120.13980.1336
AWD0.16910.14380.12080.109819.907116.296812.965211.3022
15%FICP77.18%79.37%81.45%82.49%76.22%78.70%81.00%82.24%
FINAW0.09350.08660.0810.07810.12710.11850.10960.1047
AWD0.33110.28730.24870.229834.92429.445724.146621.4982
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, Q.; Lv, Q.; Zhang, G. A Combined Forecasting System Based on Modified Multi-Objective Optimization for Short-Term Wind Speed and Wind Power Forecasting. Appl. Sci. 2021, 11, 9383. https://doi.org/10.3390/app11209383

AMA Style

Zhou Q, Lv Q, Zhang G. A Combined Forecasting System Based on Modified Multi-Objective Optimization for Short-Term Wind Speed and Wind Power Forecasting. Applied Sciences. 2021; 11(20):9383. https://doi.org/10.3390/app11209383

Chicago/Turabian Style

Zhou, Qingguo, Qingquan Lv, and Gaofeng Zhang. 2021. "A Combined Forecasting System Based on Modified Multi-Objective Optimization for Short-Term Wind Speed and Wind Power Forecasting" Applied Sciences 11, no. 20: 9383. https://doi.org/10.3390/app11209383

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop