Next Article in Journal
The Impact of Investments in Physical Capital, Labor, and Knowledge Capital on Enterprise Market Value: Estimation and Optimization
Next Article in Special Issue
A Coupled Simulated Annealing and Particle Swarm Optimization Reliability-Based Design Optimization Strategy under Hybrid Uncertainties
Previous Article in Journal
Predator–Prey Model Considering Implicit Marine Reserved Area and Linear Function of Critical Biomass Level
Previous Article in Special Issue
A New Loss Function for Simultaneous Object Localization and Classification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Residual Strength Modeling and Reliability Analysis of Wind Turbine Gear under Different Random Loadings

School of Mechanical Engineering, Xinjiang University, Urumqi 830046, China
*
Author to whom correspondence should be addressed.
Submission received: 31 August 2023 / Revised: 18 September 2023 / Accepted: 19 September 2023 / Published: 21 September 2023

Abstract

:
A novel method is proposed to investigate the pattern of variation in the residual strength and reliability of wind turbine gear. First, the interaction between loads and the effect of the loading sequence is considered based on the fatigue damage accumulation theory, and a residual strength degradation model with few parameters is established. Experimental data from two materials are used to verify the predictive performance of the proposed model. Secondly, the modeling and simulation of the wind turbine gear is conducted to analyze the types of fatigue failures and obtain their fatigue life curves. Due to the randomness of the load on the gear, the rain flow counting method and the Goodman method are employed. Thirdly, considering the seasonal variation of load, the decreasing trend of gear fatigue strength under multistage random load is calculated. Finally, the dynamic failure rate and reliability of gear fatigue failure under multistage random loads are analyzed. The results demonstrate that the randomness of residual strength increases with increasing service time. The seasonality of load causes fluctuations in the reliability of gear, providing a new idea for evaluating the reliability of the wind turbine gear.

1. Introduction

The energy resource of wind power generation is green and environmentally friendly, and the reliability of wind turbines has become a hot research topic in recent years. Pérez et al. [1] focused on the failure rate of the wind turbine components and found that the gearbox and blades were the most likely locations to cause faults, and they proposed that state detection can improve the reliability of wind turbines. Kusiak and Li [2] predicted the fault situation of the wind turbine based on data from state detection to ensure its reliability. Chen et al. [3] proposed a transfer condition monitoring based on feature alignment and parameter fine-tuning for more accurate monitoring, which can more accurately identify the status of wind turbines. Dui et al. [4] proposed a resilience optimization model of a multi-component system to determine the degree of influence of components on the reliability of the wind turbine system, thus improving the reliability of the wind turbine system. Su et al. [5] established a stochastic differential equation state model to describe the degradation of wind turbines, preventing wind turbine failure. Research has shown that focusing on the faults and reliability of wind turbine components is necessary to increase the reliability of wind turbines.
Wind turbine gears are one of the core components of the wind turbine system, which transfer power and adjust the wind speed and output torque. If the gear fails, it will directly affect the operating efficiency and output power of the wind turbine, causing the production line to stagnate and resulting in significant losses [6]. Therefore, the reliability of the gear directly affects the reliability of the wind turbine system. The reliability evaluation of wind turbine gears can predict the probability of gear failure and possible failure modes. It provides necessary information for preventive maintenance of the wind turbine, ensures safe and stable operation of the wind turbine, and can reduce shutdown costs. Reliability-based design optimization methods and control methods to improve the reliability of wind power systems have also been widely recognized and studied in recent years [7,8,9,10]. In conclusion, the calculation and analysis of gear reliability are very important.
In order to evaluate the reliability of gears more accurately, it is necessary to fully and realistically consider influencing factors such as gear load cycling and material quality. Wind turbine gears subjected to random load changes will experience fatigue damage, resulting in reduced fatigue strength. The decrease in gear strength will cause the problem of tooth surface wear or fatigue fracture during operation, which will affect the reliability of the gears [11]. The ability of a gear to resist external loads after a certain degree of fatigue damage is defined as the residual strength.
Researchers have proposed a variety of residual strength models. For instance, Schaff and Davidson [12] proposed a general residual strength model of materials, and its key parameters need to be fitted by the experimental data of materials. Based on the hypothesis of Hahn and Kim [13], Chou and Croman [14] established the residual strength model of time by the relationship between strength and life. Ganesan and Joanna [15] considered the effect of the stress ratio of the continuous load on the component and proposed a two-parameter residual strength model of glass-fiber-reinforced polymer. Cheng and Hwu [16] obtained the residual strength degradation trajectory of the composite material based on Schaff’s model.
Some researchers have also developed residual strength models for metals, in addition to the residual strength model for composite materials. For instance, in the aspect of damage development, Li et al. [17] established the residual strength degradation model of the material based on a large number of experimental data of 35CrMo. Gao and An [18] developed a residual strength model that requires a small amount of data and takes into account the dispersion of residual strength based on the stress interference theory. Based on the fatigue damage theory and the interval theory, Jiang et al. [19] established a two-parameter residual strength model using experimental data and obtained fuzzy dynamic reliability. In recent years, many researchers have used data-driven methods to predict the degradation of residual strength. For instance, Yang et al. [20] used an artificial neural network to predict the residual strength of composites after impact. Lu et al. [21] used principal component analysis combined with a multi-objective optimization vector machine to predict the residual strength of the pipeline. Miao and Zhao [22] proposed a method of deep extreme machine learning to predict the residual strength of defective pipes. Meng et al. [23] proposed an adaptive Kriging model to improve the efficiency of optimal design for offshore wind turbines.
In summary, most of the residual strength models are used to describe composite materials. Significantly, there is no consistent residual strength model for both composite materials and metal materials. Furthermore, most of the existing residual strength models require a large amount of experimental data to predict. However, many materials or parts require long experimental cycles and high costs. For this, a simple model based on the fatigue damage accumulation theory is proposed, which requires a small number of parameters. The effect of the loading condition, loading sequence, and interaction between loads on strength degradation is fully considered. Moreover, the degradation trajectory of fatigue strength can only be predicted by understanding the fatigue life data of materials or mechanical parts.
Wind energy is an unstable energy source, with variable wind speeds and directions. Variations in wind speed will impact the load on the gear directly. Furthermore, changes in the operational status of the wind turbine will also affect the gear load. From this point of view, the load on the wind turbine gear shows obvious randomness. Actually, the wind load is not completely irregular. The wind speed in different regions varies seasonally, such as in southeastern China, where the wind speed is usually higher during the winter and summer but weaker in autumn [24]. In addition, the wind loads vary in different seasons due to the seasonal nature of the loads applied to the gears of wind turbines. As mentioned above, in order to more closely characterize the residual strength of wind turbine gears and more accurately predict their reliability, the stochastic and seasonal nature of the loads to which they are subjected needs to be fully taken into account.
The remainder of the paper is organized as follows. In Section 2, the boundary conditions of the residual strength model are described, and the existing nonlinear fatigue damage accumulation model is introduced. Based on this model, a residual strength model considering the loading sequence and load interaction is established. The experimental data of two materials are used to validate the proposed residual strength model and compare it with Schaff’s model. In Section 3, the main fatigue failure modes of the wind turbine gear are analyzed, and the fatigue life curve of the gear is obtained by modeling and simulation. Then, the statistical analysis of the four-season wind load on the gear is carried out, and the degradation trend of the residual strength of the gear is obtained. In Section 4, the reliability of wind turbine gear is obtained by the Monte Carlo simulation method based on the random stress-residual strength interference theory. Section 5 summarizes some main conclusions and proposes some further study directions.

2. Residual Strength Model of Multistage Loading Based on the Fatigue Damage Accumulation Theory

The strength degradation of materials has always been a hot topic in the field of reliability. The gear will suffer fatigue damage under the continuous action of cyclic loads. From a macro perspective, the fatigue damage of materials can be evaluated by measuring the residual strength changes of the material [25]. Thus, it can be considered that the strength degradation is closely related to the degree of fatigue damage accumulation. Conversely, the trend of residual strength degradation can also be measured by fatigue damage. Based on these theories, the existing nonlinear fatigue cumulative damage theories are introduced. We then explain the boundary conditions of the residual strength model. Finally, a residual strength model for multistage loading is proposed from the perspective of cumulative fatigue damage, and the proposed model is validated.

2.1. Nonlinear Fatigue Damage Accumulation Model

To date, the linear damage accumulation theory proposed by Miner is widely used in engineering, since its form is concise and simple for calculation [26]. With the development of damage theory, it is found that damage is not only related to the size of the load but also to the loading sequence and the interaction between the loads. Due to the fact that fatigue damage of gears is a complex nonlinear process in practical engineering, establishing a nonlinear damage model is more practical. Many researchers have established nonlinear damage accumulation models from different perspectives. Manson and Halford [27] used the damage curve method to establish a nonlinear damage accumulation model. Freudenthal and Heller [28] developed a fatigue damage model considering the interaction between variable amplitude loads. Zhu and Huang [29] established a damage model based on the law of creep. Zhu et al. [30] introduced the energy damage parameter to define uncertain fatigue damage. Risitano et al. [31] quantified fatigue damage accumulation based on the relationship between temperature and loads.
Considering the loading sequence effect, Peng et al. [32] proposed a nonlinear damage accumulation model based on the energy method, which can be formulated as
d D d n = 1 ln N f 1 N f 1 D
where  D  is fatigue damage,  n  is the number of load cycles, and  N f  is the fatigue life of parts or materials.
The model is verified by 45 steel, 16 Mn steel, and Al-2024 aluminum alloy. It is found that the damage prediction has good accuracy.
By integrating Equation (1), the following expression can be obtained:
D = ln 1 n N f ln N f
From the above equation, it can be seen that when  n = 0 , the cumulative damage  D = 0 . When  n = N f 1 , the total cumulative damage  D  reaches 1. If the load is applied once more, the material will fail. Therefore, the material fails when the number of load cycles is  N f  [26,33].

2.2. Boundary Conditions of the Residual Strength Model

In general, the residual strength of the material is related to the initial strength of the material, the quality of processing and manufacturing, the loading method of the load, and the size of the load. According to the existing residual strength model [34,35], it can be found that the formula generally satisfies the following two conditions:
(1) The initial residual strength is the static strength of the material.
r 0 = σ b
where  r 0  is the residual strength of the load 0 times and  σ b  is the static strength.
(2) When the material or part fails, the residual strength is equal to the peak value of the load.
r N f 1 = S max
where  S max  is the maximum value of the load on the part or material.
Among numerous residual strength models, the model proposed by Schaff is quite consistent with various practical situations. Due to the fewer parameters required to determine and its relatively high accuracy, this model is widely used in practical engineering applications. Schaff’s model contains two parameters: The initial residual strength and the maximum value of the load, which also meet the boundary conditions of Equations (3) and (4) mentioned above. Therefore, a residual strength model based on Schaff’s model [12] is proposed, which can be formulated as
r n = r 0 r 0 S max Y
where  Y  should make the residual strength expression satisfy the above two boundary conditions, which can be formulated as
Y = 0 , n = 0 1 , n = N f 1
The decrease in residual strength is related to the degree of fatigue damage accumulation. It is also found that  D  satisfies the condition of  Y . Thus, let  D = Y . We substitute Equation (2) into Equation (5) to obtain the residual strength model considering the loading sequence effect, which can be formulated as
r n = r 0 r 0 S max ln 1 n N f ln N f
From Equation (7), it can be seen that the proposed residual strength model satisfies the initial state of static strength, and the load on the material at failure is equal to the maximum load. With the increase in loading times  n , the residual strength  r n  decreases monotonously. Using this model, the residual strength degradation law of the material under the load can be obtained.

2.3. Residual Strength Model under Multistage Loading

In engineering practice, mechanical components are usually subjected to constantly changing loads. Therefore, it is necessary to consider how to characterize residual strength when loading multistage loads. If the residual strength model of the two-level load is known, then the residual strength model of the multistage load only needs to be analogized on this basis. Therefore, in this paper, we derive the residual strength model expression of two levels for load.
If the first-level loads and the second-level loads are applied separately to the same material, the rate of decline of the two residual strengths will necessarily be different. In other words, there are two different curves. According to the equivalent theory of fatigue damage accumulation, it is necessary to equalize the residual strength of the first-level loads to the residual strength curve of the second-level loads. Therefore, a suitable equivalence formula needs to be established to perform the equivalence of the residual strength under different levels of loading. The equal percentage reduction in residual strength under different loads is defined as the equivalent basis. The percentage expression of the decrease in residual strength can be formulated as
P = r 0 r n   r 0 S max
The percentage of residual strength reduction after the first-level loads is equal to the percentage of residual strength reduction after the second-level loads, which can be formulated as
P 1 = P 2
In the theory of fatigue damage accumulation, some researchers considered the interaction between loads [33,36]. The ratio of the two-level load is introduced into the expression of damage equivalence to characterize this influencing factor. In this paper, this theory is also applied to the residual strength equivalent theory, and the load ratio factor is introduced into the equivalent expression. Then, the equivalent residual strength can be formulated as
P 1 = P 2 σ 2 σ 1
r 1 0 r 1 n 1   r 1 0 S 1 max = r 2 0 r 2 n 2 r 2 0 S 2 max σ 2 σ 1
where  n 1  is the number of first-level actions to load,  n 2  is the number of times that the residual strength after the first level of loads is equivalent to the second level of loads, and the second level of loads need to be applied.  σ 1  is the mean stress of the first level of loads and  σ 2  is the mean stress of the second level of loads.
Based on the concept of fatigue damage theory, the residual strength of two levels of load is equivalent. The equivalent loading number  n 2  of the second-level load can be obtained and added to the original loading number  n 2  of the second-level load to obtain the remaining strength after the second-level load is loaded. Therefore, it is necessary to derive the equivalent loading number  n 2  of the second-level load. Substituting the residual strength formula (7) into formula (11) can obtain
r 1 0 r 1 0 r 1 0 S 1 max ln 1 n 1 N f 1 ln N f 1   r 1 0 S 1 max = r 2 0 r 2 0 r 2 0 S 2 max ln 1 n 2 N f 2 ln N f 2 r 2 0 S 2 max σ 2 σ 1
By simplifying Equation (12), the equivalent number of actions  n 2  can be obtained, which can be formulated as
n 2 = N f 2 N f 2 1 + ln 1 n 1 N f 1 ln N f 1 σ 1 σ 2
The first level of loads are  n 1  times, and the residual strength after  n 2  times of the second level of loads are equivalent to  n 2 + n 2  times of the second level of loads. The residual strength after two levels of loads can be obtained, which can be formulated as
r n = r 0 r 0 S max ln 1 n 2 + N f 2 N f 2 1 + ln 1 n 1 N f 1 ln N f 1 σ 1 σ 2 N f ln N f
In summary, when the material is subjected to the multistage load, the calculation is performed according to the above ideas based on the residual strength equivalent theory.

2.4. Verification of Residual Strength Model

At present, the typical residual strength model is Schaff’s model. Due to its few parameters and high precision, it is widely used in practical engineering. The degradation model described by Schaff’s model is generally consistent with the actual degradation situation. Therefore, the proposed model is compared with the typical Schaff’s model in this section. The materials used to verify model performance include the residual strength data of 45 steel under first-level loading and the experimental data of 30CrMnSiA steel under second-level loading from the published literature [32].

2.4.1. One-Level Loading

45 steel was subjected to a constant amplitude load of 310.37 MPa. The number of load cycles, the experimental residual strength data, the residual strength predicted by Schaff’s model, and the residual strength predicted by the proposed model are shown in Table 1.
The calculation results of the two models are compared, as shown in Figure 1.
In Table 1, it can be seen that the calculation results of Schaff’s model are closer to the second and third points of the experiment. However, the overall results of the proposed model perform better predictive performance. In Figure 1, it can be seen that the maximum error of Schaff’s model is 7.45% and the maximum error of the proposed model is 5.07%, indicating that the error of the proposed model is closer to the true value. The  R 2  of Schaff’s model is 0.9870, while the  R 2  of the proposed model is 0.9901, indicating that the predicted residual strength of the proposed model fits better with the actual value. Both Schaff’s model and the proposed model show a decreasing pattern of strength in their residual strength curves. The residual strength decays slowly in the early stages and decreases abruptly in the later stages, leading to sudden failure of the material.

2.4.2. Two-Level Loading

The experimental data of 30CrMnSiA and aluminum alloy Al-2024 is also used to validate the proposed model [32]. Both materials are subjected to two levels of horizontal loading, and the experimental data of the two materials in different load sequences are shown in Table 2 and Table 3.
Schaff’s model and the proposed model are used to calculate the residual strength degradation curve of the material under two levels of sequential loading, as shown in Figure 2 and Figure 3.
In order to evaluate the predictive performance of the model, the maximum error and the goodness of fit between the predicted and true values are calculated, as shown in Figure 4 and Figure 5.
From Figure 4 and Figure 5, it can be clearly seen that the  R 2  of the proposed model is higher than that of Schaff’s model for both materials, regardless of the high-to-low loading order or the low-to-high loading order. The maximum error of the proposed model is smaller than that of Schaff’s model. Therefore, it can be considered that the proposed residual strength model has high accuracy for both high-to-low loading and low-to-high loading. Moreover, in Figure 2 and Figure 3, it can be seen that the loading sequence has an impact on the strength degradation trend of the load. Therefore, it is necessary to consider the loading sequence of the load in predicting the trend of strength degradation. The proposed model reflects this influencing factor well, which indicates that it is suitable for residual strength prediction.

3. Modeling of Residual Strength of Wind Turbine Gears

The forms of fatigue failure of gears with different materials and parameters are also different, due to tooth root bending fatigue failure or tooth surface contact fatigue failure [37]. Therefore, it is necessary to analyze whether the maximum stress on the gear occurs at the bending of the tooth root or at the contact of the tooth surface. In addition, the proposed residual strength model requires the S-N curve of the gear, and the S-N curve of the material is different from the S-N curve of the part. If the S-N curve of the gear material is used directly for the calculation of the residual strength, there will be a greater deviation in the calculation results. Therefore, it is necessary to obtain the force analysis and S-N curve of the gear by experimental or simulation methods.

3.1. Fatigue Failure Mode Analysis of Gears and Fatigue Life Curve Fitting

It is a common research method to analyze the force on machine parts by software simulation. Some researchers compared the simulation results of fatigue analysis software or finite element software with experimental results [38,39]. It is proven that software simulation tools can quickly obtain the stress and fatigue life results of components, which have certain reference significance. Furthermore, it significantly reduces the analysis cycle time and cost. Thus, computer simulation technology can be used to efficiently and quickly complete the stress analysis and fatigue life analysis of gears.
The working principle of the gear fatigue testing machine is simulated, and the Workbench software is used for static simulation. The version of the Workbench is 2022 R1. The first step is to simplify the model appropriately, and single-tooth loading is adopted. In order to eliminate the impact of non-critical components on the simulation results, the model is simplified based on the working principle of the fatigue testing machine. Moreover, only a portion of the gear and support heads are retained, as shown in Figure 6.
According to experimental principles, the position of the load action is set at the intersection point  O  between the tangent line of the gear base circle and the tooth profile. The loading direction  F  of the load is the perpendicular downward direction of the normal tooth profile passing through point  O .
The simulation analysis process is as follows:
(1)
Establish a three-dimensional model as shown in Figure 7.
(2)
Set the material parameters. The elastic modulus of the material 20CrMnTi is 207 GPa, Poisson’s ratio is 0.25, the hardness is 40 HB, and the tensile strength is 2747 MPa. Considering the influence of the deformation of the contact head on the simulation results, it is set as a rigid body.
(3)
Mesh generation. Firstly, the whole mesh is divided by a multi-zone. Then, the contact surface between the gear and the upper and lower support heads is densified locally, and the mesh of the tooth root part is refined.
(4)
Set the contact type. In the part of the contact between the support head and the tooth surface, the contact is not separated.
(5)
Set loads and boundary conditions. The upper support head only retains the movement of the z-axis. Fully fixed constraints are applied to the lower support head. The inner hole of the gear only retains the rotational freedom of the y-axis. A vertical downward load is applied to the upper end of the upper support head. Five groups of loads are set up, which are 200 KN, 400 KN, and 600 KN, respectively. In this study, the simulation results of 400 KN are employed to illustrate the results.
The equivalent stress cloud diagram of the gear obtained by the software is shown in Figure 8.
The static analysis shows that the maximum equivalent stress of the gear is 805.22 Mpa on the side of the tooth root. From the results of modeling and simulation, it can be concluded that the main failure form of the gear is tooth root bending fatigue failure. The accuracy of the model is evaluated by theoretical calculation. According to the calculation formula of tooth root bending fatigue stress in the published literature [40], when the gear is subjected to the tangential load of 400 KN, the bending stress of the tooth root is 809.47 Mpa. The theoretical calculation value is close to the simulation value, so it can be considered that the model is reasonable and can be used for subsequent life analysis. Finally, using Ncode (2022 R1) software analysis and setting the load as symmetric cyclic stress loading, the stress ratio is −1. Then the life values of each point at the tooth root can be obtained. The cloud diagram is illustrated in Figure 9.
Points are taken at the location of the maximum stress at the root bends under multistage load, and the stress and life value of each point are taken. Thus, the stress life curve of the gear can be drawn, as shown in Figure 10.

3.2. Bending Stress Distribution of Gear under Four-Season Loads

The strength degradation is closely related to the load level. It is very necessary to appreciate the bending stress of the gear when calculating the degradation path of the bending strength of the gear. In the natural environment, the wind load on the wind turbine is random, so the bending stress of the gear is also random. In this section, random stress is converted into regular stress that can be directly used for calculation. Taking a wind speed of 78 m as an example [41], the bending stress of the tooth root of the gear is calculated and the residual strength trajectory is analyzed.

3.2.1. The Bending Stress of Gear Is Calculated by Wind Speed

The force on the gear of the wind turbine is determined by the wind speed at the location of the wind turbine. The Weibull distribution is generally used to describe the wind speed [42]. The mean wind speed and variance of the four seasons are known. Then, the calculation formulas of the shape parameter  f  and scale parameter  g  of the Weibull distribution can be formulated as [43]
f = H v E v 1.086
g = E v Γ 1 f + 1
where  v  is the random wind speed,  H v  is the variance of the random wind speed, and  E v  is the mean value of the random wind speed.
The shape and scale parameters of the Weibull distribution of the four seasons are calculated using Equations (15) and (16) based on the known mean and variance of wind speed in reference [36]. The estimated calculation results of the wind speed mean, variance, and distribution parameters are shown in Table 4.
The torque of the gear can be obtained from the random wind speed [44]. According to the torque of the gear, the tangential load can be easily obtained. Then, the bending stress of the gear is calculated. The bending stress can be formulated as [40]
σ = K A K V K β K α F t b m Y F Y S Y ε Y β
where  K A  is the service coefficient,  K V  is the dynamic load coefficient,  K β  is the tooth load distribution coefficient,  K α  is the tooth load distribution coefficient,  F t  is the circumferential force on the pitch circle of the end face,  b  is the smaller tooth width in meshing gears,  m  is the normal modulus,  Y F  is the tooth profile coefficient under the load and addendum,  Y S  is the stress correction coefficient under the action of the load and tooth tip,  Y ε  is the coefficient of the coincidence degree, and  Y β  is the helix angle coefficient.
The tooth width of the sun gear is 170 mm, and its normal modulus is 16 [45]. The parameter values in the calculation formula of gear bending stress are shown in Table 5.
Thus, the gear bending stress of four seasons can be obtained, as shown in Figure 11.

3.2.2. Random Stress Is Treated as Symmetrical Cyclic Stress

The stress calculated by the random wind load in each season is random stress, which is also known as variable amplitude stress. The proposal of the residual strength model is based on the concept of fatigue damage, which requires the stress life curve of gear for residual strength model calculation. The S-N curve of the gear is simulated and fitted under symmetric cyclic loading with a load stress ratio of −1. Therefore, it is necessary to convert the actual random load into a symmetric load, as shown in Figure 12.
First, the rain flow counting method is used to perform cyclic counting of stresses with the same mean and amplitude [46]. Then, the Goodman method is used to unify the stress ratios of each group, and the equivalent stresses for four seasons can be obtained. By fitting the distribution of equivalent force, it can be used to calculate residual strength.
After processing the four seasons’ loads using the rain flow counting method, the following results can be obtained, as shown in Figure 13.
The mean stress of the load obtained by the rain flow counting method is usually not equal to 0, which means the stress ratio is not equal to −1. This problem can also be observed from the two-dimensional load spectrum, as shown in Figure 13. Therefore, in order to effectively estimate the fatigue life of gear using S-N curves, it is necessary to convert the mean and amplitude of each level of load in the two-dimensional load spectrum into equivalent loads corresponding to the S-N curve [47]. Further processing of seasonal loads was carried out according to Goodman’s fatigue empirical formula
S = σ d σ d S m
where  S  is the equivalent load,  σ d  is the tensile strength,  S a  is the amplitude of the load, and  S m  is the mean load.
In general, the stress distribution of the gear can be fitted using the Weibull distribution [48]. Therefore, a two-parameter Weibull distribution fitting is performed on the equivalent stress of the four seasons, and the cumulative probability density function is used to represent the fitting effect. The fitting results of equivalent stress in four seasons are shown in Figure 14.
The results show that the CDF of the Weibull distribution almost coincides with the CDF of equivalent stress samples. The estimated values of Weibull distribution parameters for stress in each season are shown in Table 6.
In summary, by fitting the random stress with a mean of 0, the bending stress distribution of the gear in spring, summer, autumn, and winter is obtained.

4. Analysis of the Residual Strength and Dynamic Reliability of Gear

In general, the strength degradation of machine parts and materials is closely related to the properties of the materials and the loads to which they are subjected. The load on the wind turbine gear is random. The uncertainty of the load makes the strength degradation uncertain. The randomness of strength degradation makes the reliability analysis and prediction of mechanical structures more complex and difficult. Therefore, in order to better predict wind turbine gear bending failure, it is necessary to fully consider the randomness of strength degradation. In this section, the residual strength variation trend and dynamic reliability of gear under multistage random loads in spring, summer, autumn, and winter are calculated and analyzed.

4.1. Residual Strength Model for Tooth Root Bending Strength

If the equivalent stress of a season is considered a one-level load, then the gear will be subjected to four-level loads every year, in the order of spring, summer, autumn, and winter. After each level of load is applied, Equation (14) needs to be used for residual strength equivalence, and then the residual strength after the next level of load is calculated. The corresponding life after  n  times of bending stress is also the random variable that follows a certain distribution, so the residual strength has randomness at the determined time  t . Next, we calculate the residual strength of the gear according to Equations (7) and (14) for residual strength. However, it is difficult to calculate the residual strength of the gear under such a multistage random load using conventional probability methods. Therefore, the Monte Carlo method is used to deal with such problems because of its good adaptability and randomness. It can be applied to address various problems and can simulate the relationship between a variety of complex probability distributions and random variables [49,50]. As the number of simulations increases, the simulation results are closer to the real situation and can provide more accurate estimates [51]. The flowchart for calculating the residual strength of gear using the Monte Carlo method is shown in Figure 15.
We calculate the residual strength at the end of each season and obtain the residual strength path of the gear within 15 years. The calculation results are shown in Figure 16.
In Figure 16, it can be seen that the residual strength of the gear shows the degradation trend over time. Due to the randomness of stress, the residual strength of the gear also represents randomness. At time  t , the residual strength of the gear is not a fixed value but a random variable that follows a distribution. As the service life increases, the range of gear residual strengths gradually widens. In order to better describe the law of the residual strength of gear, the distribution of the residual strength is fitted for each season. In general, a two-parameter Weibull distribution can be used to describe the residual strength distribution after any given fatigue cycle [16]. The Q-Q and CDF tests show that the residual intensity at the end of each season follows a Weibull distribution. Taking the spring residual strength of the first year as an example, the shape parameter of the fitted Weibull distribution is 34,678.07, and its scale parameter is 3746.73. The test results of the Q-Q chart are shown in Figure 17.
The fitted Weibull distribution is close to the actual residual strength quantile and the Q-Q plot is in a straight line, and  R 2  between each quantile and the diagonal is 0.9938. From the non-parametric CDF test, it can be seen that the maximum vertical difference between the actual data and the fitted data is very small, and the two CDF curves almost coincide. Therefore, it can be considered that using the Weibull distribution to fit the residual strength at time  t  is reasonable.
Since the residual strength of time  t  can be described by the Weibull distribution, the distribution of residual strength changing with time can be described by the Weibull distribution of parameters changing with time. After obtaining the scale and shape parameters of the Weibull distribution fitted by residual strength at the end of each season, the variation of the parameters over time is analyzed. The fitting results of parameters over time are shown in Figure 18.
The coefficient of determination of scale parameters is 0.9855 and the coefficient of determination of the shape parameter is 0.9986. Both values are very close to 1, indicating that the fitting result is relatively suitable. Therefore, the residual strength of gears follows a two-parameter Weibull distribution, and the scale parameter  k  and shape parameter  w  at time  t  can be formulated as
k = 6.901 × 10 8 t 2.725 + 2706
w = 2.863 × 10 8 t 1.984

4.2. Gear Reliability Analysis Considering the Failure of Tooth Root Bending

There are many methods for reliability analysis. For instance, Huang et al. [52] combined the chi-square approximation and numerical integration methodologies to calculate the reliability of positioning accuracy. Li et al. [53] used a Bayesian copula network based on a fuzzy rough set to evaluate and make decisions. Yazdi et al. [54] used an improved fault tree method to analyze reliability. Yan et al. [55] evaluated the reliability of multiple failure modes based on the method of failure modes and effects criticality analysis. The performance of mechanical components usually varies over time and there are multiple failure modes. Yu et al. [56,57,58,59] have studied the time-varying reliability calculation method for complex products. In this section, in addition to calculating the static reliability of gears for four seasons, the time-varying reliability of gears is also considered.
The stress level of gear varies in different seasons, and it can be assumed that the stress distribution of the four seasons does not change with the increase in gear operating years. In other words, within the prediction of 15 years, it can be assumed that the stress level remains constant throughout the four seasons of each year. Without taking into account the degradation in the strength of the gear, the stress–strength interference theory suggests that the reliability of the gear in the same season will remain the same regardless of the year. According to the calculation, the reliability of gears at the end of spring, summer, autumn, and winter is 0.9995, 0.9988, 0.9973, and 0.9964, respectively. It can be seen that the reliability of gear is the highest in spring.
When considering the strength degradation of a gear, the stress on the gear is a seasonal distribution, and the remaining strength of the gear is a temporal distribution. Based on the stress strength interference theory, the criterion for gear failure is that if the residual strength is less than the random load, the gear will fail. The reliability and failure rate results are calculated by the Monte Carlo method, as shown in Figure 19 and Figure 20.
In Figure 19 and Figure 20, it can be seen that the reliability of gear does not monotonically decrease over time but rather has a significant span from the end of winter to the end of spring each year. Although the decreasing trend of residual strength is continuous, the load on the gear varies from season to season as a result of the influence of stress levels. The explanation diagram of the reliability jump is shown in Figure 21.
The remaining strength of the gear gradually decreases over time. The strength of the gear at the end of spring in the first year is  Y 1 , and after one year of service, the strength of the gear at the end of winter in the second year is  Y 4 . The stress level  S 1  of gear in spring is lower than the stress level  S 2  in winter. The stress level is only related to the season and does not change with the increase in service life. The strength  Y 1  at the end of spring in the first year is higher than the strength  Y 2  at the end of winter in the first year. Under the influence of stress and strength, the reliability of spring in the first year is higher than that of winter. The residual strength  Y 3  at the end of spring in the second year is lower than the strength  Y 2  in winter in the first year, but the stress level  S 1  in spring is much lower than the stress level  S 2  in winter. The reliability at the end of spring in the second year will be higher than that at the end of winter in the first year.
As mentioned above, it can be concluded that the differences in stress levels among different seasons have a significant impact on gear reliability. The reliability trend over the four seasons is determined by both the stress level and residual strength of the gear concurrently. In the spring of each year, the reliability of the gear is the highest, since the load on the gear is the smallest, and the residual strength of the gear is the highest. This result is consistent with the reliability results of the gear without considering strength degradation. The changes in stress levels throughout the four seasons do not make the reliability of gear monotonically decrease over time, but rather experience a sudden increase in reliability.
By observing the variation pattern of reliability at the same stress level, the following three aspects can be determined:
  • The overall decreasing trend of reliability is determined by the degradation trend of gear strength. The gear’s reliability in each season decreases monotonically from year to year, which indicates that the reliability of the gear has decreased compared to the previous year in all seasons. Due to the sudden increase in the decay rate of the residual strength, the reliability of the gear suddenly decreases rapidly in the 15th year.
  • By observing the distance between the four seasonal reliability curves, it can be found that the distance between the summer and autumn reliability curves is the largest, the distance between the spring and summer curves ranks second, and the distance between the autumn and winter curves is the smallest. It can be considered that the impact of load amplitudes in the four seasons causes a change in the distance between the reliability curves of adjacent seasons. In Table 4, it can be seen that there is a significant difference in load amplitudes between summer and autumn, resulting in a significant downward shift in the reliability curve from summer to autumn. The difference in amplitude between autumn and winter loads is the smallest, so the overall downward movement of the reliability curve from autumn to winter is relatively small.
  • The difference in reliability between the four seasons is relatively small in the first few years. In the first nine years, the reliability of gears fluctuated between 1 and 0.99 per year. As the service life increases, the reliability difference of the gear in the four seasons of the same year becomes increasingly significant. In the 15th year, the reliability fluctuation range of the gear within one year is 0.99 to 0.94. This is affected by the randomness of decreasing residual gear strength. As time passes, the range of residual gear strength increases, resulting in a greater range of changes in gear reliability throughout the year.

5. Conclusions

In this study, from the perspective of fatigue damage accumulation, a residual strength model is established that considers the interaction of multistage loading and the influence of loading sequence. The load characteristics of wind turbine gears are focused, and the randomness and regularity of load level fluctuations throughout the four seasons are considered. Furthermore, the residual strength and dynamic reliability of wind turbine gears under multistage random load levels are calculated. Overall, the main conclusions can be summarized as follows:
(1)
The proposed residual strength model is validated using experimental data from two materials, and the calculation results showed that the proposed model has good predictive performance. In addition, by comparing its predictive performance with the classical Schaff’s model, it can be concluded that the proposed model can more accurately predict the residual strength of wind turbine gears.
(2)
The residual strength of the gear at a determined time point is a random variable that follows the Weibull distribution and is affected by the randomness of the load. The overall residual strength of the gear shows a decreasing trend. Moreover, the randomness of the residual strength of the gear at a certain time becomes stronger, and the range of residual strength values becomes wider as the service time increases.
(3)
Wind load amplitudes in each season influence the reliability of wind turbine gears. Initially, there is a small difference in reliability between seasons, but as the service time increases, the difference becomes more significant. Overall, the reliability of gears decreases monotonically over time in each season. However, there is a jump in reliability due to the effect of wind speed on the gears in different seasons.

Author Contributions

Methodology, software, investigation, writing—review and editing, funding acquisition, J.G.; formal analysis, software, original draft preparation, Y.L.; investigation, writing—review and editing, supervision, Y.Y.; data curation, writing—original draft preparation, F.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 52065062), Fundamental Research Funds for Universities in Xinjiang Uygur Autonomous Region (Grant No. XJEDU2023P007), the Natural Science Foundation of Xinjiang Uygur Autonomous Region (Grant No. 2020D01C056), and the Key Research and Development Program of Xinjiang Uygur Autonomous Region (Grant No. 2021B01003).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pérez, J.M.P.; Márquez, F.P.G.; Tobias, A.; Papaelias, M. Wind turbine reliability analysis. Renew. Sustain. Energy Rev. 2013, 23, 463–472. [Google Scholar] [CrossRef]
  2. Kusiak, A.; Li, W.Y. The prediction and diagnosis of wind turbine faults. Renew. Energy 2011, 36, 16–23. [Google Scholar] [CrossRef]
  3. Chen, S.; Qiao, Z.J.; Xie, C.Y.; Zhu, R.H.; Lai, Z.H.; Xu, X.F.; Zhou, S.T. Transfer condition monitoring across wind turbines using feature alignment and parameter fine-tuning. Meas. Sci. Technol. 2023, 34, 125901. [Google Scholar] [CrossRef]
  4. Dui, H.; Wang, X.; Zhou, H. Redundancy-based resilience optimization of multi-component systems. Mathematics 2023, 11, 3151. [Google Scholar] [CrossRef]
  5. Su, H.; Zhao, Y.; Wang, X. Analysis of a state degradation model and preventive maintenance strategies for wind turbine generators based on stochastic differential equations. Mathematics 2023, 11, 2608. [Google Scholar] [CrossRef]
  6. Costa, Á.M.; Orosa, J.A.; Vergara, D.; Fernández-Arias, P. New tendencies in wind energy operation and maintenance. Appl. Sci. 2021, 11, 1386. [Google Scholar] [CrossRef]
  7. Meng, D.B.; Yang, S.Y.; De Jesus, A.M.P.; Zhu, S.P. A novel Kriging-model-assisted reliability-based multidisciplinary design optimization strategy and its application in the offshore wind turbine tower. Renew. Energy 2023, 203, 407–420. [Google Scholar] [CrossRef]
  8. Meng, D.B.; Yang, S.Y.; Lin, T.; Wang, J.P.; Yang, H.F.; Lv, Z.Y. RBMDO using Gaussian mixture model-based second-order mean-value saddlepoint approximation. CMES-Comp. Model. Eng. Sci. 2022, 132, 553–568. [Google Scholar] [CrossRef]
  9. Meng, D.B.; Yang, S.Y.; He, C.; Wang, H.T.; Lv, Z.Y.; Guo, Y.P.; Nie, P. Multidisciplinary design optimization of engineering systems under uncertainty: A review. Int. J. Struct. Integr. 2022, 13, 565–593. [Google Scholar] [CrossRef]
  10. Mseddi, A.; Naifar, O.; Rhaima, M.; Mchiri, L.; Makhlouf, A.B. Robust Control for Torque Minimization in Wind Hybrid Generators: An H Approach. Mathematics 2023, 11, 3557. [Google Scholar] [CrossRef]
  11. Yan, Y. Load characteristic analysis and fatigue reliability prediction of wind turbine gear transmission system. Int. J. Fatigue 2020, 130, 105259. [Google Scholar] [CrossRef]
  12. Schaff, J.R.; Davidson, B.D. Life prediction methodology for composite structures. Part I—Constant amplitude and two-stress level fatigue. J. Compos. Mater. 1997, 31, 128–157. [Google Scholar] [CrossRef]
  13. Hahn, H.T.; Kim, R.Y. Proof testing of composite materials. J. Compos. Mater. 1975, 9, 297–311. [Google Scholar] [CrossRef]
  14. Chou, P.C.; Croman, R. Residual strength in fatigue based on the strength-life equal rank assumption. J. Compos. Mater. 1978, 12, 177–194. [Google Scholar] [CrossRef]
  15. Ganesan, C.; Joanna, P.S. Fatigue life and residual strength prediction of GFRP composites: An experimental and theoretical approach. Lat. Am. J. Solids Struct. 2018, 15, e72. [Google Scholar] [CrossRef]
  16. Cheng, H.C.; Hwu, F.S. Fatigue reliability analysis of composites based on residual strength. Adv. Compos. Mater. 2006, 15, 385–402. [Google Scholar] [CrossRef]
  17. Li, L.; Xie, L.Y.; He, X.H.; Hao, G.B. Strength degradation law of metallic material under fatigue loading. Chin. J. Mech. Eng. 2010, 32, 967–971. [Google Scholar]
  18. Gao, J.; An, Z. A new probability model of residual strength of material based on interference theory. Int. J. Fatigue 2019, 118, 202–208. [Google Scholar] [CrossRef]
  19. Jiang, C.J.; Liu, X.T.; Wang, X.L.; Wang, X.; Su, S.C. Interval dynamic reliability analysis of mechanical components under multistage load based on strength degradation. Qual. Reliab. Eng. Int. 2021, 37, 567–582. [Google Scholar] [CrossRef]
  20. Yang, B.; Fu, K.K.; Lee, J.; Li, Y. Artificial neural network (ANN)-based residual strength prediction of carbon fibre reinforced composites (CFRCs) after impact. Appl. Compos. Mater. 2021, 28, 809–833. [Google Scholar] [CrossRef]
  21. Lu, H.F.; Xu, Z.D.; Iseley, T.; Matthews, J.C. Novel data-driven framework for predicting residual strength of corroded pipelines. J. Pipeline Syst. Eng. Pract. 2021, 12, 04021045. [Google Scholar] [CrossRef]
  22. Miao, X.; Zhao, H. Novel method for residual strength prediction of defective pipelines based on HTLBO-DELM model. Reliab. Eng. Syst. Saf. 2023, 237, 109369. [Google Scholar] [CrossRef]
  23. Meng, D.B.; Yang, S.Y.; De Jesus, A.M.P.; Ferradosaet, T.F.; Zhu, S.P. A novel hybrid adaptive Kriging and water cycle algorithm for reliability-based design and optimization strategy: Application in offshore wind turbine monopile. Comput. Methods Appl. Mech. Eng. 2023, 412, 116083. [Google Scholar] [CrossRef]
  24. Yu, L.J.; Zhong, S.Y.; Bian, X.D.; Heilman, W.E. Climatology and trend of wind power resources in China and its surrounding regions: A revisit using Climate Forecast System Reanalysis data. Int. J. Climatol. 2016, 36, 2173–2188. [Google Scholar] [CrossRef]
  25. Whitworth, H.A. Evaluation of the residual strength degradation in composite laminates under fatigue loading. Compos. Struct. 2000, 48, 261–264. [Google Scholar] [CrossRef]
  26. Miner, M.A. Cumulative damage in fatigue. J. Appl. Mech. 1945, 12, A159–A164. [Google Scholar] [CrossRef]
  27. Manson, S.S.; Halford, G.R. Practical implementation of the double linear damage rule and damage curve approach for treating cumulative fatigue damage. Int. J. Fract. 1981, 17, 169–192. [Google Scholar] [CrossRef]
  28. Freudenthal, A.M.; Heller, R.A. On stress interaction in fatigue and a cumulative damage rule. J. Aerosp. Sci. 1959, 26, 431–442. [Google Scholar] [CrossRef]
  29. Zhu, S.P.; Huang, H.Z. A generalized frequency separation–strain energy damage function model for low cycle fatigue–creep life prediction. Fatigue Fract. Eng. Mater. Struct. 2010, 33, 227–237. [Google Scholar] [CrossRef]
  30. Zhu, S.P.; Huang, H.Z.; Ontiveros, V.; He, L.P.; Modarres, M. Probabilistic low cycle fatigue life prediction using an energy-based damage parameter and accounting for model uncertainty. Int. J. Damage Mech. 2012, 21, 1128–1153. [Google Scholar] [CrossRef]
  31. Risitano, A.; Risitano, G. Cumulative damage evaluation of steel using infrared thermography. Theor. Appl. Fract. Mech. 2010, 54, 82–90. [Google Scholar] [CrossRef]
  32. Peng, Z.C.; Huang, H.Z.; Wang, H.K.; Zhu, S.P.; Lv, Z.Q. A new approach to the investigation of load interaction effects and its application in residual fatigue life prediction. Int. J. Damage Mech. 2016, 25, 672–690. [Google Scholar] [CrossRef]
  33. Huang, B.X.; Wang, S.C.; Geng, S.L.; Liu, X.T. Improved numerical model for fatigue cumulative damage of mechanical structure considering load sequence and interaction. Adv. Mech. Eng. 2021, 13, 1687814021995309. [Google Scholar] [CrossRef]
  34. Gao, J.X.; Heng, F.; Yuan, Y.P.; Liu, Y.Y. Fatigue reliability analysis of composite material considering the growth of effective stress and critical stiffness. Aerospace 2023, 10, 785. [Google Scholar] [CrossRef]
  35. Ganesan, C.; Joanna, P.S. Modeling the residual strength and fatigue life of carbon fiber composites under constant amplitude loading. Mech. Adv. Mater. Struct. 2020, 27, 1840–1848. [Google Scholar] [CrossRef]
  36. Fu, Z.; Hu, X.; Zhang, J. A new fatigue life prediction method based on nonlinear fatigue cumulative damage generalized expression. J. Mech. Sci. Technol. 2022, 36, 205–212. [Google Scholar] [CrossRef]
  37. Osman, T.; Velex, P. A model for the simulation of the interactions between dynamic tooth loads and contact fatigue in spur gears. Tribol. Int. 2012, 46, 84–96. [Google Scholar] [CrossRef]
  38. Grbovic, A.; Rasuo, B. FEM based fatigue crack growth predictions for spar of light aircraft under variable amplitude loading. Eng. Fail. Anal. 2012, 26, 50–64. [Google Scholar] [CrossRef]
  39. Khan, M.A.A.; Sheikh, A.K.; Gasem, Z.M.; Asad, M. Fatigue life and reliability of steel castings through integrated simulations and experiments. Metals 2022, 12, 339. [Google Scholar] [CrossRef]
  40. Meng, D.B.; Lv, Z.Y.; Yang, S.Y.; Wang, H.T.; Xie, T.W.; Wang, Z.H. A time-varying mechanical structure reliability analysis method based on performance degradation. Structures 2021, 34, 3247–3256. [Google Scholar] [CrossRef]
  41. Ouammi, A.; Dagdougui, H.; Sacile, R.; Mimet, A. Monthly and seasonal assessment of wind energy characteristics at four monitored locations in Liguria region (Italy). Renew. Sustain. Energy Rev. 2010, 14, 1959–1968. [Google Scholar] [CrossRef]
  42. Wais, P. A review of Weibull functions in wind sector. Renew. Sustain. Energy Rev. 2017, 70, 1099–1107. [Google Scholar] [CrossRef]
  43. Altunkaynak, A.; Erdik, T.; Dabanlı, Í.; Şen, Z. Theoretical derivation of wind power probability distribution function and applications. Appl. Energy 2012, 92, 809–814. [Google Scholar] [CrossRef]
  44. Qian, B.Y.; Zhou, K.C.; Gao, X.X.; Liu, W. Torque transmission law of wind turbine drive system under the excitation of uneven wind load. J. Phys. Conf. Ser. 2022, 2218, 012057. [Google Scholar] [CrossRef]
  45. Gao, J.X. Reliability Analysis of Wind Turbine Gearbox Considering Strength Degradation and Failure Dependence. Master’s Thesis, Lanzhou University of Technology, Gansu, China, 2014. (In Chinese). [Google Scholar]
  46. Kebir, T.; Correia, J.; Benguediab, M.; Jesus, A.M.D. Numerical study of fatigue damage under random loading using Rainflow cycle counting. Int. J. Struct. Integr. 2021, 12, 149–162. [Google Scholar] [CrossRef]
  47. Jiang, S.; Li, W.; Xin, G.F.; Sheng, L.C.; Wang, Y.Q. Study on dynamic reliability of permanent magnet gear transmission system with wear and failure correlation. Eng. Fail. Anal. 2022, 131, 105802. [Google Scholar] [CrossRef]
  48. Nejad, A.R.; Gao, Z.; Moan, T. On long-term fatigue damage and reliability analysis of gears under wind loads in offshore wind turbine drivetrains. Int. J. Fatigue 2014, 61, 116–128. [Google Scholar] [CrossRef]
  49. Kroese, D.P.; Brereton, T.; Taimre, T.; Botev, Z.I. Why the Monte Carlo method is so important today. Wiley Interdiscip. Rev.-Comput. Stat. 2014, 6, 386–392. [Google Scholar] [CrossRef]
  50. Papadopoulos, C.E.; Yeung, H. Uncertainty estimation and Monte Carlo simulation method. Flow Meas. Instrum. 2001, 12, 291–298. [Google Scholar] [CrossRef]
  51. Nakazawa, M. Examining the Advantages and Disadvantages of Pilot Studies: Monte Carlo Simulations. Master’s Thesis, The University of New Mexico, Albuquerque, NM, USA, 2011. [Google Scholar]
  52. Huang, P.; Gu, Y.K.; Li, H.; Yazdi, M.; Qiu, G.Q. An optimal tolerance design approach of robot manipulators for positioning accuracy reliability. Reliab. Eng. Syst. Saf. 2023, 237, 109347. [Google Scholar] [CrossRef]
  53. Li, H.; Yazdi, M.; Huang, H.Z.; Huang, C.G.; Peng, W.; Nedjati, A.; Adesina, K.A. A fuzzy rough copula Bayesian network model for solving complex hospital service quality assessment. Complex Intell. Syst. 2023, 1–27. [Google Scholar] [CrossRef] [PubMed]
  54. Yazdi, M.; Mohammadpour, J.; Li, H.; Huang, H.Z.; Zarei, E.; Pirbalouti, R.G.; Adumene, S. Fault tree analysis improvements: A bibliometric analysis and literature review. Qual. Reliab. Eng. Int. 2023, 39, 1639–1659. [Google Scholar] [CrossRef]
  55. Yan, Y.T.; Luo, Z.Q.; Liu, Z.Y.; Liu, Z.B. Risk Assessment Analysis of Multiple Failure Modes Using the Fuzzy Rough FMECA Method: A Case of FACDG. Mathematics 2023, 11, 3459. [Google Scholar] [CrossRef]
  56. Yu, S.; Wang, Z.L. A general decoupling approach for time-and space-variant system reliability-based design optimization. Comput. Methods Appl. Mech. Eng. 2019, 357, 112608. [Google Scholar] [CrossRef]
  57. Yu, S.; Wang, Z.L.; Li, Y. Time and space-variant system reliability analysis through adaptive Kriging and weighted sampling. Mech. Syst. Signal Process. 2022, 166, 108443. [Google Scholar] [CrossRef]
  58. Yu, S.; Zhang, Y.W.; Li, Y.; Wang, Z.L. Time-variant reliability analysis via approximation of the first-crossing PDF. Struct. Multidiscip. Optim. 2020, 62, 2653–2667. [Google Scholar] [CrossRef]
  59. Yu, S.; Guo, P.; Wu, X. Adaptive first-crossing approach for life-cycle reliability analysis. Chin. J. Mech. Eng. 2023, 145, 091707. [Google Scholar] [CrossRef]
Figure 1. Residual strength prediction results of the proposed model and Schaff’s model comparison. (a) Residual strength of 45 steel under one-level constant amplitude load; (b) goodness of fit and the maximum error of the results predicted by the two models.
Figure 1. Residual strength prediction results of the proposed model and Schaff’s model comparison. (a) Residual strength of 45 steel under one-level constant amplitude load; (b) goodness of fit and the maximum error of the results predicted by the two models.
Mathematics 11 04013 g001
Figure 2. Residual strength of 30CrMnSiA steel with different loading levels. (a) Residual strength of 30CrMnSiA steel under lowhigh loading sequence; (b) residual strength of 30CrMnSiA steel under highlow loading sequence.
Figure 2. Residual strength of 30CrMnSiA steel with different loading levels. (a) Residual strength of 30CrMnSiA steel under lowhigh loading sequence; (b) residual strength of 30CrMnSiA steel under highlow loading sequence.
Mathematics 11 04013 g002
Figure 3. Residual strength of aluminum alloy Al-2024 with different loading levels. (a) Residual strength of aluminum alloy Al-2024 under lowhigh loading sequence; (b) residual strength of aluminum alloy Al-2024 under highlow loading sequence.
Figure 3. Residual strength of aluminum alloy Al-2024 with different loading levels. (a) Residual strength of aluminum alloy Al-2024 under lowhigh loading sequence; (b) residual strength of aluminum alloy Al-2024 under highlow loading sequence.
Mathematics 11 04013 g003
Figure 4. Maximum error and goodness of fit between model predictions and experimental values of 30CrMnSiA. (a) lowhigh loading sequence; (b) highlow loading sequence.
Figure 4. Maximum error and goodness of fit between model predictions and experimental values of 30CrMnSiA. (a) lowhigh loading sequence; (b) highlow loading sequence.
Mathematics 11 04013 g004
Figure 5. Maximum error and goodness of fit between model predictions and experimental values of aluminum alloy Al-2024. (a) lowhigh loading sequence; (b) highlow loading sequence.
Figure 5. Maximum error and goodness of fit between model predictions and experimental values of aluminum alloy Al-2024. (a) lowhigh loading sequence; (b) highlow loading sequence.
Mathematics 11 04013 g005
Figure 6. Simplified structure diagram of gear fatigue testing machine.
Figure 6. Simplified structure diagram of gear fatigue testing machine.
Mathematics 11 04013 g006
Figure 7. Results of the three-dimensional model establishment.
Figure 7. Results of the three-dimensional model establishment.
Mathematics 11 04013 g007
Figure 8. Equivalent stress cloud map of gear.
Figure 8. Equivalent stress cloud map of gear.
Mathematics 11 04013 g008
Figure 9. Life cloud map of gear.
Figure 9. Life cloud map of gear.
Mathematics 11 04013 g009
Figure 10. S-N curve of gear.
Figure 10. S-N curve of gear.
Mathematics 11 04013 g010
Figure 11. Root bending stress in 180 s in four seasons.
Figure 11. Root bending stress in 180 s in four seasons.
Mathematics 11 04013 g011
Figure 12. Flow chart of stress treatment.
Figure 12. Flow chart of stress treatment.
Mathematics 11 04013 g012
Figure 13. Resulting diagram of the rain flow counting method.
Figure 13. Resulting diagram of the rain flow counting method.
Mathematics 11 04013 g013
Figure 14. CDF fitting diagrams of equivalent stress for four seasons.
Figure 14. CDF fitting diagrams of equivalent stress for four seasons.
Mathematics 11 04013 g014
Figure 15. Flowchart for calculating residual strength.
Figure 15. Flowchart for calculating residual strength.
Mathematics 11 04013 g015
Figure 16. Residual strength path of gear with randomness. (a) Three-dimensional diagram of residual strength; (b) two-dimensional diagram of residual strength.
Figure 16. Residual strength path of gear with randomness. (a) Three-dimensional diagram of residual strength; (b) two-dimensional diagram of residual strength.
Mathematics 11 04013 g016
Figure 17. Weibull distribution test of spring residual strength. (a) Q-Q diagram of distribution test; (b) test of cumulative distribution function.
Figure 17. Weibull distribution test of spring residual strength. (a) Q-Q diagram of distribution test; (b) test of cumulative distribution function.
Mathematics 11 04013 g017
Figure 18. Curve fitting of shape and scale parameters over time. (a) Fitting result of scale parameters; (b) fitting result of shape parameters.
Figure 18. Curve fitting of shape and scale parameters over time. (a) Fitting result of scale parameters; (b) fitting result of shape parameters.
Mathematics 11 04013 g018
Figure 19. Reliability variation result of gear.
Figure 19. Reliability variation result of gear.
Mathematics 11 04013 g019
Figure 20. Failure rate variation result of gear.
Figure 20. Failure rate variation result of gear.
Mathematics 11 04013 g020
Figure 21. Explanation diagram of reliability jump.
Figure 21. Explanation diagram of reliability jump.
Mathematics 11 04013 g021
Table 1. Test results and model prediction of residual strength of 45 steel.
Table 1. Test results and model prediction of residual strength of 45 steel.
Stress/MPaCycle RatioCyclesExperimental Residual Strength/MPaSchaff’s Model/MPaProposed Model/MPa
310.37001183.261183.261183.26
0.269701180.941183.221164.65
0.413,9401179.381180.491140.65
0.620,9701067.81147.351106.49
0.827,900998.20967.761048.86
134,866310.37310.37310.37
Table 2. Experimental data of residual life of 30CrMnSiA under two-level loading.
Table 2. Experimental data of residual life of 30CrMnSiA under two-level loading.
Stress/MPaLoading Sequence   n 1   n 1 N f 1   n 2   n 2 N f 2
732836LowHigh13,0000.23366020.917
15,0000.26965010.903
25,0000.44854000.750
35,0000.62844280.615
45,0000.80732540.425
836732HighLow12000.16736,9110.833
18000.20832,4500.792
30000.41716,0020.583
50000.69469690.306
Table 3. Experimental data of residual life of aluminum alloy Al-2024 under two-level loading.
Table 3. Experimental data of residual life of aluminum alloy Al-2024 under two-level loading.
Stress/MPaLoading Sequence   n 1   n 1 N f 1   n 2   n 2 N f 2
150200LowHigh86,0000.2144,5000.9633
172,0000.4133,5000.8900
258,0000.681,7000.5447
200150HighLow30,0000.228,7000.5319
60,0000.4101,0500.2350
90,0000.676,0500.1769
Table 4. List of wind speed information.
Table 4. List of wind speed information.
SeasonAverage of Wind Speed/m·s−1Variance of Wind Speed/m2·s−2Scale Parameter  g Shape Parameter  f
Spring5.48 185.95 1.32
Summer6.92 11.67.70 1.56
Autumn11.81 26.5413.28 1.78
Winter8.63 19.129.57 1.51
Table 5. Relevant parameters for calculating bending stress of tooth root.
Table 5. Relevant parameters for calculating bending stress of tooth root.
ParametersMean ValueStandard ErrorParametersMean ValueStandard Error
  K A 1.10.036   Y F 2.210.073
  K V 1.150.035   Y F 1.640.066
  K β 1.320.02   Y ε 0.780
  K α 1.010.003   Y β 10
Table 6. Weibull distribution fitting results of equivalent stress in four seasons.
Table 6. Weibull distribution fitting results of equivalent stress in four seasons.
SeasonScale Parameter  h  of Weibull DistributionShape Parameter  j  of Weibull Distribution
Spring1.28550.97
Summer1.27600.17
Autumn1.32704.99
Winter1.25679.85
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

Gao, J.; Liu, Y.; Yuan, Y.; Heng, F. Residual Strength Modeling and Reliability Analysis of Wind Turbine Gear under Different Random Loadings. Mathematics 2023, 11, 4013. https://0-doi-org.brum.beds.ac.uk/10.3390/math11184013

AMA Style

Gao J, Liu Y, Yuan Y, Heng F. Residual Strength Modeling and Reliability Analysis of Wind Turbine Gear under Different Random Loadings. Mathematics. 2023; 11(18):4013. https://0-doi-org.brum.beds.ac.uk/10.3390/math11184013

Chicago/Turabian Style

Gao, Jianxiong, Yuanyuan Liu, Yiping Yuan, and Fei Heng. 2023. "Residual Strength Modeling and Reliability Analysis of Wind Turbine Gear under Different Random Loadings" Mathematics 11, no. 18: 4013. https://0-doi-org.brum.beds.ac.uk/10.3390/math11184013

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