Next Article in Journal
A Class of Fifth-Order Chebyshev–Halley-Type Iterative Methods and Its Stability Analysis
Next Article in Special Issue
Smooth and Efficient Path Planning for Car-like Mobile Robot Using Improved Ant Colony Optimization in Narrow and Large-Size Scenes
Previous Article in Journal
Existence and Uniqueness of Some Unconventional Fractional Sturm–Liouville Equations
Previous Article in Special Issue
Fuzzy Fractional Order PID Tuned via PSO for a Pneumatic Actuator with Ball Beam (PABB) System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid LSTM-Based Fractional-Order Neural Network for Jeju Island’s Wind Farm Power Forecasting

1
School of Electrical Engineering, Vellore Institute of Technology, Vellore 632014, India
2
Department of Electrical and Electronics Engineering, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Malaysia
*
Author to whom correspondence should be addressed.
Fractal Fract. 2024, 8(3), 149; https://doi.org/10.3390/fractalfract8030149
Submission received: 1 February 2024 / Revised: 17 February 2024 / Accepted: 20 February 2024 / Published: 5 March 2024
(This article belongs to the Special Issue Applications of Fractional-Order Calculus in Robotics)

Abstract

:
Efficient integration of wind energy requires accurate wind power forecasting. This prediction is critical in optimising grid operation, energy trading, and effectively harnessing renewable resources. However, the wind’s complex and variable nature poses considerable challenges to achieving accurate forecasts. In this context, the accuracy of wind parameter forecasts, including wind speed and direction, is essential to enhancing the precision of wind power predictions. The presence of missing data in these parameters further complicates the forecasting process. These missing values could result from sensor malfunctions, communication issues, or other technical constraints. Addressing this issue is essential to ensuring the reliability of wind power predictions and the stability of the power grid. This paper proposes a long short-term memory (LSTM) model to forecast missing wind speed and direction data to tackle these issues. A fractional-order neural network (FONN) with a fractional arctan activation function is also developed to enhance generated wind power prediction. The predictive efficacy of the FONN model is demonstrated through two comprehensive case studies. In the first case, wind direction and forecast wind speed data are used, while in the second case, wind speed and forecast wind direction data are used for predicting power. The proposed hybrid neural network model improves wind power forecasting accuracy and addresses data gaps. The model’s performance is measured using mean errors and R2 values.

1. Introduction

Optimising the integration of wind energy into the power grid and ensuring grid stability relies heavily on accurate wind power prediction. Over the years, researchers have explored various techniques, including neural networks, machine learning, and deep learning methods, to enhance the precision and reliability of wind power predictions. Machine learning creates a generalised model from previous input data and output results, then predicts outcomes in the future using multiple learning methods. In machine learning approaches, artificial neural networks (ANNs) [1] and support vector machines (SVMs) [2] are commonly used. ANNs can predict non-linear data and analyse the correlation between impact data and wind power. Training ANNs requires a lot of data and time, while high-dimensional data limits computational speed, leading to local optimum solutions. SVM avoids these issues and generalises them effectively. In [3], the integration of least squares and SVM (LS-SVM) was used to estimate the wind power load, enhancing computation efficiency and predicting accuracy. Incorporating LS-SVM principles, Zhang et al. introduced modifications to the model that effectively minimised prediction errors [4]. In [5], the researchers developed a fuzzy neural network for wind power forecasting coupled with online risk assessment and, in addition, investigated the effectiveness and potential improvements in enhancing wind energy forecasting models. Jie Shi et al. combined the Hilbert–Huang transform with artificial intelligence (AI) for power forecasting and further explored the effectiveness of this integrated model in improving prediction accuracy and enhancing renewable energy integration [6]. In [7], the authors developed the application of radial basis function neural networks in wind power forecasting, incorporating probabilistic methods to enhance forecasting accuracy and uncertainty assessment. The researchers employed the empirical mode decomposition (EMD) model with a neural network to forecast wind power and speed [8]. Further, they investigated the effectiveness of EMD-based models in improving short-term wind forecasting accuracy. The authors in [9] created an emotional neural network technique for predicting weather patterns and wind power generation. Additionally, they emphasised that this method can be applied to real-world scenarios. The authors proposed Gaussian processes integrated with numerical weather prediction (NWP) and complex-valued ANN for day-ahead wind power forecasting and examined their effectiveness in optimising wind energy generation and improving prediction accuracy [10,11].
Jyotirmayee et al. presented a variational mode decomposition technique in combination with a multi-kernel regularised pseudo-inverse ANN for wind power forecasting [12]. The authors developed 3D convolutional neural networks (CNNs) for extracting numerical weather prediction data in wind power forecasting, investigated similar methods to enhance prediction accuracy, and considered the potential benefits of utilising 3D CNNs in this context [13]. In [14], the structured neural network model for predicting short-term wind power emphasises the developed model’s effectiveness and potential in achieving accurate short-term predictions. The researchers in [15,16] implemented an ANN for predicting wind power’s discrete wavelet-transform-based wind speed and highlighted the network’s effectiveness and potential in improving renewable energy integration. Additionally, the researchers emphasised the need for further investigation into model enhancements to address uncertainties and improve forecasting precision. In addition, the authors of [17] explored the current machine learning techniques for power forecasting, identifying emerging trends, and highlighted the key challenges faced in this domain. AI has shown promise in enhancing wind power generation forecasting through hybrid approaches, but significant challenges still need to be addressed for practical implementation and improved accuracy [18]. Despite these obstacles, the prospects for AI-based forecasting in the renewable energy sector remain encouraging. The authors of [19] employed an ANN model to forecast the wind power generation of the Pawan Danawi wind farm in Sri Lanka and highlighted that the model could also be applied to the environmental and climatic conditions to identify the wind power potential of the area. Machine learning methods are more effective than statistical approaches in predicting non-linear wind power data due to their adaptability and self-learning capabilities. However, these models have limitations in expressing complex data. This is because of the advancement of big data technology. Deep learning algorithms can overcome these challenges by extracting higher-level abstract features from the original samples. This enables the discovery of complex rules in high-dimensional data.
Deep learning models have advanced significantly in recent years, and deep neural network (DNN) algorithms have been introduced [20,21,22]. Recurrent neural networks (RNNs) and LSTM networks are robust architectures for sequence data, demonstrating advantages in non-linear feature learning [23]. Thus, RNN and LSTM are the most often used deep learning models in wind power prediction research. The authors of [24] conducted wind power generation prediction using multivariate LSTM time series. The researchers of [21] implemented deep feature extraction and LSTM techniques for data-driven wind speed forecasting and explored the effectiveness of these techniques in improving wind speed predictions. In [25], the authors used data cleaning and feature extraction techniques for power prediction. In [26], the authors used machine learning algorithms such as light gradient boosting machines (GBMs) and LSTM networks for short-term wind forecasting of weather stations in India and also aimed to enhance wind energy prediction accuracy, contributing to efficient renewable energy integration and management. The authors of [27] implemented an ensemble approach combining algorithms, namely, deep learning and gradient descent, for wind power forecasting and explored the model’s effectiveness in improving forecasting accuracy and reliability. The deep-learning-based methods in [28] were developed to generate accurate and reliable prediction intervals for wind power forecasting, addressing the multi-objective nature of the problem. The researchers in [29] employed a Seq2Seq wind power output prediction method developed using deep learning and a clustering algorithm to forecast wind power with NWP data and real-time historical wind data. Adam Kisvari et al. [30] applied a deep learning approach using a data-driven and a gated recurrent unit (GRU) to forecast wind power. Further, the researchers in [31] implemented a temporal convolution network (TCN)-based approach for day-ahead wind power forecasting and compared the implemented method with the LSTM and GRU models. In [32], the authors developed LSTM-based RNNs for wind power forecasting, focusing on variable selection techniques. The authors in [33] implemented the GRU neural network method for wind power prediction, utilising evolutionary network architecture search for optimisation. The researchers in [34,35] constructed multi-modal spatio-temporal neural networks and optimised deep autoregressive RNNs for multi-horizontal wind power forecasting.
The use of attention-based models has become more popular for predicting long-term series. In [36,37], the authors demonstrated self-attention’s effectiveness in capturing complex patterns and dynamics, particularly in capturing long-distance dependencies within time-series data. Juan Ren et al. [38] developed the CNN-LSTM-LightGBM framework with an attention mechanism for short-term wind power forecasting, which aimed to enhance forecasting accuracy by efficiently capturing temporal dependencies and extracting relevant features from wind power data. In [39], the authors proposed wind power forecasting methods using variational mode decomposition, and LSTM attention networks showed the encoder–decoder structure’s superiority over a dual attention–LSTM neural network in enhancing prediction performance. Lei Wang et al. constructed an advanced transformer model for ultra-short-term wind power prediction [40]. Nevertheless, challenges such as space–time complexity and input and output sequence limitations remain. Furthermore, Ref. [41] developed a novel method for ultra-short-term wind power prediction, addressing previous limitations through feature extractions. The approach shows promising results, improving prediction accuracy and addressing space–time complexity issues.
A fractional-order activation function is a specific activation function used in artificial neural networks. It allows the use of non-integer exponents to calculate the output of a neuron, which can improve the performance of specific neural networks. These activation functions possess unique properties that make them suitable for specific modelling tasks and data. They are beneficial for capturing long-range dependencies and non-linear relationships in data, which cannot be effectively handled by conventional activation functions such as arctan. The fractal nature of these activation functions is attributed to their self-similarity property, which enables them to capture complex patterns in data at multiple scales. Hence, they benefit time-series forecasting applications where the data may exhibit fractal-like behaviour [42]. Using fractional-order activation functions can enhance the performance of FONN model-based forecasting by allowing for more accurate and efficient modelling of complex non-linear relationships in the data. Additionally, these functions can help to reduce overfitting, a common problem in traditional neural networks. The use of fractional-order activation functions may increase the computational load during training. However, this can be offset by the improved performance and accuracy of the model, leading to faster and more efficient forecasting [43]. Therefore, it is justified to use fractional-order activation functions because of their ability to capture complex non-linear relationships in data, which can improve the efficiency and accuracy of FONN model-based forecasting.
Motivated by the above literature, this paper presents a new hybrid model that uses LSTM to forecast missing input data and FONN to predict generated wind power. The performance of the proposed approach is evaluated based on the wind data collected from Jeju Island’s wind farm in three different island sites. The key contributions of this research are outlined below:
  • The LSTM model is designed to predict missing input parameters, including wind speed and direction. Its performance is evaluated through root mean squared error (RMSE) assessment.
  • The FONN model predicts wind power using the LSTM’s forecast data and evaluates performance with a coefficient of determination (R2) and mean squared error (MSE).
  • The models developed were evaluated in two case studies involving missing data scenarios for specific parameters.
The subsequent sections of the manuscript are organised to explore the research comprehensively. Section 2 describes the dataset from Jeju Island in three different sites and presents the data visualisation and correlation analysis in various scenarios. Section 3 describes the proposed hybrid LSTM-based fractional-order neural network model for wind power forecasting. Section 4 shows the results and discussion of the proposed models’ performance evaluation to handle the missing parameter data. Section 5 concludes the proposed work.

2. Dataset Description

In South Korea, Jeju Island has a prosperous wind energy landscape with advanced wind farms strategically placed throughout scenic terrain. These wind farms take advantage of the island’s plentiful wind resources, significantly contributing to its renewable energy portfolio. Sites A, B, and C are among the top wind farms on Jeju Island, each with unique specifications and characteristics. Table 1 provides an overview of the data collection period, collection time interval, and detailed wind turbine specifications for Sites A, B, and C [44].
The wind turbine specifications presented in the table highlight each site’s customised design and engineering considerations. These specifications, which include the model, output, wind speed capacity, rotor dynamics, voltage, and power control, showcase Jeju Island’s commitment to harnessing wind energy efficiently and sustainably. The island’s wind farms are characterised by their meticulous data collection and cutting-edge turbine specifications, which testify to their dedication to renewable energy and their aspiration to create a cleaner and greener future.
As shown in Figure 1, the data from sites A, B, and C include wind power, direction, and speed, indicating the chaotic behaviour. There were 1080, 432, and 720 samples from sites A, B, and C, respectively. These samples’ pair plots are shown in Figure 2 and Figure 3. Figure 1 shows the pairwise relationships in a dataset, while Figure 3 shows correlation coefficients between wind direction, speed, and power at the three sites. As shown in the figure, the diagonal elements are one, indicating a perfect correlation with each variable. The off-diagonal elements in the figure show the correlation between the two parameters.
Utilising the Pearson correlation coefficient, a numerical measure ranging from −1 to 1, the correlation analysis shown in Figure 3 can quantify the strength and direction of relationships. This method is invaluable in providing essential insights into how changes in one parameter might correspond to changes in another. Upon examination of the wind farms at each site, it was discovered that a positive linear relationship exists between wind direction, speed, and power. The correlation coefficient between wind direction and speed at Site A is 0.25, at Site B it is 0.22, and at Site C it is 0.36, indicating a weak correlation. Similarly, the correlation coefficient between wind direction and power at Site A is 0.34, at Site B it is 0.17, and at Site C it is 0.38, indicating a weak correlation. Lastly, the correlation coefficient between wind speed and wind power at Site A is 0.82, at Site B it is 0.95, and at Site C it is 0.98, indicating a robust correlation. It is important to note that all correlation coefficients fall within the range of −1 to 1, demonstrating that the relationships are positive and linear. Additionally, it should be noted that the correlations observed at Site C are stronger than those at Site A, and the correlations at Site A are the most robust of the three sites.

2.1. Correlation Analysis of Wind Speed Parameter with Missing Data

The first case examines the correlation between wind direction, speed, and power across Sites A, B, and C, where the wind speed parameter has missing data. The correlation matrix for Site A, shown in Figure 4, presents a comprehensive view of these relationships. Notable correlations emerge, with wind power and direction exhibiting a moderate positive correlation of 0.31, indicating a tendency for increased wind direction to coincide with heightened wind power. Additionally, wind power and speed show a stronger positive correlation of 0.63, suggesting that higher wind power values correspond to elevated wind speed values. Conversely, wind direction and speed demonstrate a more modest positive correlation of 0.22. The correlation matrix for Site B mirrors these trends, showing modest positive correlations of 0.13, 0.21, and 0.23 between wind power and direction, wind power and speed, and wind direction and speed, respectively. At Site C, wind power and wind direction exhibit a moderately positive correlation of 0.31. In contrast, the correlation between wind power and speed is characterised by a relatively moderate positive correlation of 0.43. Similarly, wind direction and speed display a moderately positive correlation of 0.34. Together, these correlation matrices provide a comprehensive understanding of the relationships between wind power, direction, and speed, shedding light on the nature and strength of these associations and emphasising the need for thorough analysis, particularly in instances of missing data.

2.2. Correlation Analysis of Wind Direction Parameter with Missing Data

The second case examines the correlation between wind direction, speed, and power across Sites A, B, and C, where the wind direction parameter has missing data. As shown in Figure 5, the correlation matrices for these sites reveal insightful connections. At Site A, wind power displays a moderate positive correlation with wind direction of 0.22 and a strong positive correlation with wind speed of 0.82. Additionally, a moderately positive correlation of 0.3 between wind direction and wind speed becomes evident. Site B’s matrix unveils distinctive relationships. Wind power and wind direction exhibit a moderate negative correlation of −1.19, while wind power and wind speed share a robust positive correlation of 0.92. In contrast, wind direction and speed show a weak negative correlation of −1.084. Similarly, Site C’s matrix reveals meaningful correlations. Wind power positively correlates with wind direction, with a correlation value of 0.19, and strongly correlates with wind speed, with a correlation value of 0.93. Further, wind direction and speed demonstrate a weak positive correlation of 0.081. In both scenarios, the correlation coefficients provide essential insights into the strength and nature of these relationships. Such understanding holds substantial implications for informed decision making and predictive analyses, particularly within renewable energy and meteorology.

3. Proposed Methodology

The proposed methodology outlines a step-by-step framework for predicting missing wind parameters and power generation within the Jeju Island wind farm context across different sites, as shown in Figure 6. This approach incorporates well-defined techniques to offer unique insights into predictive modelling and performance assessment. The initial step involves data collection and pre-processing. Specifically, the wind power, speed, and direction datasets from the Jeju Island wind farm at Sites A, B, and C are collected. This dataset provides the implementation for subsequent analysis and predictive modelling. Next, the correlation analysis is conducted with missing input data. This step considers two scenarios: missing wind speed data and missing wind direction data. The correlation analysis examines the complex relationships between wind speed, direction, and power under both scenarios. This analysis provides insights into how these parameters influence wind power generation. The detailed outcomes of this analysis are depicted in Figure 4 and Figure 5.
In the third step of the process, an LSTM model is used to forecast the missing wind speed and direction data. The process begins with data normalisation within the range of −1 to +1, followed by the compilation of time-series data. The dataset is then partitioned into distinct subsets for training and testing. This step involves building and training the LSTM model, progressively improving its predictive capabilities by learning from the data. Continuous evaluation ensures that the model reaches the desired level of accuracy within the allocated number of training iterations. If this level is not achieved, adjustments are made through loss calculations and model updates. Once the desired level of proficiency is attained, the model is used to forecast missing wind speed and direction data. These predicted data are then compared against actual data for comprehensive analysis. The model’s accuracy is evaluated through performance metrics such as the RMSE and visualisation of original and forecast waveforms.
The final stages of the methodology focus on predicting wind power using the FONN model. The process begins with developing a fractional-order arctan activation function using fractional derivatives. This newly developed function enhances the predictive capabilities of the FONN model. Subsequently, the FONN model is designed by integrating the developed fractional-order arctan function, rendering it adept at accurate wind power prediction. A precise parametrization of the model follows, encompassing the determination of the number of hidden layers and neurons. The FONN model is then iteratively trained to enhance its predictive performance. If necessary, model parameters are adjusted and retrained, ensuring continuous optimisation. A main evaluation criterion is whether the model achieves improved accuracy. Finally, the trained model is tested using a test dataset, and its performance is evaluated in terms of the coefficient of determination (R2) and MSE, with comparisons made against a conventional neural network model. This comprehensive methodology offers a structured approach to predicting wind power and leveraging predictive modelling techniques for renewable energy applications. As shown in Figure 6 and as explained earlier, the first part of the methodology is the LSTM model’s development for forecasting missing input data of wind speed and direction. The next part presents the FONN model that is used to make predictions of the generated wind power.

3.1. LSTM Model

The LSTM model’s architecture for forecasting 20% of missing input data of wind speed and direction is demonstrated in Figure 7. The input to the LSTM consists of three time-steps x t 1 ,   x t , and x t + 1 . The architecture includes memory blocks comprising each memory cell, input, forget, and output dates, which will be explained below.
In the forget gate, f t is computed using the sigmoid function; σ ( · ) determines the past information to forget:
f t = σ ( w f · h t 1 , x t + b f ) ,
In the above equation, h t 1 , w f , and b f represent the previous cell’s output, and the gate’s weight and bias, respectively. The sigmoid function σ ( · ) output varies between 0 and 1, representing complete forgetting at 0 and full retention at 1.
As for the input gate i t , it is also calculated based on the information to be stored in the cell state:
i t = σ ( w i · h t 1 , x t + b i ) .
This gate’s “tanh” layer adds weight to the cell state. The update equation for the memory cell, represented as C ˜ t , is
C ˜ t = t a n h ( w C · h t 1 , x t + b C ) ,
Here, w C and b C denote the memory cell’s weight and bias.
Using the output gate O t to determine the output information from the current cell, it can be calculated as
O t = σ ( w O · h t 1 , x t + b O ) .
The current cell’s output ( h t ) will be calculated as
h t = O t × t a n h ( C t ) .
where C t is the cell state and O t is the output gate.
Two critical factors that influence the performance of an LSTM model are input delays and the number of hidden units. An LSTM network with 10 hidden units was trained to achieve adequate performance using a trial-and-error approach. Incomplete learning, limited generalisation, and underfitting can occur if the LSTM model is not sufficiently trained. Incomplete learning can cause suboptimal performance because the model fails to capture all underlying patterns. Limited generalisation means that the model needs to extend its predictions beyond the training data, leading to poor performance on missing data. On the other hand, underfitting causes poor performance on both training and test datasets. However, overtraining is unlikely due to the need for more learning from the data. Therefore, adequate training is crucial for accurate predictions. To achieve this, the Adam solver was introduced with variable learning and dropout rates of 0.005 and 0.2, respectively, over 1000 epochs.

3.2. FONN Model

The FONN model’s architecture, designed to predict generated wind power using forecast missing wind direction and speed data with an LSTM model, is illustrated in Figure 8 [42]. In this configuration, there are 2 input nodes, 30 hidden nodes, and 1 output node, and their ratios are 2:30:1. The number of nodes in a hidden layer plays a significant role in determining the predictive capabilities of a neural network. Too few nodes may result in underfitting, while too many can lead to overfitting. In this case, a trial-and-error approach was used to determine that a hidden layer with 30 nodes would achieve satisfactory results. Within the architecture, bias values at the hidden and output layers were represented as “b”, with values [30, 1]. For the output layer, the activation function selected was “Purelin”. In contrast, the hidden layers’ activation function “F” employed can vary between developed and standard tangential functions. This variation was assessed to determine the model’s performance. For the training algorithm, the Levenberg–Marquardt algorithm was selected. This neural network training algorithm helps to fine-tune the model’s parameters effectively. The model’s performance evaluation was conducted using the performance measures outlined in the subsequent section, ensuring a comprehensive assessment of its predictive capabilities.

3.3. Fractional-Order Tangential Activation Functions

The tansig activation function, also known as the hyperbolic tangent sigmoid, is commonly used in hidden layers for classification tasks [45]. This function maps input values from the range of ( , + ) to ( 1 , 1 ) . Its mathematical expression is given below [42]:
f ( x ) = 2 1 + e 2 x 1 .
The tansig function is known to have a higher derivative compared to the sigmoid function. Additionally, its output mean is 0 when the average input values approach 0. These properties make the tansig function a valuable tool for training neural networks, as it can significantly improve convergence rates and expedite the training process. However, similar to the sigmoid function, tansig is also prone to the vanishing gradient problem [46]. Incorporating fractional-order derivatives into the tansig function to introduce a non-linear component can address the gradient problem. The fractional ordering of tansig can be derived by expressing Equation (6) using the MacLaurin series expansion as follows:
f ( x ) = n = 0 4 n ( 4 n 1 ) B 2 n ( 2 n ) ! x 2 n 1 .
The fractional ordering of the tansig activation function can be computed for an order α ( 0 , 0.9 ) as follows [42]:
D α f ( x ) = g ( x ) = D α n = 0 4 n ( 4 n 1 ) B 2 n ( 2 n ) ! x 2 n 1 , g ( x ) = n = 0 4 n ( 4 n 1 ) B 2 n ( 2 n 1 ) ! ( 2 n ) ! Γ ( 2 n α ) x 2 n 1 α .
Figure 9a shows the response of the fractional-order derivative of the tansig activation function for various values of α , as compared to the behaviour of the regular tansig function. The conventional tansig function has an S-shaped curve, similar to the sigmoid function and its variations. On the other hand, the fractional-order derivative of tansig has an S-shaped curve for lower values of α . However, for higher values of α , the function becomes non-linear due to the fractional ordering, which can help solve the vanishing gradient problem.
The hard tansig function is a commonly used version of the tansig activation function in deep learning applications. Unlike the tansig function, the hard tansig function is more efficient and computationally cheaper. It has a range of [ 1 , 1 ] and is defined as follows [45,47]:
f ( x ) = 1 if x < 1 , x if 1 x 1 , 1 if x > 1 .
The fractional-order derivative of the hard tansig function can be computed for an order α ( 0 , 0.9 ) as follows [42]:
D α f ( x ) = g ( x ) = D α 1 if x < 1 x if 1 x 1 1 if x > 1 , g ( x ) = 1 Γ ( 1 α ) x α if x < 1 1 Γ ( 2 α ) x 1 α if 1 x 1 1 Γ ( 1 α ) x α if x > 1 .
The comparison shown in Figure 9b highlights the response of the fractional-order derivative of the hard tansig activation function for different values of α orders compared to the conventional derivative. The analysis reveals that the functions take the α value derivative in specific intervals while exhibiting zero gradients in others. This aspect indicates that the vanishing gradient problem is less likely to occur in the fractional-order derivative of the hard tansig function as long as most of these units operate within the periods when the gradient is 1. Moreover, the analysis suggests that the fractional ordering has introduced non-linearity into the function, which will help resolve the vanishing gradient problem.
The LiSHT (linearly scaled hyperbolic tangent) is a popular activation function used in deep learning to address the “dead ReLU” issue. When the ReLU function is given negative input, it can become inactive, resulting in a zero gradient that prevents weight updates during backpropagation. As a result, to solve this problem, the LiSHT function multiplies the input with the element-wise hyperbolic tangent output. Additionally, since the hyperbolic tangent function has a range of [−1, 1], negative gradients are not eliminated like with ReLU functions, which helps maintain the optimal learning for training deep neural networks. The LiSHT function can be computed by multiplying the tansig function with its input, as shown in [47].
f ( x ) = x · δ ( x ) ,
The following expression defines the tansig function δ ( x ) , which can be found in Equation (6):
δ ( x ) = 2 1 + e 2 x 1 .
The LiSHT function in Equation (11) can be expressed using a MacLaurin series expansion as follows:
f ( x ) = n = 0 4 n ( 4 n 1 ) B 2 n ( 2 n ) ! x 2 n .
The equation above enables computation of the fractional ordering of the LiSHT activation function for an order α ( 0 , 0.9 ) as follows [42]:
D α f ( x ) = g ( x ) = D α n = 0 4 n ( 4 n 1 ) B 2 n ( 2 n ) ! x 2 n , g ( x ) = n = 0 4 n ( 4 n 1 ) B 2 n Γ ( 2 n + 1 ) ( 2 n ) ! Γ ( 2 n α + 1 ) x 2 n α .
The response of the fractional-order derivative of the LiSHT activation function for various α orders, compared with the conventional one, is shown in Figure 9c. The response indicates that the conventional LiSHT produces a positive output response. For lower α values, the fractional ordering of LiSHT achieves similar behaviour. However, the response shows fractional ordering introduces more significant non-linearity than other activation functions for higher α values.
Additionally, the conventional arctan activation function is employed at the hidden layer in neural networks. However, its non-monotonic nature can pose optimisation challenges. Mathematically, the arctan function is expressed as [48]
f ( x ) = tan 1 ( x ) .
This equation can be expanded using the MacLaurin series as follows:
f ( x ) = n = 0 ( 1 ) n 2 n + 1 x 2 n + 1 .
The arctan function is enhanced by introducing fractional-order derivatives to tackle these challenges, improving its smoothness and optimisation potential within the FONN model. The fractional-order derivative of the arctan activation function for an order α ( 0 , 0.9 ) can be computed as [49]
D α f ( x ) = g ( x ) = D α n = 0 ( 1 ) n 2 n + 1 x 2 n + 1 , g ( x ) = n = 0 ( 1 ) n Γ ( 2 n + 3 ) ( 2 n + 1 ) Γ ( 2 n + 2 α ) x 2 n + 1 α .
This enhancement results in smoother derivatives for the fractional-order arctan function, facilitating more effective gradient-based optimisation, which makes it better at capturing complex dynamics and long-range dependencies in wind power data, and its response at different α values is shown in Figure 9d. Compared to conventional functions, fractal activation functions like fractional-order arctan provide more flexibility in modelling non-linear systems. They are better at adapting to intricate patterns in renewable energy data.
Furthermore, the Purelin activation function is employed at the networks’ output layer. It is a linear function that directly relates output to input, giving a response of k x for an input of x. The response of Purelin is shown in Figure 10. For k = 1 , it functions as an identity. This function, with a hyperparameter k, is described as [50]
f ( x ) = k x .

3.4. Performance Metrics

The MSE and RMSE are widely recognised performance metrics that assess the difference between predicted and actual values. Extensive studies have demonstrated their effectiveness as error measures in numerical prediction tasks [51,52]. The MSE is computed between actual ( Y i ) and predicted ( Y i ^ ) as follows:
MSE = 1 N i = 1 N ( Y i Y i ^ ) 2 .
The RMSE provides an interpretable measure of the average forecasting error, which is computed as follows:
RMSE = MSE .
Additionally, the coefficient of determination, denoted as R2, is frequently used to show the predictive capability of forecasting methods in fitting actual data ( Y i ), calculated as [52]
R 2 = 1 i = 1 N ( Y i Y i ^ ) 2 i = 1 N ( Y i Y i ¯ ) 2 .
where Y ¯ signifies the average of the predicted values. R2 yields values ranging from 0 (indicating a poor match) to 1 (representing a perfect fit).
In all the above equations, ‘N’ represents the sample size and Y ^ i denotes predicted values. These metrics provide valuable insights into numerical forecasting approaches’ accuracy and predictive performance.

4. Results and Discussion

This section evaluates the LSTM model’s accuracy in forecasting missing wind speed and direction data and the FONN model’s performance in predicting wind power using the forecast data of missing wind speed and direction data in the Jeju Island wind farm for all sites under different cases.

4.1. Performance of LSTM Model

In Jeju Island’s wind farm, 20% of wind speed and direction data are missing at three sites (A, B, and C). This missing data can negatively impact the accuracy of power predictions, operational planning, efficiency, safety, and overall system reliability. An LSTM model has been developed to forecast missing wind speed and direction data to address the issue, as mentioned in Section 3 and compared with non-linear autoregressive (NAR) [53], and autoregressive integrated moving average (ARIMA) [54] models. The LSTM model has one input, 200 hidden units, and one output. The model uses learning and dropout rates of 0.005 and 0.2, respectively, for 1000 iterations. Table 2 displays RMSE values for various models used in forecasting missing wind speed and direction data at three sites. The table compares the performance of three different forecasting models: LSTM, NAR, and ARIMA. Based on Table 2, the performance analysis of various forecasting models is as follows:
  • The LSTM model exhibits the lowest RMSE values compared to the NAR and ARIMA models for forecasting missing wind speed data across all sites.
  • At Site A, the LSTM model achieved the lowest RMSE value of 0.16, followed by NAR with an RMSE of 0.353 and ARIMA with an RMSE of 0.583.
  • Similarly, the LSTM model at Site B outperformed the other models with an RMSE of 0.185, while the NAR and ARIMA models showed higher RMSE values of 0.297 and 0.458, respectively.
  • Finally, at Site C, the LSTM model exhibited the lowest RMSE of 0.112, followed by ARIMA with an RMSE of 0.387 and NAR with the highest RMSE of 0.457.
  • The following analysis is related to missing wind direction data forecasting, where the performance of the models varies across different sites.
  • At Site A, the LSTM model had the lowest RMSE of 0.18, followed by ARIMA with an RMSE of 0.386 and NAR with the highest RMSE of 0.442.
  • Similarly, at Site B, the LSTM model performed best with an RMSE of 0.425, followed by NAR with an RMSE of 0.185, and ARIMA with the highest RMSE of 0.572.
  • Finally, at Site C, the NAR model had the lowest RMSE of 0.395, followed by LSTM with an RMSE of 0.126, and ARIMA with the highest RMSE of 0.454.
The results show that the LSTM model performs better than the NAR and ARIMA models in forecasting wind speed and wind direction missing data across different sites. However, the performance may vary depending on the specific site and the nature of the wind data. Figure 11 and Figure 12 depict the actual wind speed and direction and those forecast by the LSTM model. Further, the numerical values in Table 2 indicate the model’s best performance, with RMSEs of around 0.11 and 0.12, respectively.

4.2. Performance of FONN Model

This section presents the FONN model’s performance in predicting wind power using the forecast missing wind speed and direction data with the LSTM model at different sites. As per the previous section, the LSTM model showed the best performance among the forecasting models. There are two case studies considered for predicting wind power. The first case study involves predicting generated wind power using wind direction data and forecast missing wind speed data. The second case study presents generated wind power using wind speed data and forecast missing wind direction data.

4.2.1. Case Study 1

As mentioned, a FONN model with a single hidden layer was used in the first case study, as depicted in Figure 8. This neural network predicts wind power using wind direction data and forecasts missing wind speed data. The architecture comprises 2 nodes in the input layer, 30 in the hidden layer, and 1 in the output layer. The activation function “Purelin” was chosen for the network’s output layer. In contrast, the hidden layer’s activation function “F” varied between conventional and developed tangential functions to evaluate the performance of the FONN model in terms of R2 and MSE. The analysis of the performance of the activation function is given in Table 3, based on the results obtained, as follows.
The following analysis was conducted to determine the accuracy of different activation functions on Site A during the training and testing phases. The results indicate that the fractional arctan function had the highest accuracy, with R2 values of 0.9749 and 0.9831 during the training and testing phases, respectively, with MSE values of 0.0205 and 0.0142. Similarly, the fractional hard tansig function performed the best, with R2 values of 0.9263 and 0.9369 and MSE values of 0.0424 and 0.0397 during the training and testing phases, respectively. The hard tansig function also performed well, with R2 values of 0.8954 and 0.9075 in the training and testing phases, respectively, and an MSE value of 0.0521 in both phases. On the other hand, the conventional tansig function had the lowest accuracy, with R2 values of 0.8578 and 0.8642 during the training and testing phases, respectively, and MSE values of 0.0753 and 0.0764, respectively.
At Site B, the fractional tansig function performed the best, with R2 values of 0.9428 and 0.9497 during training and testing phases, respectively, and MSE values of 0.0534 and 0.0529. The hard tansig function also performed well, with R2 values of 0.9489 and 0.9517 during the training and testing phases, respectively, and MSE values of 0.0583 and 0.0578, respectively. On the other hand, the worst-performing function for Site B was the conventional tansig function, with R2 values of 0.9328 and 0.9436 in the training and testing phases, respectively, and MSE values of 0.0662 and 0.0652, respectively. Moreover, the fractional hard tansig function performed better during the training and testing phases, with R2 values of 0.9543 and 0.9609 and MSE values of 0.0464 and 0.0432, respectively. The conventional LiSHT function had R2 values of 0.9532 and 0.9584 and MSE values of 0.0428 and 0.0414 during the training and testing phases, respectively. The fractional LiSHT function performed better, with R2 values of 0.9572 and 0.9621 and MSE values of 0.0399 and 0.0386 during the training and testing phases, respectively. For the highest accuracy, the fractional arctan function proved to be the best option, with R2 values of 0.9929 and 0.9952 in the training and testing phases, respectively, and MSE values of 0.0046 and 0.0032, respectively. Similarly, the conventional arctan function also performed well, with R2 values of 0.9901 and 0.9948 during the training and testing phases, respectively, and MSE values of 0.0063 and 0.0035, respectively.
For Site C, the fractional tansig function performed better than the conventional tansig function, with R2 values of 0.8931 and 0.9026 and MSE values of 0.0742 and 0.0629 during the training and testing phases, respectively. Similarly, the fractional hard tansig function performed better than the conventional hard tansig function, with R2 values of 0.9035 and 0.9163 and MSE values of 0.0598 and 0.0586 during the training and testing phases, respectively. For the LiSHT function, the fractional LiSHT function performed slightly better than the conventional LiSHT function, with R2 values of 0.8864 and 0.8973, and MSE values of 0.0752 and 0.0745 during the training and testing phases, respectively. The highest accuracy was achieved using the fractional arctan function, with R2 values of 0.9573 and 0.9635 and MSE values of 0.0123 and 0.0115 during the training and testing phases, respectively. The conventional arctan function also performed well, with R2 values of 0.9469 and 0.9529 and MSE values of 0.0158 and 0.0134 during the training and testing phases, respectively.
Therefore, from the results across all the sites shown in Table 3, the FONN model’s best performance with the conventional arctan function is depicted in Figure 13 and the fractional arctan function is shown in Figure 14 at the hidden layer. The neural network’s performance varied across sites and activation functions. The developed arctan arctan function provided improved results compared to other functions, reflected in higher R2 values and lower MSE values for training and testing across different sites.

4.2.2. Case Study 2

The following analysis presents the second case study, which uses the same network as the previous case, with an identical node count and activation functions. However, this network uses wind speed data and forecast wind direction data as inputs to predict the generated wind power. The obtained results are shown in Table 4, and the analysis of the performance of the activation functions at the three sites is as follows.
Table 4 shows that conventional and fractional functions performed well across all the sites during the training and testing phases. Among the conventional functions, hard tansig and arctan outperformed the others in terms of R2 and MSE values during both training and testing phases. On the other hand, the fractional function performed better overall, with higher R2 values and lower MSE values than the corresponding conventional functions. For instance, for Site A, arctan had the highest R2 value of 0.9898 and the lowest MSE value of 0.0081 for training, while for testing, it had the highest R2 value of 0.9931 and the lowest MSE value of 0.0059. Similarly, the fractional arctan function had the highest R2 value of 0.9899 and the lowest MSE value of 0.0081 for training, while for testing, it had the highest R2 value of 0.9946 and the lowest MSE value of 0.0048. These values indicate that arctan and its corresponding fractional function were the best-performing functions for Site A. The second-best performing function for Site A was hard tansig and its corresponding fractional function. For training, hard tansig had the highest R2 value of 0.9264 and the lowest MSE value of 0.0372, while for testing, it had the highest R2 value of 0.9378 and the lowest MSE value of 0.0346. Similarly, the fractional hard tansig function had the highest R2 value of 0.9726 and the lowest MSE value of 0.0218 for training, while for testing, it had the highest R2 value of 0.9832 and the lowest MSE value of 0.0169. On the other hand, the tansig function had the lowest R2 value of 0.8973 and the highest MSE value of 0.0621 during training. Similarly, during testing, it had the lowest R2 value of 0.9043 and the highest MSE value of 0.0594. Similarly, the fractional tansig function had the lowest R2 value of 0.9264 and the highest MSE value of 0.0519 during training, while it had the lowest R2 value of 0.9329 and the highest MSE value of 0.0497 when tested. These values indicate that tansig and its corresponding fractional function were the worst-performing functions at Site A.
At Site B, the activation function arctan performed the best in both training and testing phases, with R² values of 0.9826 and 0.9875, respectively, and MSE values of 0.0129 and 0.0094, respectively. The arctan arctan function also performed well, with R² values of 0.9835 and 0.9867 and MSE values of 0.0124 and 0.0094, respectively. The best-performing activation function was arctan hard tansig, with R² values of 0.8864 and 0.8949 in the training and testing phases, respectively, and MSE values of 0.0682 and 0.0617, respectively. The tansig and LiSHT activation functions performed well but not as well as the arctan and arctan hard tansig functions. In conclusion, the arctan and arctan hard tansig activation functions performed the best for Site B, with arctan being slightly better regarding the R² value. Similarly, for Site C, the activation function arctan performed the best in both training and testing phases, with R² values of 0.9793 and 0.9866, respectively, and MSE values of 0.0085 and 0.0054, respectively. The arctan arctan function also performed well, with R² values of 0.9816 and 0.9865 and MSE values of 0.0076 and 0.0052, respectively. The second-best-performing activation function was arctan hard tansig, with R² values of 0.9273 and 0.9526 in the training and testing phases, respectively, and MSE values of 0.0341 and 0.0252, respectively. The tansig and LiSHT activation functions also performed well but not as well as the arctan and arctan hard tansig functions. Thus, the arctan and arctan hard tansig activation functions performed the best for Site C, with arctan being slightly better regarding the R² value. The worst-performing activation function was tansig for both conventional and fractional functions.
Therefore, the results presented in Table 4, the best performance of the FONN model with the conventional arctan function, is shown in Figure 15, and the developed arctan arctan activation function is depicted in Figure 16 at the hidden layer during training and testing at all three sites. The conventional arctan and the developed arctan arctan functions exhibit strong predictive abilities compared to other functions at all three sites. Minor variations in the R2 and MSE values demonstrate the consistent and dependable performance of both functions for predicting generated wind power using the provided input data.
The results in both case studies compared the performance of the arctan tangential functions and the conventional tangential functions in predicting wind power across various sites. The arctan arctan function consistently achieved higher R2 values and lower MSE values than the other functions, indicating better predictive capabilities of the FONN model. These findings have important implications for fields that rely on predictive modelling, such as finance, economics, and engineering.

5. Conclusions

A hybrid approach combining LSTMs and FONNs has been presented in this paper to forecast data missing from wind parameters and predict generated wind power across all the sites in the Jeju Island wind farm. An LSTM model was employed to forecast missing wind speed and direction data, obtaining RMSE values of approximately 0.11 and 0.12, respectively. In addition, the FONN model was used to predict wind power with forecast missing wind parameters data through two case studies. In the first case, using wind direction and forecast wind speed data, the developed arctan arctan activation function outperformed the conventional arctan function in the neural network, with high R2 and low MSE values, around 0.97 and 0.003, respectively, during training and testing. Similarly, both activation functions exhibited strong predictive capabilities in predicting wind power using wind speed. During training and testing, the forecast wind direction in the second case achieved high R2 and low MSE values, around 0.98 and 0.004, respectively. The results highlight the potential of the developed arctan arctan function, which consistently proved its effectiveness in enhancing predictive capabilities compared to the conventional arctan function and among all the tangential functions in both case studies. The study provides valuable insights into predicting generated wind power and fills gaps in missing data, demonstrating the potential of advanced neural networks in renewable energy applications. The developed arctan tangential activation functions have improved predictive capabilities compared to the conventional tangential functions, but their increased complexity may limit their practical implementation. In future work, there is a possibility of expanding the analysis carried out on fractional activation functions at α = 0.1 to determine the optimal α value. This extension of α could potentially increase the predictive accuracy of power in wind farms.

Author Contributions

Conceptualisation, B.R. and K.B.; methodology, B.R. and V.R.K.; software, B.R.; validation, B.R., K.B., and V.R.K.; formal analysis, B.R.; investigation, B.R.; resources, V.R.K.; data curation, B.R.; writing—original draft preparation, B.R.; writing—review and editing, K.B. and V.R.K.; visualisation, B.R.; supervision, K.B.; project administration, K.B.; funding acquisition, K.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Short-Term Internal Research Funding (STIRF) with grant number 015LA0-048.

Data Availability Statement

Datasets related to this article can be found at https://figshare.com/articles/dataset/J_2014_01_C_txt/8330285 (accessed on 14 November 2023).

Acknowledgments

The authors further thank the Vellore Institute of Technology in Vellore, India, for their assistance.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kariniotakis, G.; Stavrakakis, G.; Nogaret, E. Wind power forecasting using advanced neural networks models. IEEE Trans. Energy Convers. 1996, 11, 762–767. [Google Scholar] [CrossRef]
  2. Han, Z.; Jing, Q.; Zhang, Y.; Bai, R.; Guo, K.; Zhang, Y. Review of wind power forecasting methods and new trends. Power Syst. Prot. Control 2019, 47, 178–187. [Google Scholar]
  3. Cui, Y.; Li, L.; Chen, D. Ultra-Short-Term Wind Power Load Forecast Based on Least Squares SVM. Electr. Autom. Pap. 2014, 5, 35–37. [Google Scholar]
  4. Zhang, Y.; Wang, P.; Ni, T.; Cheng, P.; Lei, S. Wind power prediction based on LS-SVM model with error correction. Adv. Electr. Comput. Eng. 2017, 17, 3–8. [Google Scholar] [CrossRef]
  5. Pinson, P.; Kariniotakis, G. Wind power forecasting using fuzzy neural networks enhanced with on-line prediction risk assessment. In Proceedings of the 2003 IEEE Bologna Power Tech Conference Proceedings, Bologna, Italy, 23–26 June 2003; Volume 2, p. 8. [Google Scholar]
  6. Shi, J.; Lee, W.J.; Liu, Y.; Yang, Y.; Wang, P. Short term wind power forecasting using Hilbert-Huang Transform and artificial neural network. In Proceedings of the 2011 4th International Conference on Electric Utility Deregulation and Restructuring and Power Technologies (DRPT), Weihai, China, 6–9 July 2011; pp. 162–167. [Google Scholar]
  7. Sideratos, G.; Hatziargyriou, N.D. Probabilistic wind power forecasting using radial basis function neural networks. IEEE Trans. Power Syst. 2012, 27, 1788–1796. [Google Scholar] [CrossRef]
  8. Hong, Y.Y.; Yu, T.H.; Liu, C.Y. Hour-ahead wind speed and power forecasting using empirical mode decomposition. Energies 2013, 6, 6137–6152. [Google Scholar] [CrossRef]
  9. Lotfi, E.; Khosravi, A.; Akbarzadeh-T, M.; Nahavandi, S. Wind power forecasting using emotional neural networks. In Proceedings of the 2014 IEEE International Conference on Systems, Man, and Cybernetics (SMC), San Diego, CA, USA, 5–8 October 2014; pp. 311–316. [Google Scholar]
  10. Chen, N.; Qian, Z.; Nabney, I.T.; Meng, X. Wind power forecasts using Gaussian processes and numerical weather prediction. IEEE Trans. Power Syst. 2013, 29, 656–665. [Google Scholar] [CrossRef]
  11. Çevik, H.H.; Acar, Y.E.; Çunkaş, M. Day ahead wind power forecasting using complex valued neural network. In Proceedings of the 2018 International Conference on Smart Energy Systems and Technologies (SEST), Seville, Spain, 10–12 September 2018; pp. 1–6. [Google Scholar]
  12. Naik, J.; Dash, S.; Dash, P.K.; 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]
  13. Higashiyama, K.; Fujimoto, Y.; Hayashi, Y. Feature extraction of NWP data for wind power forecasting using 3D-convolutional neural networks. Energy Procedia 2018, 155, 350–358. [Google Scholar] [CrossRef]
  14. Abesamis, K.; Ang, P.; Bisquera, F.I.; Catabay, G.; Tindogan, P.; Ostia, C.; Pacis, M. Short-Term Wind Power Forecasting Using Structured Neural Network. In Proceedings of the 2019 IEEE 11th International Conference on Humanoid, Nanotechnology, Information Technology, Communication and Control, Environment, and Management (HNICEM), Laoag, Philippines, 29 November–1 December 2019; pp. 1–4. [Google Scholar]
  15. Ansari, S.; Sampath Vinayak Kumar, T.G.; Dhillon, J. Wind Power Forecasting using Artificial Neural Network. In Proceedings of the 2021 4th International Conference on Recent Developments in Control, Automation & Power Engineering (RDCAPE), Noida, India, 7–8 October 2021; pp. 35–37. [Google Scholar] [CrossRef]
  16. Khelil, K.; Berrezzek, F.; Bouadjila, T. DWT-based Wind Speed Forecasting Using Artificial Neural Networks in the region of Annaba. In Proceedings of the 2020 1st International Conference on Communications, Control Systems and Signal Processing (CCSSP), El Oued, Algeria, 16–17 May 2020; pp. 508–512. [Google Scholar]
  17. Jørgensen, K.L.; Shaker, H.R. Wind power forecasting using machine learning: State of the art, trends and challenges. In Proceedings of the 2020 IEEE 8th International Conference on Smart Energy Grid Engineering (SEGE), Oshawa, ON, Canada, 12–14 August 2020; pp. 44–50. [Google Scholar]
  18. Lipu, M.H.; Miah, M.S.; Hannan, M.; Hussain, A.; Sarker, M.R.; Ayob, A.; Saad, M.H.M.; Mahmud, M.S. Artificial intelligence based hybrid forecasting approaches for wind power generation: Progress, challenges and prospects. IEEE Access 2021, 9, 102460–102489. [Google Scholar] [CrossRef]
  19. Peiris, A.T.; Jayasinghe, J.; Rathnayake, U. Forecasting wind power generation using artificial neural network:“Pawan Danawi”—A case study from Sri Lanka. J. Electr. Comput. Eng. 2021, 2021, 5577547. [Google Scholar] [CrossRef]
  20. He, Y.; Li, H. Probability density forecasting of wind power using quantile regression neural network and kernel density estimation. Energy Convers. Manag. 2018, 164, 374–384. [Google Scholar] [CrossRef]
  21. Wu, Y.X.; Wu, Q.B.; Zhu, J.Q. Data-driven wind speed forecasting using deep feature extraction and LSTM. IET Renew. Power Gener. 2019, 13, 2062–2069. [Google Scholar] [CrossRef]
  22. Ramadevi, B.; Bingi, K. Chaotic time series forecasting approaches using machine learning techniques: A review. Symmetry 2022, 14, 955. [Google Scholar] [CrossRef]
  23. Vaswani, A.; Shazeer, N.; Parmar, N.; Uszkoreit, J.; Jones, L.; Gomez, A.N.; Kaiser, Ł.; Polosukhin, I. Attention is all you need. arXiv 2017, arXiv:1706.03762. [Google Scholar] [CrossRef]
  24. Zhu, Q.; Li, H.; Wang, Z.; Chen, J.; Wang, B. Ultra-short-term prediction of wind farm power generation based on long-and short-term memory networks. Power Grid Technol. 2017, 41, 3797–3802. [Google Scholar]
  25. Wang, S.; Li, B.; Li, G.; Yao, B.; Wu, J. Short-term wind power prediction based on multidimensional data cleaning and feature reconfiguration. Appl. Energy 2021, 292, 116851. [Google Scholar] [CrossRef]
  26. Khochare, J.; Rathod, J.; Joshi, C.; Laveti, R.N. A short-term wind forecasting framework using ensemble learning for indian weather stations. In Proceedings of the 2020 IEEE International Conference for Innovation in Technology (INOCON), Bangluru, India, 6–8 November 2020; pp. 1–7. [Google Scholar]
  27. Kumar, D.; Abhinav, R.; Pindoriya, N. An ensemble model for short-term wind power forecasting using deep learning and gradient boosting algorithms. In Proceedings of the 2020 21st National Power Systems Conference (NPSC), Gandhinagar, India, 17–19 December 2020; pp. 1–6. [Google Scholar]
  28. Zhou, M.; Wang, B.; Guo, S.; Watada, J. Multi-objective prediction intervals for wind power forecast based on deep neural networks. Inf. Sci. 2021, 550, 207–220. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Li, Y.; Zhang, G. Short-term wind power forecasting approach based on Seq2Seq model using NWP data. Energy 2020, 213, 118371. [Google Scholar] [CrossRef]
  30. Kisvari, A.; Lin, Z.; Liu, X. Wind power forecasting—A data-driven method along with gated recurrent neural network. Renew. Energy 2021, 163, 1895–1909. [Google Scholar] [CrossRef]
  31. Lin, W.H.; Wang, P.; Chao, K.M.; Lin, H.C.; Yang, Z.Y.; Lai, Y.H. Wind power forecasting with deep learning networks: Time-series forecasting. Appl. Sci. 2021, 11, 10335. [Google Scholar] [CrossRef]
  32. Cali, U.; Sharma, V. Short-term wind power forecasting using long-short term memory based recurrent neural network model and variable selection. Int. J. Smart Grid Clean Energy 2019, 8, 103–110. [Google Scholar] [CrossRef]
  33. Zhang, K.; Jin, H.; Jin, H.; Wang, B.; Yu, W. Gated Recurrent Unit Neural Networks for Wind Power Forecasting based on Surrogate-Assisted Evolutionary Neural Architecture Search. In Proceedings of the 2023 IEEE 12th Data Driven Control and Learning Systems Conference (DDCLS), Xiangtan, China, 12–14 May 2023; pp. 1774–1779. [Google Scholar]
  34. Arora, P.; Jalali, S.M.J.; Ahmadian, S.; Panigrahi, B.; Suganthan, P.; Khosravi, A. Probabilistic Wind Power Forecasting Using Optimized Deep Auto-Regressive Recurrent Neural Networks. IEEE Trans. Ind. Inform. 2022, 19, 2814–2825. [Google Scholar] [CrossRef]
  35. Miele, E.S.; Ludwig, N.; Corsini, A. Multi-Horizon Wind Power Forecasting Using Multi-Modal Spatio-Temporal Neural Networks. Energies 2023, 16, 3522. [Google Scholar] [CrossRef]
  36. Wu, N.; Green, B.; Ben, X.; O’Banion, S. Deep transformer models for time series forecasting: The influenza prevalence case. arXiv 2020, arXiv:2001.08317. [Google Scholar]
  37. Han, K.; Xiao, A.; Wu, E.; Guo, J.; Xu, C.; Wang, Y. Transformer in transformer. Adv. Neural Inf. Process. Syst. 2021, 34, 15908–15919. [Google Scholar]
  38. Ren, J.; Yu, Z.; Gao, G.; Yu, G.; Yu, J. A CNN-LSTM-LightGBM based short-term wind power prediction method based on attention mechanism. Energy Rep. 2022, 8, 437–443. [Google Scholar] [CrossRef]
  39. Zhou, X.; Liu, C.; Luo, Y.; Wu, B.; Dong, N.; Xiao, T.; Zhu, H. Wind power forecast based on variational mode decomposition and long short term memory attention network. Energy Rep. 2022, 8, 922–931. [Google Scholar] [CrossRef]
  40. Wang, L.; He, Y.; Li, L.; Liu, X.; Zhao, Y. A novel approach to ultra-short-term multi-step wind power predictions based on encoder–decoder architecture in natural language processing. J. Clean. Prod. 2022, 354, 131723. [Google Scholar] [CrossRef]
  41. Wei, H.; Wang, W.s.; Kao, X.x. A novel approach to ultra-short-term wind power prediction based on feature engineering and informer. Energy Rep. 2023, 9, 1236–1250. [Google Scholar] [CrossRef]
  42. Ramadevi, B.; Kasi, V.R.; Bingi, K. Fractional ordering of activation functions for neural networks: A case study on Texas wind turbine. Eng. Appl. Artif. Intell. 2024, 127, 107308. [Google Scholar] [CrossRef]
  43. Esquivel, J.Z.; Vargas, J.A.C.; Lopez-Meyer, P. Fractional adaptation of activation functions in neural networks. In Proceedings of the 2020 25th International Conference on Pattern Recognition (ICPR), Milan, Italy, 10–15 January 2021; pp. 7544–7550. [Google Scholar]
  44. Son, N.; Yang, S.; Na, J. Hybrid forecasting model for short-term wind power prediction using modified long short-term memory. Energies 2019, 12, 3901. [Google Scholar] [CrossRef]
  45. Dubey, S.R.; Singh, S.K.; Chaudhuri, B.B. A comprehensive survey and performance analysis of activation functions in deep learning. arXiv 2021, arXiv:2109.14545. [Google Scholar]
  46. Ding, B.; Qian, H.; Zhou, J. Activation functions and their characteristics in deep neural networks. In Proceedings of the 2018 Chinese Control and Decision Conference (CCDC), Shenyang, China, 9–11 June 2018; pp. 1836–1841. [Google Scholar]
  47. Nwankpa, C.; Ijomah, W.; Gachagan, A.; Marshall, S. Activation functions: Comparison of trends in practice and research for deep learning. arXiv 2018, arXiv:1811.03378. [Google Scholar]
  48. Lederer, J. Activation functions in artificial neural networks: A systematic overview. arXiv 2021, arXiv:2101.09957. [Google Scholar]
  49. Job, M.S.; Bhateja, P.H.; Gupta, M.; Bingi, K.; Prusty, B.R. Fractional Rectified Linear Unit Activation Function and Its Variants. Math. Probl. Eng. 2022, 2022, 1860779. [Google Scholar] [CrossRef]
  50. Sharma, S.; Sharma, S.; Athaiya, A. Activation functions in neural networks. Towards Data Sci. 2017, 6, 310–316. [Google Scholar] [CrossRef]
  51. Adhikari, R.; Agrawal, R.K. An introductory study on time series modeling and forecasting. arXiv 2013, arXiv:1302.6613. [Google Scholar]
  52. Bingi, K.; Prusty, B.R.; Kumra, A.; Chawla, A. Torque and temperature prediction for permanent magnet synchronous motor using neural networks. In Proceedings of the 2020 3rd International Conference on Energy, Power and Environment: Towards Clean Energy Technologies, Shillong, India, 5–7 March 2021; pp. 1–6. [Google Scholar]
  53. Ramadevi, B.; Bingi, K. Time Series Forecasting Model for Sunspot Number. In Proceedings of the 2022 International Conference on Intelligent Controller and Computing for Smart Power (ICICCSP), Hyderabad, India, 21–23 July 2022; pp. 1–6. [Google Scholar] [CrossRef]
  54. Siami-Namini, S.; Tavakoli, N.; Siami Namin, A. A Comparison of ARIMA and LSTM in Forecasting Time Series. In Proceedings of the 2018 17th IEEE International Conference on Machine Learning and Applications (ICMLA), Orlando, FL, USA, 17–20 December 2018; pp. 1394–1401. [Google Scholar] [CrossRef]
Figure 1. The data from three wind farm sites on Jeju Island.
Figure 1. The data from three wind farm sites on Jeju Island.
Fractalfract 08 00149 g001
Figure 2. Pairwise relationships in the data from three wind farm sites on Jeju Island.
Figure 2. Pairwise relationships in the data from three wind farm sites on Jeju Island.
Fractalfract 08 00149 g002
Figure 3. Correlation analysis on the data from three wind farm sites on Jeju Island.
Figure 3. Correlation analysis on the data from three wind farm sites on Jeju Island.
Fractalfract 08 00149 g003
Figure 4. Correlation analysis of wind speed parameter with missing data.
Figure 4. Correlation analysis of wind speed parameter with missing data.
Fractalfract 08 00149 g004
Figure 5. Correlation analysis of wind direction parameter with missing data.
Figure 5. Correlation analysis of wind direction parameter with missing data.
Fractalfract 08 00149 g005
Figure 6. Flowchart of the proposed methodology.
Figure 6. Flowchart of the proposed methodology.
Fractalfract 08 00149 g006
Figure 7. Architecture of LSTM model for forecasting missing input time-series data.
Figure 7. Architecture of LSTM model for forecasting missing input time-series data.
Fractalfract 08 00149 g007
Figure 8. FONN model’s architecture for generated wind power prediction.
Figure 8. FONN model’s architecture for generated wind power prediction.
Fractalfract 08 00149 g008
Figure 9. Response of FONN activation functions at hidden layer.
Figure 9. Response of FONN activation functions at hidden layer.
Fractalfract 08 00149 g009
Figure 10. Response of Purelin activation function at output layer.
Figure 10. Response of Purelin activation function at output layer.
Fractalfract 08 00149 g010
Figure 11. Actual and forecast wind speed data at different sites.
Figure 11. Actual and forecast wind speed data at different sites.
Fractalfract 08 00149 g011
Figure 12. Actual and forecast wind direction data at different sites.
Figure 12. Actual and forecast wind direction data at different sites.
Fractalfract 08 00149 g012
Figure 13. Performance of conventional neural network model during training and testing with forecast missing wind speed.
Figure 13. Performance of conventional neural network model during training and testing with forecast missing wind speed.
Fractalfract 08 00149 g013
Figure 14. Performance of FONN model with forecast missing wind speed.
Figure 14. Performance of FONN model with forecast missing wind speed.
Fractalfract 08 00149 g014
Figure 15. Performance of conventional neural network model during training and testing with forecast missing wind direction.
Figure 15. Performance of conventional neural network model during training and testing with forecast missing wind direction.
Fractalfract 08 00149 g015
Figure 16. Performance of FONN model with forecast missing wind direction.
Figure 16. Performance of FONN model with forecast missing wind direction.
Fractalfract 08 00149 g016
Table 1. Jeju Island’s wind farm data and specifications.
Table 1. Jeju Island’s wind farm data and specifications.
Data AspectSite ASite BSite C
Data Collection Period11 January 2014–25 January 201411 January 2014–20 January 201411 January 2014–25 January 2014
Collection Time Interval10 min10 min10 min
Wind Turbine Specifications
ModelU88U50U50
Output2000 kW750 kW750 kW
Wind SpeedUp to 12 m/sUp to 12.5 m/sUp to 12.5 m/s
Rotor Speed Range6–17.5 rpm9–28 rpm9–28 rpm
Voltage and Frequency690 V/60 Hz690 V/60 Hz690 V/60 Hz
Rotor Diameter88 m50 m50 m
Hub Height80 m50 m50 m
Power ControlPitch RegulationPitch RegulationPitch Regulation
Table 2. Performance comparison of various forecasting models for missing data of wind speed and direction at different sites.
Table 2. Performance comparison of various forecasting models for missing data of wind speed and direction at different sites.
ModelSiteWind Speed (m/s)Wind Direction (deg)
RMSERMSE
LSTMSite A0.180.16
Site B0.4250.185
Site C0.1120.126
NARSite A0.3530.442
Site B0.2970.185
Site C0.4570.395
ARIMASite A0.5830.386
Site B0.4580.572
Site C0.3870.454
Table 3. Performance comparison of different functions in training and testing phases for various sites under case study 1.
Table 3. Performance comparison of different functions in training and testing phases for various sites under case study 1.
SiteConventional
Function
TrainingTestingFractional
Function
TrainingTesting
R2 MSE R2 MSE R2 MSE R2 MSE
Site ATansig0.85780.07530.86420.0764Tansig0.87390.06280.88640.0612
Hard tansig0.89540.05210.90750.0516Hard tansig0.92630.04240.93690.0397
LiSHT0.87490.06830.88730.0621LiSHT0.90250.06120.91730.0598
Arctan0.97270.02270.97330.0207Arctan0.97490.02050.98310.0142
Site BTansig0.93280.06620.94360.0652Tansig0.94280.05340.94970.0529
Hard tansig0.94890.05830.95170.0578Hard tansig0.95430.04640.96090.0432
LiSHT0.95320.04280.95840.0414LiSHT0.95720.03990.96210.0386
Arctan0.99010.00630.99480.0035Arctan0.99290.00460.99520.0032
Site CTansig0.82160.08530.83620.0817Tansig0.89310.07420.90260.0629
Hard tansig0.84530.07320.85640.0695Hard tansig0.90350.05980.91630.0586
LiSHT0.87620.07890.87580.0778LiSHT0.88640.07520.89730.0745
Arctan0.94690.01580.95290.0134Arctan0.95730.01230.96350.0115
Table 4. Performance comparison of different functions in training and testing phases for various sites under case study 2.
Table 4. Performance comparison of different functions in training and testing phases for various sites under case study 2.
SiteConventional
Function
TrainingTestingFractional
Function
TrainingTesting
R2 MSE R2 MSE R2 MSE R2 MSE
Site ATansig0.89730.06210.90430.0594Tansig0.92640.05190.93290.0497
Hard tansig0.92640.03720.93780.0346Hard tansig0.97260.02180.98320.0169
LiSHT0.91630.05830.92890.0542LiSHT0.95170.04870.96190.0453
Arctan0.98980.00810.99310.0059Arctan0.98990.00810.99460.0048
Site BTansig0.82450.09820.84630.0968Tansig0.85620.08410.86780.0832
Hard tansig0.86740.07210.86890.0708Hard tansig0.88640.06820.89490.0617
LiSHT0.84620.08190.85730.0798LiSHT0.86930.07390.87150.0716
Arctan0.98260.01290.98750.0094Arctan0.98350.01240.98670.0094
Site CTansig0.90410.05280.91460.0512Tansig0.93170.04250.94620.0419
Hard tansig0.90890.04810.91630.0479Hard tansig0.92730.03410.95260.0252
LiSHT0.89320.05140.90230.0506LiSHT0.91720.04590.92510.0445
Arctan0.97930.00850.98660.0054Arctan0.98160.00760.98650.0052
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ramadevi, B.; Kasi, V.R.; Bingi, K. Hybrid LSTM-Based Fractional-Order Neural Network for Jeju Island’s Wind Farm Power Forecasting. Fractal Fract. 2024, 8, 149. https://doi.org/10.3390/fractalfract8030149

AMA Style

Ramadevi B, Kasi VR, Bingi K. Hybrid LSTM-Based Fractional-Order Neural Network for Jeju Island’s Wind Farm Power Forecasting. Fractal and Fractional. 2024; 8(3):149. https://doi.org/10.3390/fractalfract8030149

Chicago/Turabian Style

Ramadevi, Bhukya, Venkata Ramana Kasi, and Kishore Bingi. 2024. "Hybrid LSTM-Based Fractional-Order Neural Network for Jeju Island’s Wind Farm Power Forecasting" Fractal and Fractional 8, no. 3: 149. https://doi.org/10.3390/fractalfract8030149

Article Metrics

Back to TopTop