Next Article in Journal
Study on the Effect of Cable Group Laying Mode on Temperature Field Distribution and Cable Ampacity
Previous Article in Journal
Investigations on EMI Mitigation Techniques: Intent to Reduce Grid-Tied PV Inverter Common Mode Current and Voltage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of an SVR Model for the Fault Diagnosis of Large-Scale Doubly-Fed Wind Turbines Using SCADA Data

School of Energy and Power Engineering, Changsha University of Science & Technology, Changsha 410114, China 2 Hubei Key Laboratory of Power System Design and Test for Electrical Vehicle, Hubei University of Arts and Science, Xiangyang 441053, China 3 College of Engineering, Northeastern University, 360 Huntington Ave, Boston, MA 02115, USA 4 Guizhou Key Laboratory of Economics System Simulation, Guizhou University of Finance & Economics, Guiyang, Guizhou 550004, China 5 School of Engineering, University of South Australia, Adelaide, SA 5095, Australia
*
Authors to whom correspondence should be addressed.
Energies 2019, 12(17), 3396; https://doi.org/10.3390/en12173396
Submission received: 16 July 2019 / Revised: 23 August 2019 / Accepted: 27 August 2019 / Published: 3 September 2019
(This article belongs to the Section A3: Wind, Wave and Tidal Energy)

Abstract

:
Fault diagnosis and forecasting contribute significantly to the reduction of operating and maintenance associated costs, as well as to improve the resilience of wind turbine systems. Different from the existing fault diagnosis approaches using monitored vibration and acoustic data from the auxiliary equipment, this research presents a novel fault diagnosis and forecasting approach underpinned by a support vector regression model using data obtained by the supervisory control and data acquisition system (SCADA) of wind turbines (WT). To operate, the extraction of fault diagnosis features is conducted by measuring SCADA parameters. After that, confidence intervals are set up to guide the fault diagnosis implemented by the support vector regression (SVR) model. With the employment of confidence intervals as the performance indicators, an SVR-based fault detecting approach is then developed. Based on the WT SCADA data and the SVR model, a fault diagnosis strategy for large-scale doubly-fed wind turbine systems is investigated. A case study including a one-year monitoring SCADA data collected from a wind farm in Southern China is employed to validate the proposed methodology and demonstrate how it works. Results indicate that the proposed strategy can support the troubleshooting of wind turbine systems with high precision and effective response.

Graphical Abstract

1. Introduction

With environmental concerns of fossil-fuel-based energy consumption, renewable energy has been playing a predominant role in clean and cost-efficient energy production, contributing significantly to the sustainability of human society. One of these renewable energy sources is the large-scale wind turbines with a doubly-fed induction generator (DFIG), which have been widely installed attributing to their low lifecycle costs. In a DFIG wind turbine system, the gearbox plays an important role in the drive train of a wind turbine. However, due to the variable heavy load from turbulent wind and other uncertain factors, gears usually suffer unbalanced and varying loads, which results in dramatic temperature rising and viscosity reduction in the lubricant. As a result, gears might be gradually deteriorated. These faults take the responsibility of unexpected system breakdown and extra expenditures on turbine maintenance [1]. Therefore, fault diagnosis and forecasting are of great importance to the resilience of turbine systems and reduction of maintenance costs. Over the years, a significant body of studies has been conducted targeting the troubleshooting of wind turbine systems. Chen et al. proposed an empirical wavelet transform method for fault diagnosis in wind turbine generator bearing fault diagnosis and found that this method can enhance performance in the weak feature detection of wind turbine drivetrain [2]. Biswal et al. utilized vibration analysis for wind turbine fault diagnosis and conclude that the fault size for gear root cracks could be identified successfully [3]. Moreover, Zhang et al. employed acoustic emission signals for wind turbine gearbox fault diagnosis, as expected, found that the locations of faults can then be determined [4]. The literature review indicates that most of those fault diagnosis methods are based on vibration and acoustic signals, however, vibration signals monitored with additional vibration sensors and equipment will result in extra investment on maintenance, as well as the increase of computational load and difficulty in signal processing.
As to the fault detection and prediction of key components in turbines, recent advances in sensors and signal processing have been contributing significantly to the monitoring and prognosis of WT gearboxes. However, accurate fault detection and reliable diagnosis in wind turbine gearboxes are still challenging due to the inherited complexities in mechanical systems and operation conditions [5]. Thus, more accurate and cost-efficient online monitoring is of essential importance on troubleshooting and enhancement of turbine system resilience. An effective condition monitoring might either evaluates component health conditions or helps with the detection and prediction of marginal changes that indicate incipient faults [6]. The WT SCADA system with key parameters can be easily measured and monitored, including bearing and lubricant temperatures, rotor speed, environment temperature, and active power. Besides, it can continuously record the historical and present status of a wind turbine system.
The literature review also shows some successful studies in predicting faults caused by complex nonlinearity reasons using operational SCADA data combined with diverse approaches such as adaptive neuron-fuzzy interference systems (ANFIS), Bayesian Networks and Deep Learning Networks [7,8,9]. A hybrid stochastic technique is proposed in reference [10], which is based on an augmented observer combined with Eigen structure assignment for the parameterization and the genetic algorithm (GA) optimization to address the attenuation of uncertainty mostly generated by disturbances. Data mining techniques are also applied to this data combination to model the power curve [11]. The results of reference [12] demonstrate that the proposed method can effectively analyze nonlinear data trends, continuously monitor the WT and reliably detect abnormal problems by using six process parameters. An exponentially weighted moving average (EWMA) control chart was proposed in reference [13] and used in Statistical Process Control to determine if the process is out of control. Wang et al. also presented a deep auto-encoder (DAE) method for anomaly detection and fault analysis of wind turbine components [14]. Although some successful applications have been achieved, the limitations of the mentioned approaches cannot be overlooked. Firstly, it is difficult to process and interpret a large body of data collected from the SCADA system simultaneously. Secondly, conventional artificial intelligent techniques are inefficient in the extraction of features from raw data [15]. Thirdly, the potential of the SCADA data might have not been fully explored [16]. In this study, an algorithmic-level online WT fault detection method based on SCADA data using support vector regression (SVR) is proposed. The employed SVR algorithm, which is based on statistical learning and structural risk minimization is a nonlinear generalization of the Generalized Portrait algorithm developed in 1964 [17,18].
In light of the above, this research is established with the following operations. Firstly, in the case of decreasing structural risk, SVR can maximally excavate the implicit classification knowledge in the SCADA data. Secondly, as the number of faulty samples is relatively smaller compared with the number of right samples, SVR can address the imbalanced data issue. Thirdly, SVR can address the nonlinear system in fault diagnosis. In the developed SVR approach, SCADA data will be preprocessed and divided into two groups. The first group termed as “Training Data” is used to obtain the optimization function with the “soft margin”, which allows some “ ϵ insensitive   loss ”. The second group termed as “Test Data” is then employed to validate the model performance. Once the measured data exceeds the confidential range, a developing fault then can be spotted.
To present the proposed methodology in a logic manner, the rest of this paper is organized as follows: the developed methodology is introduced in Section 2, followed by the statistic control methods demonstrated. Section 3 is focused on the development of an SVR model for wind turbine gearbox systems using 45 SCADA parameters. After that, a case study is employed to demonstrate and validate the established system in Section 4. Finally, Section 5 draws to concluding remarks and recommendations for future studies.

2. Methodology

Support vector machines (SVM), known as a kind of learning machine underpinned by the statistical learning theory, were firstly proposed by Vapnik et al. [19]. Support vector regression (SVR) is a machine learning algorithm based on the principle of structural risk minimization is a common application form of SVMs. The SVR is a powerful function approximation technique of outstanding generalization performance, which can be used to obtain globally optimal solutions, as well as map a nonlinear regression problem into a linear regression problem by applying a kernel function. SCADA monitoring provides a series dataset including bearing and lubricant temperatures, rotor speed and active power. Compared to the condition monitoring based on vibration and acoustic signals, SCADA monitoring independent of additional sensors and previous fault datasets. However, raw SCADA data acquired from wind turbines requires precise mathematical modeling. And the mathematical model is sensitive to external disturbance. In addition, accurate modeling requires specification parameters and prior data under abnormal working conditions, which limited the application in diverse gearboxes under variable working conditions. Therefore, an SVR algorithm is developed in the research to explore the ability of self-study and disposal of complex nonlinearity. Moreover, only standardized real-time SCADA data is used in the construction of the SVR model to reduce the impacts of external disturbances. The methodology and framework of this study can be shown in Figure 1.
To support the construction and optimization of the proposed SVR model, a grid search method based on cross-valid training data which is initially introduced in reference [20], is employed for the SVR parameters optimization. After that, the temperature detection model of wind turbine gearboxes was setup, followed by the bearing and lubricate temperatures under normal working conditions were predicted. Then real-time data was fed into the SVR model for the calculation of the estimated residual. In order to detect the changes in online data, a moving average calculation was applied. Finally, the residual with distributing was operated in statistical process control. Once the residuals step out of the thresholds, incipient faults can then be predicted.

2.1. Support Vector Regression

SVM is a machine learning tool widely used in classification and regression analysis. The SVM regression is known as a nonparametric technique since it relies on kernel functions. Compared to other existing fault detection methods, the SVM is specialized in address problems with small sample size, nonlinear and high-dimensional datasets. Moreover, with the introduction of soft margin, the generalization ability of SVM is significantly enhanced. The initial design of SVM is to map the input vectors { ( x 1 , y 1 ) , , ( x l , y l ) } ϵ X × into a higher dimensional feature space. Then the linear regression in this high dimensional space will be solved. The input space is defined by Kernel functions K ( x t ,   y j ) [21].
Given the training data { ( x 1 , y 1 ) , , ( x l , y l ) } ϵ X × , where l denotes the total numbers of data and x i demonstrate the values of independent variables or training data in this study. y i is the output while X denotes the space of the input patterns (e.g., X d ), one objective of this research is to find a function f ( x ) that has at most ε deviation from the actually obtained targets y i for all the training data, and at the same time to set it as flat as possible. One of these functions is the Gaussian kernel, with the maximum tolerance of the deviation between f ( x ) and y is ϵ . By introducing a dual set of variables, the function then can be expressed as:
m i n i m i z e 1 2 w 2 + C i = 1 l ( ξ i + ξ i * )
s u b j e c t   t o { y i w , x i b ε + ξ i w , x i + b y i ε + ξ i * ξ i , ξ i *                                         0
where the trade-off is determined by the constant C . The ε i n s e n s i t i v e loss function | ξ | ε in the context of deviations larger than ε , can be described by
| ξ | ε = { 0                               i f   | ξ | ε | ξ | ε         o t h e r w i s e
with the introduction of Lagrange multipliers, Equation (4) is then obtained.
L = 1 2 w 2 + C i = 1 l ( ξ i + ξ i * ) i = 1 l ( η i ξ i + η i * ξ i * ) i = 1 l α i ( ε + ξ i y i + w , x i + b ) i = 1 l α i * ( ε + ξ i * + y i w , x i b )
In this equation, L is the Lagrangian and η i , η i * , α i , α i * are Lagrange multipliers. Hence the dual variables in Equation (5) have to satisfy the positivity constraints shown in Equation (5). Note that α i ( * ) refers to α i and α i * here.
α i ( * ) ,   η i ( * )   0
It follows from the saddle point condition that the partial derivatives of L with respect to the prime variables ( w ,   b ,   ξ i ,   ξ i * ) must vanish for the optimality as shown.
b L = i = 1 l ( α i * α i ) = 0
w L = w i = 1 l ( α i * α i ) x i = 0
ξ i * L = C α i ( * ) η i * = 0
By substituting Equations (6)–(8) into Equation (4), the dual optimization problem can then be obtained.
m a x i m i z e { 1 2 i = 1 l ( α i α i * ) ( α j α j * ) x i ,   x j ε i = 1 l ( α i + α i * ) + i = 1 l ( α i α i * )
s u b j e c t   t o ε i = 1 l ( α i α i * ) = 0   a n d   α i ,   α i * [ 0 , C ]
In deriving Equation (9) the dual variables η i , η i * were eliminated through a condition Equation (8) which can be reformulated as η i * = C α i * . Equation (7) can then be rewritten as
w = i = 1 l ( α i α i * ) x i
Thus, the SVR detection model can be expressed as
f ( x ) = i = 1 l ( α i α i * ) x i ,   x + b
The complete algorithm can then be described in terms of dot products between the data acquired by the wind turbine SCADA system.

2.2. Statistic Control Methods

In order to identify if the wind turbine is working in normal, the statistic control method is applied to the time series of sample data. Mean, standard deviation, minimum and maximum of the data are collected within control limits. There are several rules to determine whether the data is out of control limits. In this study, the normal distribution was applied for the wind turbine SCADA data. Here, the normal distribution set up by central limit theorem as the distribution to which the mean value of almost any set of independent and randomly generated variables rapidly converges is an important class of statistical distributions.
The probability density function of the normal distribution is given as:
f ( x ) = 1 σ 2 π e ( x μ ) 2 σ 2
where μ is the mean or expectation of the distribution, σ is the standard deviation and σ 2 is the variance. Different values of μ and σ would produce different normal density curves and different normal distributions. The normal density can be specified by means of an equation. The probability distribution of a random variable x is identified to be normal if it has a probability density. Figure 2 shows the changes in the normal distribution with the standard deviation. As shown in Figure 2, 68% of the data falls in [ μ σ ,   μ + σ ] standard deviation of the average, 95% of the data is within [ μ 2 σ ,   μ + 2 σ ] the standard deviations of the average, while 99.7% of the data falls in [ μ 3 σ , μ + 3 σ ] the standard deviations of the average. If the data is within [ μ 3 σ ,   μ + 3 σ ], the process will be identified as working in normal. However, once the data goes beyond the control limits [ μ 3 σ ,   μ + 3 σ ], the process will be considered as out of control, and then faults are detected.

3. Model Construction

3.1. Faults Detection Model of Wind Turbine Gearboxes

SVR is a detection algorithm based on statistical learning theory. The initial design of this algorithm is to minimize error, individualize the hyperplane and set a margin of tolerance. Analysis of the SCADA data in this study found that the values of parameters such as bearing temperature, lubricant temperature, and rotor speed are of high interactions.
The SVR detection model can be expressed as
K ^ b ( x b ) = i = 1 l ( α i α i * ) x i ,   x + b
K ^ o ( x o ) = i = 1 l ( α i α i * ) x i ,   x + b
where f ^ b ( x b ) and f ^ o ( x o ) denote the estimated values of bearing temperature and lubricate temperature. x b = [ T b ( t ) ,   T b ( t 1 ) , ,   T b ( t k 1 ) ,   T e ( t k 1 ) ,   ,   T e ( t k 2 ) , v ( t 1 ) , ,   v ( t k 3 ) ,   P ( t 1 ) , ,   P ( t k 4 ) ] T is the bearing input. x o = [ T o ( t ) ,   T o ( t 1 ) ,   ,   T o ( t l 1 ) ,   T c ( t 1 ) ,   ,   T e ( t l 2 ) ,   v ( t 1 ) ,   ,   v ( t l 3 ) ,   P ( t 1 ) ,   ,   P ( t l 4 ) ] T is the temperature input. Note that T b ( t i ) ,   T o ( t i ) ,   T e ( t i )   , v ( t i ) ,   P ( t i ) are previous sampling data of bearing temperature, lubricant temperature, environment temperature, rotor speed and active power at time i . k 1 ,   k 2 ,   k 3 ,   k 4 ,   l 1 ,   l 2 ,   l 3 and l 4 are historical sampling time of state variables. K ^ b ( x b t ,   x b ) and K ^ o ( x o t , x o ) is the kernel function of bearing temperature data and the lubricate temperature data. The Gaussian kernel is given as
K ( x t ,   x j ) = e 0.5 x t x j 2 σ 2
where σ is a given constant. One significant advantage of using Gaussian kernel is the possibility of operating in infinite dimension space, which results in that the feature space is easier to handle due to the linearity in feature space. Three of the parameters can be adjusted for setting an SVR with the Gaussian kernel, namely, C ,   ε , and σ . The SVR performance has a heavy dependence on the triplet of parameters. The parameter selection aligns with the value selection of the decision variables C ,   ε , and σ , which can maximize the SVR performance on test data. In SVR models, the number of data sets is measured by the sliding window N with N is further divided into 5 elements. After that, the optimal model parameters are obtained by using the fivefold cross-validation. Finally, the optimal model is used in online fault detection. In a 50% cross-validation, 40% of the sample is used for training and 10% of the sample is used for testing.

3.2. Experimental Setup and Procedure

As discussed in the previous section, parameter selection is of great importance in SVR operations. To address this concern, grid search is adopted in the scheme. Regarding the optimizations of the regularization parameter C and the size of the error cross-insensitive zone ε , the grid search was functioned with an exhaustive parameter optimization method. The SVR model will perform optimally only in the case those parameters have been selected properly. The grid search is based on a defined subset of the hyper-parameter space with minimal value, maximal value, and the number of steps input. As shown in Figure 3, the SCADA datasets had been divided into five subsets to perform the fivefold cross-validation [17]. Where, one of these subsets was used for testing, while the others were used in training. Thus, different combinations of parameters C and σ combine the 2-dimensional grid. Only the best combination will be selected for the SVR model training.

3.3. Gearbox Faults Detection Methods and Residual Control

In order to detect the incipient faults of gearboxes, warning and alarming limits need to be set. In the normal working condition, the residual of detected data and the online real-time data fall within the control limits [ μ 3 σ ,   μ + 3 σ ]. However, when the wind turbine suffers from sudden uneven loads, the residual could deviate from the normal working space. The significant changes in the estimated residual could result in a varied distribution of the mean value and standard deviation. As this sudden change would not appear as a trend in time series, thus, once the estimated residual of the data keeping increase and step out of the thresholds, a fault is then identified.
In the proposed model, one year of monitoring data was collected as a baseline to perform a validation of normal distribution. The mean value μ and the standard deviation σ of the residual value, as well as the estimated residual value, is then used Equation (17) to support troubleshooting. In this formula, with the mean value-centered, [ μ σ ,   μ + σ ] employed as the warning limits while [ μ 3 σ ,   μ + 3 σ ] will play as the alarming limits.
μ = 1 n i = 1 n e i = 1 n i = 1 n ( T i T ^ i )
σ 2 = 1 n i = 1 n ( e i μ ) 2

3.4. Gearbox Temperature SVR Model Residual Statistic Analysis

In wind turbine SCADA systems, the bearing temperature and the lubricant temperature of gearboxes are recorded with an interval of two seconds, which helps real-time monitoring and fault detection. When the wind turbine gearboxes experienced unexpected faults, the recorded vector in the normal working space would be updated. This would result in a significant change in the residual of the predicted data and real-time data. The mean value and the standard deviation indicated that the distribution of the residual would be abnormal. To detect the changes of parameters in the time series, a moving average calculation is then applied. For a given period, the residual sequence of gearbox temperature from the SVR model can be expressed as
ε T = [ ε 1 ,   ε 2   ,   ,   ε N ,   ]
A time window of width N is used to calculate the moving average or mean, as well as the standard deviation for the N consecutive residuals in the window.
X ¯ ε = 1 N i = 1 N ε
S ε = 1 N 1 i = 1 N ( ε i X ¯ ε ) 2
In normal conditions, the residual mean value should be around zero, and the standard deviation should be with a constant value. As mentioned above, gearbox faults can be identified when the residual mean value and standard deviation start to deviate. Specifically, in the first situation, the residual mean value remains zero, but the standard deviation steps out of the lower and upper boundary. However, in the second situation, the residual mean value varies significantly, while the standard deviation remains relatively constant. In the third situation, both the residual mean value and the standard value deviate significantly.

4. Case Study and Discussions

In this section, a one-year SCADA dataset of the WT from a wind farm located in Southern China has been employed for the validation of the proposed methodology. The framework of the studied wind turbine system is shown as Figure 4, which consists of blades, a multi-stage gearbox, a DFIG, and a frequency converter and control unit. The gearbox is driven by blades to spin the generator to electricity production. The stator of the DFIG converts the mechanical energy with high-speed rotation into power output via the gearbox. Wind turbine faults are found mostly from key elements such as the gearbox, blades, and the generator. These faults are usually associated with high maintenance costs and long downtime of turbine systems. Therefore, troubleshooting and efficient fault identification of gearboxes are of great significance to reduce the lifecycle expenditures on wind turbines, as well as to increase productivities of wind farms.

4.1. Structure of the Wind Turbine Gearbox and the SCADA System

Large-scale wind turbine systems are developed for wind energy harvesting in remote regions, where strong winds are available. The turbine nacelle is sitting on top of a tower which is usually more than 80meters (m) in height. Figure 4 illustrates the major components of a typical utility-scale wind turbine drivetrain, which is composed of main bearing, main shaft, gearbox, brake, generator shaft, and generator. The generator converts the mechanical energy with high speed to power energy. The blades and rotor hub are supported by main bearings.
The SCADA system at the wind farm records all wind turbine parameters with a frequency of two seconds. As shown in Figure 5, each record includes a timestamp, output power, stator current and voltage, wind speed, environmental and nacelle temperatures, as well as the generator stator winding and cooling air temperatures [18]. Indeed, there are 34 variables in total including internal parameters such as bearing and lubricant temperatures, as well as external parameters like environmental temperature obtained from the SCADA systems installed at the wind farm. In which the lubricant temperature will be sampled with a resolution of two seconds. These data will be preprocessed through data normalization before analysis. At the same time, the SCADA system keeps a recording of turbine operating and fault information such as startup, shutdown, generator over temperature and pitch system fault. Each fault record is composed of a timestamp, a state number, and the fault information.

4.2. Data Preprocessing

To reduce the calculation errors contributed by the numerical differences result from diverse wind turbine systems and to maintain the consistency of the original data structure, the SCADA data is normalized within an interval of [0, 1]. The normal area value is 3 m/s–25 m/s, and the rated wind speed is 12 m/s. The cut-in wind speed is 3 m/s and cut-out wind speed at 25 m/s. As can be seen from Table 1.
Firstly, the mechanism of wind turbines is analyzed; after that, by combining with expert experience, the comprehensive characteristics of collected data are obtained; and finally, 34 variables are set to support the fault detecting and predicting. Backpropagation neural network (BPNN) including three types of layers: the input layer, the hidden layers, and the output layer. The hidden layers is established to fit the characteristics of our sample data and prevent the over fit. We proposed L2 regularization function in the BPNN.
The architecture of the three-layer BPNN is shown in Figure 6. The circles in Figure 6 denotes the neurons of different layers. In this article, we adopted sigmoid function as activation function:
y = 1 1 + e x #
From the wind turbine SCADA datasets, the 34 systematic features listed in Table 1 can be obtained. To explore the interactions among SCADA parameters, all the monitored data should be normalized. In addition, data normalization can speed up the coverage of the gradient descent algorithm which then leads to the performance enhancement in the machine learning process. In this study, features were calibrated with a linear function, which ensures the values of each feature in data unit-variance. We use the normalization to avoid overweight that certain features of different dimensions may cause. The calibration can be expressed as
x ¯ i = x i x m i n   x m a x x m i n  
where x m i n is the lower bound; x m a x is the upper bound, and x ¯ i is the normalized value. Please note that the parameters x m a x and x m i n can only be computed within the training data, which would be used in training, validation and testing in the following stages.

4.3. Results and Discussions

With the establishment of the temperature detection model and the processing of all the monitoring data, the WT gearbox working conditions can then be analyzed.
(1) Normal Operation Conditions of Gearboxes
Figure 7 shows the predicted and actual operating temperatures of the gearbox lubricant and bearing. Although there are numerical differences between the measured and estimated values, they present consistency in value changes, or it can be concluded that they are of the same tendency in value variations. Figure 8 is further employed to find errors in the values of residuals. The diagram indicated that the results detected by the SVM model are desirable when the gearbox is working in normal conditions. With the intervals [ μ 2 σ ,   μ + 2 σ ] and [ μ 3 σ ,   μ + 3 σ ] are established as warning limit and alarming limit respectively (shown as threshold 1 and threshold 2 respectively in Figure 8, most values fall in within the thresholds, which indicates that the gearbox was working in normal conditions. In this study, mean absolute percentage error (MAPE), root-mean-square error (RMSE), mean absolute error (MAE), mean absolute percentage error (MAPE) are employed to measure the accuracy of the model. RMSE is an estimator that can measure the deviation between the observed value and the real value. While the mean absolute MAE is a measure of the difference between two continuous variables that can be used to reflect the accuracy of the model. MAPE considering both the error between the true value and the predicted value, and the ratio between the error and the true value, thus, it supports a good evaluation of the stability of the model. In Figure 7, the MAPE, RMSE, MAE are 0.4572, 0.3500, and 0.0074, respectively.
(2) Abnormal Operating Conditions of Gearboxes
Figure 9 illustrates a detection analysis of an abnormal gearbox installed at the 5# wind turbine on July 10th. As shown in the diagram, the MAPE, RMSE, MAE is 0.4572, 0.3500, and 0.0074, respectively. While the oil temperatures of the gearbox predicted with the SVM model demonstrates a close matching with the monitoring values. That is further supported by the diagram shown in Figure 10. It then can be concluded that the residual between predicting and monitoring values is growing. As a result, the residual has gradually deviated from the thresholds. This trend also indicates that there is an unbalanced load resulting in friction between the main shaft and the bearing.
Back Propagation Neural Network (BPNN) is a basic neural network widely used in fault diagnosis, which is generally built as a feed-forward multi-layer network according to the error backpropagation algorithm. To further validate the proposed methodology, a comparative study was also conducted between the methods developed in this research with a BPNN. The results are shown in Table 2 and Figure 11. As shown in Table 2, the average troubleshooting rates achieved by SVR are 0.97%, 7.47%, 3.24%, and 24.26% higher than BPNN in dataset 1 to dataset 4 respectively. While Figure 11 presents the mean and Standard Deviation of the BPNN and SVR fault detection accurracy. It then can be concluded that SVR shows lower variability than BPNN in the experiments conducted.

5. Concluding Remarks

This study presents an SVR based method which is of the ability in troubleshooting for large-scale wind turbines. An SVR model has been developed for the implementation of fault diagnosis and forecasting with the monitoring data collected from a SCADA system involved for analysis. A case study was carried out to validate the proposed methodology, and the results indicated that the SVR model can forecast faults effectively. In addition, a comparative study was employed to validate the efficacy of the proposed method. And the results indicated that the SVR based method developed in this research is of better performance in troubleshooting compared with a BPNN. Experiments also demonstrated that the SVR model can provide a feasible solution for the real-time online detection of wind turbine faults.
In light of the above findings, suggestions for future studies might be focused on the following fields: (1) preprocessing of the wind turbine fault diagnosis data and the selecting of fault features, which are the foundations of SVR-based troubleshooting operations. (2) Studies of the imbalance problems in wind turbine troubleshooting categories, where the amount of normal class samples is of predominate importance while the size of fault class samples is of unneglectable effects. However, it is often difficult to obtain desirable troubleshooting with single-disciplinary machine learning algorithms.

Author Contributions

M.T. conceived and designed the topic and performed the experiments of the paper; W.C. wrote original draft; Q.Z. made revisions on this paper; H.W. and B.H. contributed materials and made suggestions for revision; M.T., W.L., L.L. and K.Z. provided guidance for modifying the manuscript; All of the authors read and approved the final manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 61403046 and 51908064), the Natural Science Foundation of Hunan Province, China (Grant No. 2019JJ40304), Changsha University of Science and Technology “The Double First Class University Plan” International Cooperation and Development Project in Scientific Research in 2018 (Grant No. 2018IC14 and 2018IC19), Hubei Superior and Distinctive Discipline Group of Mechatronics and Automobiles (Grant No. XKQ2018002), Hunan Provincial Department of Transportation 2018 Science and Technology Progress and Innovation Plan Project (Grant No. 201843), the Key Laboratory of Renewable Energy Electric-Technology of Hunan Province, the Key Laboratory of Efficient & Clean Energy Utilization of Hunan Province, as well as Major Fund Project of Technical Innovation in Hubei (Grant No. 2017AAA133).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Artigao, E.; Martín-Martínez, S.; Honrubia-Escribano, A.; Gómez-Lázaro, E. Wind turbine reliability: A comprehensive review towards effective condition monitoring development. Appl. Energy 2018, 228, 1569–1583. [Google Scholar] [CrossRef]
  2. Chen, J.; Pan, J.; Li, Z.; Zi, Y.; Chen, X. Generator bearing fault diagnosis for wind turbine via empirical wavelet transform using measured vibration signals. Renew. Energy 2016, 89, 80–92. [Google Scholar] [CrossRef]
  3. Biswal, S.; George, J.D.; Sabareesh, G. Fault size estimation using vibration signatures in a wind turbine test-rig. Procedia Eng. 2016, 144, 305–311. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Lu, W.; Chu, F. Planet gear fault localization for wind turbine gearbox using acoustic emission signals. Renew. Energy 2017, 109, 449–460. [Google Scholar] [CrossRef]
  5. Salameh, J.P.; Cauet, S.; Etien, E.; Sakout, A.; Rambault, L.; Processing, S. Gearbox condition monitoring in wind turbines: A review. Mech. Syst. Signal Process. 2018, 111, 251–264. [Google Scholar] [CrossRef]
  6. Gibert, K.; Marti-Puig, P.; Cusidó, J.; Solé-Casals, J.J.E. Identifying health status of wind turbines by using self organizing maps and interpretation-oriented post-processing tools. Energies 2018, 11, 723. [Google Scholar]
  7. Bangalore, P.; Letzgus, S.; Karlsson, D.; Patriksson, M. An artificial neural network-based condition monitoring method for wind turbines, with application to the monitoring of the gearbox. Wind Energy 2017, 20, 1421–1438. [Google Scholar] [CrossRef]
  8. De Bessa, I.V.; Palhares, R.M.; D’Angelo, M.F.S.V.; Chaves Filho, J. Data-driven fault detection and isolation scheme for a wind turbine benchmark. Renew. Energy 2016, 87, 634–645. [Google Scholar] [CrossRef]
  9. Zhao, H.; Liu, H.; Hu, W.; Yan, X. Anomaly detection and fault analysis of wind turbine components based on deep learning network. Renew. Energy 2018, 127, 825–834. [Google Scholar] [CrossRef]
  10. Odofin, S.; Bentley, E.; Aikhuele, D. Robust fault estimation for wind turbine energy via hybrid systems. Renew. Energy 2018, 120, 289–299. [Google Scholar] [CrossRef]
  11. Marti-Puig, P.; Gibert, K.; Cusidó, J.; Solé-Casals, J.J.E. A text-mining approach to assess the failure condition of wind turbines using maintenance service history. Energies 2019, 12, 1982. [Google Scholar]
  12. Dao, P.B.; Staszewski, W.J.; Barszcz, T.; Uhl, T. Condition monitoring and fault detection in wind turbines based on cointegration analysis of scada data. Renew. Energy 2018, 116, 107–122. [Google Scholar] [CrossRef]
  13. Cambron, P.; Masson, C.; Tahan, A.; Pelletier, F. Control chart monitoring of wind turbine generators using the statistical inertia of a wind farm average. Renew. Energy 2018, 116, 88–98. [Google Scholar] [CrossRef]
  14. Wang, Z.; Wang, J.; Wang, Y.J.N. An intelligent diagnosis scheme based on generative adversarial learning deep neural networks and its application to planetary gearbox fault pattern recognition. Neurocomputing 2018, 310, 213–222. [Google Scholar] [CrossRef]
  15. Khan, S.; Yairi, T.J.M.S.; Processing, S. A review on the application of deep learning in system health management. Mech. Syst. Signal Process. 2018, 107, 241–265. [Google Scholar] [CrossRef]
  16. Dai, J.; Yang, W.; Cao, J.; Liu, D.; Long, X. Ageing assessment of a wind turbine over time by interpreting wind farm scada data. Renew. Energy 2018, 116, 199–208. [Google Scholar] [CrossRef]
  17. Smola, A.J.; Schölkopf, B. A tutorial on support vector regression. Stat. Comput. 2004, 14, 199–222. [Google Scholar] [CrossRef] [Green Version]
  18. Shamshirband, S.; Petković, D.; Amini, A.; Anuar, N.B.; Nikolić, V.; Ćojbašić, Ž.; Kiah, M.L.M.; Gani, A.J.E. Support vector regression methodology for wind turbine reaction torque prediction with power-split hydrostatic continuous variable transmission. Energy 2014, 67, 623–630. [Google Scholar] [CrossRef]
  19. Vapnik, V.; Golowich, S.E.; Smola, A.J. Advances in neural information processing systems. In Support Vector Method for Function Approximation, Regression Estimation and Signal Processing, Proceedings of the NIPS’96 9th International Conference on Neural Information Processing Systems, Denver, CO, USA, 3–5 December 1996; MIT Press: Cambridge, MA, USA, 1996. [Google Scholar]
  20. Syarif, I.; Prugel-Bennett, A.; Wills, G.J.T. Svm parameter optimization using grid search and genetic algorithm to improve classification performance. Telkomnika 2016, 14, 1502. [Google Scholar] [CrossRef]
  21. Castellani, F.; Astolfi, D.; Sdringola, P.; Proietti, S.; Terzi, L. Analyzing wind turbine directional behavior: Scada data mining techniques for efficiency and power assessment. Appl. Energy 2017, 185, 1076–1086. [Google Scholar] [CrossRef]
Figure 1. Fault detection underpinned by support vector regression (SVR).
Figure 1. Fault detection underpinned by support vector regression (SVR).
Energies 12 03396 g001
Figure 2. Probability density curve of normal distribution.
Figure 2. Probability density curve of normal distribution.
Energies 12 03396 g002
Figure 3. SVM parameter optimization using grid search.
Figure 3. SVM parameter optimization using grid search.
Energies 12 03396 g003
Figure 4. Framework of a wind turbine system.
Figure 4. Framework of a wind turbine system.
Energies 12 03396 g004
Figure 5. Data acquisition process of a WT’s SCADA system.
Figure 5. Data acquisition process of a WT’s SCADA system.
Energies 12 03396 g005
Figure 6. The architecture of the three-layer BPNN.
Figure 6. The architecture of the three-layer BPNN.
Energies 12 03396 g006
Figure 7. Monitoring and detecting temperatures of a normal operating gearbox tank.
Figure 7. Monitoring and detecting temperatures of a normal operating gearbox tank.
Energies 12 03396 g007
Figure 8. Residual control diagram of a normal operating gearbox tank.
Figure 8. Residual control diagram of a normal operating gearbox tank.
Energies 12 03396 g008
Figure 9. Actual temperature and detection temperature of an abnormal operating gearbox tank.
Figure 9. Actual temperature and detection temperature of an abnormal operating gearbox tank.
Energies 12 03396 g009
Figure 10. Residual control diagram of an abnormal operating gearbox tank.
Figure 10. Residual control diagram of an abnormal operating gearbox tank.
Energies 12 03396 g010
Figure 11. Comparison between BPNN and SVR fault detection accuracies.
Figure 11. Comparison between BPNN and SVR fault detection accuracies.
Energies 12 03396 g011
Table 1. SCADA data list for the development of normal behavior model.
Table 1. SCADA data list for the development of normal behavior model.
Name of VariableUnitNormal BehaviorName of VariableUnitNormal Behavior
Spinner temp.°CYESPower outputWYES
Hub controller temp.°CYESReactive powerKVArYES
Pitch angle°CYESGrid rotor inverter ph.1 temp.°CYES
Hydraulic oil temp.°CYESGrid rotor inverter ph.2 temp.°CYES
Rotor speedRpmYESGrid rotor inverter ph.3 temp.°CYES
Gear bearing temp. (HSS)°CYESConverter cooling water temp.°CYES
Gear oil temp.°CYESConverter choke coil temp.°CYES
Generator speedRpmYESConverter controller temp.°CYES
Generator bearing temp. 1°CYESTop controller temp.°CYES
Generator bearing temp. 2°CYESGrid busbar temp.°CYES
Generator slip ring temp.°CYESHV transformer ph.1 temp.°CYES
Generator ph.1 temp.°CYESHV transformer ph.2 temp.°CYES
Generator ph.2 temp.°CYESHV transformer ph.3 temp.°CYES
Generator ph.3 temp.°CYESNacelle temp.°CYES
Generator current ph.1°CYESWind speedm/sYES
Generator current ph.2°CYESWind direction°NO
Generator current ph.3°CYESAmbient temp°CNO
Table 2. Wind turbine fault detection rates with BPNN and SVR.
Table 2. Wind turbine fault detection rates with BPNN and SVR.
Dataset NameBPNNSVR
Dataset184.29% (±0.0005)85.26% (±0.0012)
Dataset284.30% (±0.0006)91.77% (±0.0007)
Dataset384.73% (±0.0843)87.97% (±0.0023)
Dataset464.06% (±0.0010)88.32% (±0.0031)

Share and Cite

MDPI and ACS Style

Tang, M.; Chen, W.; Zhao, Q.; Wu, H.; Long, W.; Huang, B.; Liao, L.; Zhang, K. Development of an SVR Model for the Fault Diagnosis of Large-Scale Doubly-Fed Wind Turbines Using SCADA Data. Energies 2019, 12, 3396. https://doi.org/10.3390/en12173396

AMA Style

Tang M, Chen W, Zhao Q, Wu H, Long W, Huang B, Liao L, Zhang K. Development of an SVR Model for the Fault Diagnosis of Large-Scale Doubly-Fed Wind Turbines Using SCADA Data. Energies. 2019; 12(17):3396. https://doi.org/10.3390/en12173396

Chicago/Turabian Style

Tang, Mingzhu, Wei Chen, Qi Zhao, Huawei Wu, Wen Long, Bin Huang, Lida Liao, and Kang Zhang. 2019. "Development of an SVR Model for the Fault Diagnosis of Large-Scale Doubly-Fed Wind Turbines Using SCADA Data" Energies 12, no. 17: 3396. https://doi.org/10.3390/en12173396

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