Next Article in Journal
Sustainable Public Transport Strategies—Decomposition of the Bus Fleet and Its Influence on the Decrease in Greenhouse Gas Emissions
Next Article in Special Issue
Techno Economic Analysis of Electric Vehicle Grid Integration Aimed to Provide Network Flexibility Services in Italian Regulatory Framework
Previous Article in Journal
Calorimetric Evaluation of Thermal Stability of Organic Liquid Hydrogen Storage Materials and Metal Oxide Additives
Previous Article in Special Issue
Multi-Channel Profile Based Artificial Neural Network Approach for Remaining Useful Life Prediction of Electric Vehicle Lithium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Online Prediction of Remaining Useful Life for Li-Ion Batteries Based on Discharge Voltage Data

1
School of Mechanical and Electrical Engineering, Wuhan University of Technology, Wuhan 430070, China
2
Research Center for Intelligent Transportation Systems, Wuhan University of Technology, Wuhan 430070, China
*
Author to whom correspondence should be addressed.
Energies 2022, 15(6), 2237; https://doi.org/10.3390/en15062237
Submission received: 18 February 2022 / Revised: 15 March 2022 / Accepted: 17 March 2022 / Published: 18 March 2022
(This article belongs to the Special Issue Control and Management of Electric Power System in Vehicles)

Abstract

:
The state of health and remaining useful life of lithium-ion batteries are key indicators for the normal operation of electrical devices. To address the problem of the capacity of lithium-ion batteries being difficult to measure online, in this paper, we propose an online method based on particle swarm optimization and support vector regression to estimation the state of health and remaining useful life. First, a novel health indicator is extracted from the discharge voltage to characterize the capacity of lithium-ion batteries. Then, based on the capacity degradation characteristics, support vector regression is used to predict the remaining useful life of these batteries, and particle swarm optimization is selected to optimize the parameters of the support vector regression, which effectively enhances the predictive performance of the model. Validated for the NASA battery aging dataset, when training with the first 40% of the dataset, the maximum error of the predicted remaining useful life was four cycles, and when training with the first 50% of the dataset, the maximum error of the predicted remaining useful life was only one cycle. When comparing to a deep neural network, support vector regression, long short-term memory algorithms and existing similar methods in the literature, the particle swarm optimization and support vector regression method can obtain more accurate prediction results.

1. Introduction

With environmental degradation, energy crises, and other problems becoming more prominent, traditional fuel vehicles no longer meet the requirements of energy saving and environmental protection, and emerging electric vehicles are becoming the main transportation of the future. Lithium-ion (Li-ion) batteries have become the main energy source for electric vehicles because of their high energy density, long service life and pollution-free characteristics [1,2,3,4,5,6,7]. The state of health (SOH) and remaining useful life (RUL) are two important monitoring parameters in a battery management system [8]. There are many definitions of SOH, but there is no unified definition yet. From a capacity perspective, this paper defines SOH as the ratio of the battery’s maximum capacity in the current cycle to its initial rated capacity; when the ratio drops to 0.7~0.8, the battery is considered to reach its end of life (EoL) [9], and continued use may lead to serious safety accidents. In actual use, the accurate prediction of SOH and RUL is conducive to the active maintenance of the battery, for avoiding safety accidents, and for providing a guarantee for stable battery operation [10].
In recent years, there has been a lot of research on predicting the SOH and RUL of Li-ion batteries. According to the different prediction methods, these research works can be divided into three categories: statistical filtering methods, data-driven methods, and model fusion methods [11].
Statistical filtering-based methods are mainly mathematical models constructed by the data characteristics of Li-ion batteries, including the Kalman filter and its extension methods and the particle filter and its extension methods. Fang et al. [12] proposed a joint state of charge (SOC) and SOH estimation method based on the dual extended Kalman filter algorithm, in turn based on the relationship between ohmic internal resistance and SOH. Zeng et al. [13] established an improved second-order equivalent circuit model, and used the Bayesian method for online parameter identification, took the ohmic resistance as the characteristic parameter of the SOH, and adopted the fuzzy unscented Kalman filter algorithm for the joint estimation of the SOC and SOH. Hu et al. [14] considered the problem of particle impoverishment in the particle filter and combined kernel smoothing with the particle filter to achieve RUL prediction with a higher accuracy than the conventional particle filter. Zhang et al. [15] introduced the Markov chain Monte Carlo method to solve the sample impoverishment problem of the unscented particle filter, in order to achieve the prediction of RUL. This type of method is easy to implement and has the advantage of a high accuracy, but it is easily affected by external factors.
Model fusion-based methods mainly combine statistical filtering and data-driven methods, which overcome the limitations of a single method. Wei et al. [16] developed a battery SOH state space model based on support vector regression (SVR) and introduced the particle filter to optimize the impedance attenuation parameters to suppress measurement noise and achieve an accurate prediction of the SOH and RUL. Song et al. [17] proposed a joint state estimation method for Li-ion batteries based on data-driven least squares support vector machines and the model-based unscented particle filter. Chang et al. [18] proposed a method for RUL prediction based on the idea of error correction by combining the unscented Kalman filter, complete ensemble empirical mode decomposition, and the relevance vector machine. However, this type of method currently has difficulties in overhead modeling and computing.
In recent years, data-driven research methods, which do not require much expertise in the field of electrochemistry and only need the historical data of battery operation to achieve SOH and RUL predictions, have received a lot of attention from scholars. The main data-driven methods include neural networks [19,20,21,22], support vector machines [23,24], moving average models [25,26], Gaussian process regression [4,27], and ensemble learning [28]. Wang et al. [29] used the ant lion optimization algorithm to optimize the parameters of SVR to achieve an accurate prediction of the SOH and RUL. Phattara et al. [30] used a deep neural network (DNN) to predict the SOH and RUL of Li-ion batteries and compared it to other algorithms to highlight the superiority of DNNs. Li et al. [31] used a recurrent neural network with a gated recurrent unit and a convolutional neural network, using voltage, current, and temperature to achieve the simultaneous estimation of the SOC and SOH. Qu et al. [32] considered the existence of noise in the Li-ion battery capacity sequence, and first performed the empirical modal decomposition of the sequence, then built a long short-term memory (LSTM) network based on particle swarm optimization (PSO) for the prediction of the SOH and RUL, supplemented with an incremental learning and attention mechanism to improve the performance of the model. Similarly, Hu et al. [33] divided the decomposed components into principal component data and fluctuation data according to correlation analysis, and used a deep belief network and an LSTM network to predict them, respectively, which also obtained accurate RUL prediction results. Yang et al. [34] used a hybrid pulse power characterization test for direct parameter extraction to identify the parameters of the equivalent circuit model, and the extracted parameters were used to train the back propagation neural network. Finally, the accuracy of back propagation neural network prediction was verified by static and dynamic current profiling tests.
In addition to the research on algorithms, there have been many studies on health indicators (HIs) that characterize the SOH of Li-ion batteries in recent years [35]. There are two main types of HIs—direct and indirect health indicators. Capacity and internal resistance are usually used as direct health indicators to characterize the SOH of Li-ion batteries, but they typically need to be obtained offline by intrusive methods under laboratory conditions, making them subject to many restrictions in practical application scenarios. Indirect health indicators need to be able to indirectly characterize the SOH of Li-ion batteries and can be measured online [36]. In [37], the discharging voltage difference of equal time intervals was chosen as an indirect health indicator to characterize the SOH, while in [38,39], the time interval of equal discharging voltage difference was chosen as an indirect health indicator to characterize the SOH. Zhou et al. [40] extracted the mean voltage fall from the discharge process of the battery as an indirect health indicator, and Yang et al. [41] chose to use discharge temperature change rate as an indirect health indicator, and Gao Dong et al. [35] extracted the average charge current drop from the charging process as an indirect health indicator considering the instability of the discharging conditions. These indirect HIs make it more convenient to predict the SOH online.
Addressing the problem that the SOH and RUL of Li-ion batteries are difficult to predict online, a new indirect HI was extracted from the discharge voltage as the input of the SVR. In order to improve the prediction accuracy, the PSO was introduced to optimize the SVR parameters, and finally, to achieve the accurate online prediction of the SOH and RUL.
The remainder of this paper is arranged as follows: Section 2 introduces the relevant algorithms and the PSO–SVR algorithm; Section 3 focuses on the Li-ion battery dataset and the extraction of indirect HIs in this paper; Section 4 introduces the relevant model evaluation metrics and shows the prediction results of the PSO–SVR model using the dataset; finally, relevant conclusions are drawn in Section 5.

2. Related Algorithms

2.1. SVR

Support vector regression [42], as a branch of support vector machines, is often used to solve small-sample, nonlinear regression problems. For nonlinear regression problems, assume that the given sample set is D = { ( x 1 , y 1 ) , ( x 2 , y 2 ) , , ( x n , y n ) } , ( x i R n , y i R ) , where x i is the feature vector of sample i ; correspondingly, y i is the target value of sample i , and n denotes the total number of samples. It is desirable to obtain a regression model, f ( x ) = w φ ( x ) + b , such that the difference between f ( x ) and y is as small as possible, where w , b are the parameters to be determined, w denotes the weight, x denotes the input data, φ is the mapping function, and b denotes the intercept.
Thus, the SVR problem can be transformed as follows:
m i n L ( w , b ) = 1 2 w 2 + C i = 1 n l ( f ( x i ) y i )
where C denotes the penalty factor and l denotes the loss function.
Introducing the slack variables, ξ i and ξ ^ i , the SVR problem can be transformed into Equations (2) and (3):
m i n L ( w , b , ξ i , ξ ^ i ) = 1 2 w 2 + C i = 1 n ( ξ i + ξ ^ i )
s t { f ( x i ) y i ε + ξ i y i f ( x i ) ε + ξ ^ i ξ i 0 , ξ ^ i 0 , i = 1 , 2 , , n
where ε denotes the maximum deviation that the regression model can tolerate between f ( x ) and the target value y . Next, Lagrange multipliers are introduced, and then partial derivatives of w , b , ξ i , ξ ^ i are taken to make them zero, respectively, which can be converted into Equations (4) and (5):
m a x ( α , α ^ ) = i = 1 n y i ( α ^ i α i ) ε ( α ^ i + α i ) 1 2 i = 1 n j = 1 n ( α ^ i α i ) ( α ^ j α j ) φ ( x i ) T φ ( x j )
s t { i = 1 n ( α ^ i α i ) = 0 0 α i , α ^ i C
where α i and α ^ i are the introduced Lagrangian multipliers; therefore, we can obtain:
w = i = 1 n ( α ^ i α i ) φ ( x i )
That is, the regression model can be expressed as Equation (7):
f ( x ) = i = 1 n ( α ^ i α i ) K ( x i , x j ) + b
where K ( x i , x j ) = φ ( x i ) T ( x j ) is the kernel function, i.e., the selection and construction of the kernel function needed to satisfy Mercer’s theorem. In this paper, the radial basis function (RBF) was selected, and its expression is as per Equation (8):
K ( x i , x j ) = exp ( 1 2 σ 2 x i x j 2 )
where σ is the parameter of kernel function, which denotes the width coefficient of the RBF.

2.2. Particle Swarm Optimization

The PSO [43] is a global random search algorithm that simulates the foraging behavior of birds. Assuming that particles fly in a D-dimensional solution set space, each particle has two attributes: position and velocity. The position of particle i at the current time t is X i t = [ x i 1 t , x i 2 t , , x i D t ] , which represents a solution in the D-dimensional solution set space, and the velocity V i t = [ v i 1 t , v i 2 t , , v i D t ] represents the direction and distance of the next flight of the particle; P b e s t i = [ p i 1 , p i 2 , , p i D ] T is the optimal position experienced by particle i , and G b e s t = [ g 1 , g 2 , , g D ] T is the optimal position experienced by the particle population. Each particle updates the individual optimal position and the global optimal position according to the velocity update (Equation (9)) and the position update (Equation (10)):
V i t + 1 = ω V i t + c 1 r 1 ( P b e s t i X i t ) + c 2 r 2 ( G b e s t X i t )
X i t + 1 = X i t + V i t + 1
where ω is the inertia constant, which is usually in the range of 0.8–1.2; c 1 and c 2 are learning factors, which adjust the step size of the particle flying in the direction of the individual optimal position, as well as the direction of the global optimal position; and r 1 and r 2 are constants, which are usually random numbers between 0 and 1.
Usually, in experimental studies, when ω takes a large value, the algorithm converges with difficulty, and when ω takes a small value, the algorithm tends to fall into local extremum [44]. Thus, in this paper, adaptive ω values were adopted so that ω decreased linearly; ω takes a larger value in the early stage of the algorithm to improve the search ability and a smaller value in the later stage of the algorithm to ensure that the algorithm can search carefully around the optimal extremum, converging to the global optimal extremum. At the same time, in order to avoid the algorithm falling into local optimization and to increase the diversity of particles, we added adaptive mutation to the PSO; after the fitness of each generation of particles is calculated, the particles with low fitness value undergo a random mutation within the specified range.

2.3. The PSO–SVR Method

More and more optimization algorithms are used for parameter identification, including particle swarm optimization [32], genetic algorithm [45], bird mating optimization algorithm [46], grey wolf optimizer [47], atom search optimizer [48], selective hybrid stochastic strategy [49] and so on. In the SVR algorithm, the penalty factor C, as well as the kernel parameter γ, directly determine the accuracy of the algorithm. At present, there is no clear guideline scheme for the selection of parameters; thus, the PSO was selected here to find the appropriate parameters of the SVR.
The specific execution flow of PSO–SVR is shown in Figure 1, which mainly contains the following steps:
Step 1. Indirect health indicator extraction and data processing: First, the battery discharge voltage needs to be processed to extract the indirect HI characterizing the battery capacity degradation. Then, the extracted indirect HI of the kth cycle and the number of cycles k are used as the input features of the model to predict the capacity value of cycle k + 1, and the data are divided into a training set and a testing set.
Step 2. Parameter initialization of the PSO algorithm: In PSO, the number of particles is 40, the number of iterations is 200, the maximum value of the penalty factor C, the parameter γ is 200, and the minimum value is 0.0001.
Step 3. Set the fitness function: In this paper, the mean square error (MSE) was selected as the fitness function to evaluate the accuracy of the model prediction:
M S E = 1 n i = 1 n ( y i y ^ i ) 2
where y i and y ^ i denote the actual capacity and the predicted capacity of cycle i , respectively, and n denotes the number of predicted samples.
Step 4. The PSO algorithm performs a parameter optimization: (1) The initialized parameters of each particle are introduced into the SVR model for training and prediction, and the prediction results are introduced into Equation (11) to calculate the fitness value of each particle. (2) The global optimal position and individual optimal position of the particles are recorded. (3) Then, the position of each particle is updated according to Equations (9) and (10). (4) Repeat (1)–(3) until the cutoff condition is reached. Step 5. SOH and RUL prediction: The global optimal parameters optimized by the PSO algorithm are introduced into the SVR model to predict the SOH and RUL of Li-ion batteries.

3. Extraction of an Indirect Health Indicator

3.1. Dataset

The charge and discharge data of the B0005 (B5), B0006 (B6), B0007 (B7), and B0018 (B18) batteries of the NASA Prognostics Center of Excellence were selected for this study [50]. The batteries all had an initial capacity of 2Ah and were specified to reach EoL when the capacity decayed to 70% of the initial capacity during the charge/discharge cycle.
Each battery was experimented on with the three different modes of charging, discharging, and impedance. As shown in Table 1, the charging process started with a constant current charge of 1.5 A, and when the voltage reached 4.2 V, a constant voltage charge was carried out until the current dropped to 20 mA. In the discharge mode, a constant current discharge was carried out at a current of 2 A until the voltages dropped to their respective cutoff voltages.
Figure 2 shows the change process of the current and voltage in the first cycle of the B5 battery. Figure 3 shows the capacity degradation trend and the battery failure threshold. Overall, the battery capacity decreased as the number of cycles increased, but the capacity regeneration phenomenon after battery resting led to a local rebound of the battery capacity. It can also be seen that the B7 battery did not reach the failure threshold during the whole life cycle; therefore, this paper took 72% of the initial capacity as the failure threshold for the B7 battery to reach the EoL.

3.2. Extraction of a Novel Indirect Health Indicator

Figure 4 shows the change of the discharge voltage with time under different discharge cycles. Starting with the constant current discharge stage, the voltage gradually dropped from 4.2 V to the cutoff voltage, followed by a regeneration of the battery capacity, causing the discharge voltage to rise back up. From the figure, we can see that as the battery aged, it took a shorter time for the voltage to drop to the cutoff voltage. Therefore, from the start of discharge to the cutoff voltage, we took the integral of the voltage over time (IVT) as a new HI, as shown in the pink area of the figure below, to characterize the energy that the battery could release. We found that the area decreased as the number of cycles increased, and the pink area can be expressed as Equation (12):
S i = t 0 t c u t o f f u i d t
where S i is the area at cycle i , t 0 is the time to start discharging, t c u t o f f represents the time taken to reach the cutoff voltage, and u i is the real-time value of the discharging voltage in cycle i .
The extraction process of IVT has the following three main steps.
Step 1. Extraction of the discharge voltage and time data for each discharge cycle;
Step 2. Performing a polynomial fit to the discharge voltage and time data;
Step 3. Calculating the IVT according to Equation (12).
Based on the above extraction process, the IVTs of the above four batteries were extracted, respectively, which is shown in Figure 5.

3.3. Pearson’s Correlation Analysis

Pearson’s correlation analysis is commonly used to measure the degree of correlation between two sets of variables. Assuming that the two sets of variables are U = { U 1 , U 2 , U 3 , , U n } and V = { V 1 , V 2 , V 3 , , V n } , their Pearson’s correlation coefficients can be expressed as Equation (13):
r = i = 1 n ( U i U ¯ ) ( V i V ¯ ) i = 1 n ( U i U ¯ ) 2 i = 1 n ( V i V ¯ ) 2
where the value range of r is −1 to 1. When r = ± 1 , a strict linear correlation between the two sets of variables is indicated; when r = 0 , a linear independence between the two sets of variables is suggested; and when r is within (−1,0) or (0,1), it means that there is a correlation between the two sets of variables, and the correlation is proportional to the absolute value of r .
Table 2 shows the Pearson’s correlation coefficients between the capacity and IVTs of the four batteries, and the absolute values are above 0.999. This shows the validity of the IVTs; that is, IVTs can be used as the input features of the SVR to achieve the online prediction of the SOH and RUL of Li-ion batteries.

4. Battery SOH and RUL Prediction

4.1. Evaluation Criteria

To evaluate the effectiveness of the proposed prediction method, the mean absolute error (MAE), the root mean square error (RMSE), the R2 score, and the error were selected as the evaluation criteria:
MAE:
M A E = 1 n i = 1 n | y i y ^ i |
RMSE:
R M S E = 1 n i = 1 n ( y i y ^ i ) 2
R2:
R 2 = 1 i = 1 n ( y ^ i y i ) 2 i = 1 n ( y ¯ y i ) 2
Error:
E r r o r = | R U L ^ R U L |
where y i and y ^ i denote the actual capacity and the predicted capacity of cycle i , respectively, y ¯ denotes the average of the true capacity values, n denotes the total number of samples, and R U L ^ and R U L denote the predicted and true values of the RUL, respectively.

4.2. Prediction Results

This paper selected the first 40% and 50% of the sample data in the battery life cycle as the training set, and used the last 60% and 50% of the sample data as the testing set.
Table 3 shows the results of PSO for optimizing the SVR parameters. We compared the PSO–SVR method to DNN, LSTM (see Appendix A for more details), and SVR (C = 100, γ = 0.01) at the same prediction start point. Figure 6 and Figure 7 and Table 4 show the results of the PSO–SVR method compared to the different algorithms (where “-” indicates that the prediction capacity value is not reduced to the failure threshold). We can find that the prediction results of the PSO–SVR are significantly better than those of the other algorithms, except for the MAE of PSO–SVR, which is 0.0001 higher than that of SVR on the B18 battery when the first 40% of the dataset was used as the training set. In general, the prediction results of the model are essentially consistent with the decreasing trend in the actual capacity, at the same time, the model can also feedback the fluctuation data generated by capacity regeneration or other noise in time.
When the first 40% of the dataset is used as the training set, the maximum error, the maximum MAE and the maximum RMSE of PSO–SVR are 4, 0.0194 and 0.0269, respectively. Taking B5 as an example, the MAE of PSO–SVR is 0.0083, while the MAEs of the DNN, LSTM, and SVR are 0.1175, 0.0282, and 0.0312, respectively; similarly, the RMSE of PSO–SVR is 0.0137, while the RMSEs of DNN, LSTM, and SVR are 0.1313, 0.0330, and 0.0325, respectively. In terms of R2, the larger the R2 value, the better the prediction effect, and the R2 value of PSO–SVR is 0.9816, which is significantly larger than that of the other methods.
When the first 50% of the dataset is used as the training set, the maximum error, the maximum MAE and the maximum RMSE of PSO–SVR are 1, 0.0177 and 0.0230, respectively. Additionally, taking B5 as an example, the MAE of PSO–SVR is 0.0076, while the MAEs of DNN, LSTM and SVR are 0.0852, 0.0125, and 0.0157, respectively; similarly, the RMSE of PSO–SVR is 0.0133, while the RMSEs of the DNN, LSTM, and SVR are 0.0957, 0.0185, and 0.0183, respectively. In terms of R2, the R2 value of the PSO–SVR is 0.9726, which is also significantly larger than that of the other methods.
When predicting the B18 battery and using the first 40% of the dataset as the training set, and when predicting the B7 battery and using the first 50% of the dataset as the training set, the prediction results of PSO–SVR and SVR are almost the same, which shows that when we initialize the SVR parameters, we select the appropriate parameters.
Overall, the prediction results of PSO–SVR highly fitted the actual capacity values, which proves that the PSO–SVR method can realize the accurate online prediction of Li-ion batteries’ SOH and RUL and provide the necessary information for the maintenance and replacement of Li-ion batteries.
Figure 8 and Figure 9, respectively, show the absolute error (AE) and MAE between the true value and the predicted value obtained by different prediction algorithms when the first 40% and 50% of the dataset were used as the training set. We can find that the AE of PSO–SVR is significantly smaller than that of the DNN, LSTM, and SVR, and the AE of PSO–SVR almost remained within 0.02. In terms of the MAE, the PSO–SVR method retained it within 0.02, and the MAE of the PSO–SVR method is less than 0.01 for the B5 battery and when the first 40% of the B7 battery dataset was used as the training set.
In addition, a comparison was made to similar studies in the literature, in which the failure threshold was uniformly set to 1.44 Ah, and the results are shown in Table 5. We can find that only for the B5 battery, when the prediction start point was 68, the error value of PSO–SVR was one value larger than that of the method used in the literature [29]; meanwhile, when the prediction start point was 84, the error value of PSO–SVR was the same as that of the method in the literature [29]. For the B18 battery, when the prediction start point was 53, the error value of PSO–SVR was the same as the method in the literature [29]; meanwhile, when the prediction start point was 66, the error value of PSO–SVR was 0, which was better than that of the method in the literature [29]. For the B6 and B7 batteries, the error value of PSO–SVR was significantly smaller than that of the method in the literature [29]. When the prediction start point was 68, the error value of PSO–SVR for both was 0, and when the prediction start point was 84, the error value of PSO–SVR was 1/4 and 1/12 of that of the method in the literature [29], respectively.

5. Conclusions

For the problem of the SOH and RUL prediction of Li-ion batteries, the IVT based on discharge voltage and an online prediction method of PSO–SVR were proposed in this paper.
Firstly, the IVT used to characterize the capacity of the battery was extracted from the discharge voltage. Secondly, the SVR model was selected to predict the SOH and RUL for the characteristics of NASA battery aging data, and then the IVT and cycle number were taken as the input features of the model to achieve online prediction of the SOH and RUL, and the PSO–SVR model was constructed for the problem of parameter selection for the SVR, and the optimal parameters of the SVR were obtained by using the PSO algorithm. Finally, the prediction results showed the feasibility of this scheme and also demonstrated the superiority of this scheme when compared to other algorithms and similar studies in the literature.

Author Contributions

Conceptualization, B.W. and Y.W.; methodology, B.W.; software, B.W.; validation, B.W.; formal analysis, B.W.; resources, Y.W. and L.Z.; data curation, B.W.; writing—original draft preparation, B.W.; writing—review and editing, B.W., Y.W., Y.Z. and J.Y.; visualization, B.W.; supervision, Y.W. and L.Z.; project administration, Y.W.; funding acquisition, H.Z. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 11972268.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data of this paper came from the NASA Prognostics Center of Excellence, and the data acquisition website was: https://ti.arc.nasa.gov/tech/dash/groups/pcoe/prognostic-data-repository/publications/#battery (accessed on 23 March 2020).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

SOHState of health
RULRemaining useful life
EoLEnd of life
SOCState of charge
SVRSupport vector regression
DNNDeep neural network
LSTMLong short-term memory
PSOParticle swarm optimization
HIHealth indicator
IVTIntegral of the voltage over time

Appendix A

The details of the DNN and the LSTM networks used in this paper are shown in Table A1. In order to prevent overfitting, we have added the early stopping method to the DNN and the LSTM.
Table A1. The details of the DNN and LSTM networks.
Table A1. The details of the DNN and LSTM networks.
AlgorithmLayer TypeNodesActivationOthers
DNN Optimizer = Adam
Dense100ReluLoss = mean absolute error
Dense
Dropout
Dense
100
100 (rate = 0.25)
1
Relu
None
None
Batchsize = 16
Epoch = 200
Monitor = val_loss
Patience = 50
LSTMLSTM
Dropout
Dense
100
100 (rate = 0.3)
1
Tanh
None
None
Optimizer = Adam
Loss = mean absolute error
Batchsize = 16
Epoch = 200
Monitor = val_loss
Patience = 10

References

  1. Zhu, L.; Zhang, J. A New Model of Jointed States of Charge and Health for Lithium Batteries. Proc. CSEE 2018, 38, 3613–3621. [Google Scholar]
  2. Yang, J.; Du, C.; Liu, W.; Wang, T.; Yan, L.; Gao, Y.; Cheng, X.; Zuo, P.; Ma, Y.; Yin, G.; et al. State-of-health estimation for satellite batteries based on the actual operating parameters–Health indicator extraction from the discharge curves and state estimation. J. Energy Storage 2020, 31, 101490. [Google Scholar] [CrossRef]
  3. Dai, H.; Zhao, G.; Lin, M.; Wu, J.; Zheng, G. A novel estimation method for the state of health of Li-ion battery using prior knowledge-based neural network and Markov chain. IEEE Trans. Ind. Electron. 2018, 66, 7706–7716. [Google Scholar] [CrossRef]
  4. Yang, D.; Zhang, X.; Pan, R.; Wang, Y.; Chen, Z. A novel Gaussian process regression model for state-of-health estimation of Li-ion battery using charging curve. J. Power Sources 2018, 384, 387–395. [Google Scholar] [CrossRef]
  5. Lin, J.; Zeng, C.; Lin, X.; Xu, C.; Xu, X.; Luo, Y. Metal–organic framework-derived hierarchical MnO/Co with oxygen vacancies toward elevated-temperature Li-Ion battery. ACS Nano 2021, 15, 4594–4607. [Google Scholar] [CrossRef]
  6. Karuppiah, S.; Keller, C.; Kumar, P.; Jouneau, P.-H.; Aldakov, D.; Ducros, J.-B.; Lapertot, G.; Chenevier, P.; Haon, C. A scalable silicon nanowires-grown-on-graphite composite for high-energy lithium batteries. ACS Nano 2020, 14, 12006–12015. [Google Scholar] [CrossRef]
  7. Schmuch, R.; Wagner, R.; Hörpel, G.; Placke, T.; Winter, M. Performance and cost of materials for lithium-based rechargeable automotive batteries. Nat. Energy 2018, 3, 267–278. [Google Scholar] [CrossRef]
  8. Dong, G.; Chen, Z.; Wei, J.; Ling, Q. Battery health prognosis using Brownian motion modeling and particle filtering. IEEE Trans. Ind. Electron. 2018, 65, 8646–8655. [Google Scholar] [CrossRef]
  9. Si, X.S.; Wang, W.; Hu, C.H.; Zhou, D.H. Remaining useful life estimation—A review on the statistical data driven approaches. Eur. J. Oper. Res. 2011, 213, 1–14. [Google Scholar] [CrossRef]
  10. Duong, P.L.T.; Raghavan, N. Heuristic Kalman optimized particle filter for remaining useful life prediction of Li-ion battery. Microelectron. Reliab. 2018, 81, 232–243. [Google Scholar] [CrossRef]
  11. Liu, D.; Song, Y.; Wu, W.; Yang, C.; Peng, Y. Review of state of health estimation for Li-ion battery pack. Chin. J. Sci. Instrum. 2020, 41, 1–18. [Google Scholar]
  12. Fang, L.; Li, J.; Peng, B. Online Estimation and Error Analysis of both SOC and SOH of Li-ion Battery based on DEKF Method. Energy Procedia 2019, 158, 3008–3013. [Google Scholar] [CrossRef]
  13. Zeng, M.; Zhang, P.; Yang, Y.; Xie, C.; Shi, Y. SOC and SOH Joint Estimation of the Power Batteries based on Fuzzy Unscented Kalman Filtering Algorithm. Energies 2019, 12, 3122. [Google Scholar] [CrossRef] [Green Version]
  14. Hu, Y.; Baraldi, P.; Di Maio, F.; Zio, E. A particle filtering and kernel smoothing-based approach for new design component prognostics. Reliab. Eng. Syst. Saf. 2015, 134, 19–31. [Google Scholar] [CrossRef]
  15. Zhang, X.; Miao, Q.; Liu, Z. Remaining useful life prediction of Li-ion battery using an improved UPF method based on MCMC. Microelectron. Reliab. 2017, 75, 288–295. [Google Scholar] [CrossRef]
  16. Wei, J.; Dong, G.; Chen, Z. Remaining useful life prediction and state of health diagnosis for Li-ion batteries using particle filter and support vector regression. IEEE Trans. Ind. Electron. 2017, 65, 5634–5643. [Google Scholar] [CrossRef]
  17. Song, Y.; Liu, D.; Liao, H.; Peng, Y. A hybrid statistical data-driven method for on-line joint state estimation of Li-ion batteries. Appl. Energy 2020, 261, 114408. [Google Scholar] [CrossRef]
  18. Chang, Y.; Fang, H.; Zhang, Y. A new hybrid method for the prediction of the remaining useful life of a Li-ion battery. Appl. Energy 2017, 206, 1564–1578. [Google Scholar] [CrossRef]
  19. Lin, H.-T.; Liang, T.-J.; Chen, S.-M. Estimation of Battery State of Health Using Probabilistic Neural Network. IEEE Trans. Ind. Inform. 2013, 9, 679–685. [Google Scholar] [CrossRef]
  20. You, G.W.; Park, S.; Oh, D. Diagnosis of electric vehicle batteries using recurrent neural networks. IEEE Trans. Ind. Electron. 2017, 64, 4885–4893. [Google Scholar] [CrossRef]
  21. Zhang, Y.; Xiong, R.; He, H.; Pecht, M.G. Long short-term memory recurrent neural network for remaining useful life prediction of Li-ion batteries. IEEE Trans. Veh. Technol. 2018, 67, 5695–5705. [Google Scholar] [CrossRef]
  22. Li, X.; Zhang, L.; Wang, Z.; Dong, P. Remaining useful life prediction for Li-ion batteries based on a hybrid model combining the long short-term memory and Elman neural networks. J. Energy Storage 2019, 21, 510–518. [Google Scholar] [CrossRef]
  23. Patil, M.A.; Tagade, P.; Hariharan, K.S.; Kolake, S.M.; Song, T.; Yeo, T.; Doo, S. A novel multistage Support Vector Machine based approach for Li ion battery remaining useful life estimation. Appl. Energy 2015, 159, 285–297. [Google Scholar] [CrossRef]
  24. Wang, N.; Liu, Y.; Jiang, K.; Chen, Z. Life Prediction of Lithium Iron Phosphate Battery Based on VPSO-SVM. J. Chongqing Univ. Technol. (Nat. Sci.) 2019, 33, 173–177, 230. [Google Scholar]
  25. Zhou, Y.; Huang, M. Li-ion batteries remaining useful life prediction based on a mixture of empirical mode decomposition and ARIMA model. Microelectron. Reliab. 2016, 65, 265–273. [Google Scholar] [CrossRef]
  26. Chen, Z.; Xue, Q.; Xiao, R.; Liu, Y.; Shen, J. State of Health Estimation for Li-ion Batteries Based on Fusion of Autoregressive Moving Average Model and Elman Neural Network. IEEE Access 2019, 7, 102662–102678. [Google Scholar] [CrossRef]
  27. Wang, Z.; Ma, J.; Zhang, L. State-of-Health Estimation for Li-ion Batteries Based on the Multi-Island Genetic Algorithm and the Gaussian Process Regression. IEEE Access 2017, 5, 21286–21295. [Google Scholar] [CrossRef]
  28. Peng, J.; Zheng, Z.; Zhang, X.; Deng, K.; Gao, K.; Li, H.; Chen, B.; Yang, Y.; Huang, Z. A Data-Driven Method with Feature Enhancement and Adaptive Optimization for Lithium-ion Battery Remaining Useful Life Prediction. Energies 2020, 13, 752. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, Y.; Ni, Y.; Zheng, Y.; Shi, X.; Wang, J. Remaining Useful Life Prediction of Li-ion Batteries Based on Support Vector Regression Optimized and Ant Lion Optimizations. Proc. CSEE 2021, 41, 1445–1457, 1550. [Google Scholar]
  30. Khumprom, P.; Yodo, N. A Data-Driven Predictive Prognostic Model for Lithium-ion Batteries based on a Deep Learning Algorithm. Energies 2019, 12, 660. [Google Scholar] [CrossRef] [Green Version]
  31. Li, C.; Xiao, F.; Fan, Y.; Tang, X.; Yang, G. Joint Estimation of the State of Charge and the State of Health Based on Deep Learning for Li-ion Batteries. Proc. CSEE 2021, 41, 681–692. [Google Scholar]
  32. Qu, J.; Liu, F.; Ma, Y.; Fan, J. A Neural-Network-Based Method for RUL Prediction and SOH Monitoring of Li-ion Battery. IEEE Access 2019, 7, 87178–87191. [Google Scholar] [CrossRef]
  33. Hu, T.; Yu, J. Life prediction of Li-ion batteries based on multiscale decomposition and deep learning. J. Zhejiang Univ. (Eng. Sci.) 2019, 53, 1852–1864. [Google Scholar]
  34. Yang, D.; Wang, Y.; Pan, R.; Chen, R.; Chen, Z. A Neural Network Based State-of-Health Estimation of Li-ion Battery in Electric Vehicles. Energy Procedia 2017, 105, 2059–2064. [Google Scholar] [CrossRef]
  35. Gao, D.; Huang, M.; Zhou, Y. Li-ion battery’s capacity estimation based on charging current data. Chin. J. Power Sources 2018, 42, 1447–1450, 1581. [Google Scholar]
  36. Gao, D.; Huang, M.H.; Xie, J. A novel indirect health indicator extraction based on charging data for Li-ion batteries remaining useful life prognostics. SAE Int. J. Altern. Powertrains 2017, 6, 183–193. [Google Scholar] [CrossRef]
  37. Liu, D.; Zhou, J.; Liao, H.; Peng, Y.; Peng, X. A health indicator extraction and optimization framework for Li-ion battery degradation modeling and prognostics. IEEE Trans. Syst. Man Cybern. Syst. 2015, 45, 915–928. [Google Scholar]
  38. Liu, D.; Wang, H.; Peng, Y.; Xie, W.; Liao, H. Satellite Li-ion battery remaining cycle life prediction with novel indirect health indicator extraction. Energies 2013, 6, 3654–3668. [Google Scholar] [CrossRef]
  39. Wang, C.; Lu, N.; Wang, S.; Cheng, Y.; Jiang, B. Dynamic long short-term memory neural-network-based indirect remaining-useful-life prognosis for satellite Li-ion battery. Appl. Sci. 2018, 8, 2078. [Google Scholar] [CrossRef] [Green Version]
  40. Zhou, Y.; Huang, M.; Chen, Y.; Tao, Y. A novel health indicator for on-line Li-ion batteries remaining useful life prediction. J. Power Sources 2016, 321, 1–10. [Google Scholar] [CrossRef]
  41. Yang, L.; Zhao, L.; Su, X.; Wang, S. A Li-ion battery RUL prognosis method using temperature changing rate. In Proceedings of the 2016 IEEE International Conference on Prognostics and Health Management (ICPHM), Ottawa, ON, Canada, 20–22 June 2016; pp. 1–7. [Google Scholar]
  42. Cortes, C.; Vapnik, V. Support-vector networks. Mach. Learn. 1995, 20, 273–297. [Google Scholar] [CrossRef]
  43. Kennedy, J.; Eberhart, R. Particle swarm optimization. In Proceedings of the ICNN’95-International Conference on Neural Networks, Perth, WA, Australia, 27 November–1 December 1995; Volume 4, pp. 1942–1948. [Google Scholar]
  44. Li, W.; Feng, F.; Jiang, Q. Prediction for railway passenger volume based on modified PSO optimized LSTM neural network. J. Railw. Sci. Eng. 2018, 15, 3274–3280. [Google Scholar]
  45. Priya, K.; Babu, T.; Balasubramanian, K.; Kumar, K.; Rajaseka, N. A novel approach for fuel cell parameter estimation using simple Genetic Algorithm. Sustain. Energy Technol. Assess. 2015, 12, 46–52. [Google Scholar] [CrossRef]
  46. Askarzadeh, A.; Rezazadeh, A. A new heuristic optimization algorithm for modeling of proton exchange membrane fuel cell: Bird mating optimizer. Int. J. Energy Res. 2013, 37, 1196–1204. [Google Scholar] [CrossRef]
  47. Ali, M.; El-Hameed, M.A.; Farahat, M.A. Effective parameters’ identification for polymer electrolyte membrane fuel cell models using grey wolf optimizer. Renew. Energy 2017, 111, 455–462. [Google Scholar] [CrossRef]
  48. Agwa, A.M.; El-Fergany, A.A.; Sarhan, G.M. Steady-state modeling of fuel cells based on atom search optimizer. Energies 2019, 12, 1884. [Google Scholar] [CrossRef] [Green Version]
  49. Guarnieri, M.; Negro, E.; DI Noto, V.; Alotto, P. A selective hybrid stochastic strategy for fuel-cell multi-parameter identification. J. Power Sources 2016, 332, 249–264. [Google Scholar] [CrossRef] [Green Version]
  50. Goebel, K.; Saha, B.; Saxena, A.; Celaya, J.R.; Christophersen, J.P. Prognostics in battery health management. IEEE Instrum. Meas. Mag. 2008, 11, 33–40. [Google Scholar] [CrossRef]
Figure 1. Flowchart of the PSO–SVR method.
Figure 1. Flowchart of the PSO–SVR method.
Energies 15 02237 g001
Figure 2. Charge and discharge process in the first cycle of the B5 battery.
Figure 2. Charge and discharge process in the first cycle of the B5 battery.
Energies 15 02237 g002
Figure 3. Battery capacity degradation curve.
Figure 3. Battery capacity degradation curve.
Energies 15 02237 g003
Figure 4. Discharge voltage change curves in different cycles of the B5 battery.
Figure 4. Discharge voltage change curves in different cycles of the B5 battery.
Energies 15 02237 g004
Figure 5. Change curves of the IVTs.
Figure 5. Change curves of the IVTs.
Energies 15 02237 g005
Figure 6. Prediction results based on SVR, DNN, LSTM and PSO–SVR using the first 40% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Figure 6. Prediction results based on SVR, DNN, LSTM and PSO–SVR using the first 40% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Energies 15 02237 g006aEnergies 15 02237 g006b
Figure 7. Prediction results based on SVR, DNN, LSTM and PSO–SVR using the first 50% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Figure 7. Prediction results based on SVR, DNN, LSTM and PSO–SVR using the first 50% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Energies 15 02237 g007
Figure 8. Absolute error and mean absolute error based on SVR, DNN, LSTM and PSO–SVR using the first 40% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Figure 8. Absolute error and mean absolute error based on SVR, DNN, LSTM and PSO–SVR using the first 40% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Energies 15 02237 g008
Figure 9. Absolute error and mean absolute error based on SVR, DNN, LSTM and PSO–SVR using the first 50% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Figure 9. Absolute error and mean absolute error based on SVR, DNN, LSTM and PSO–SVR using the first 50% of the total cycle data for training: (a) B5; (b) B6; (c) B7; (d) B18.
Energies 15 02237 g009aEnergies 15 02237 g009b
Table 1. Experimental conditions for the battery dataset.
Table 1. Experimental conditions for the battery dataset.
No.Temperature (°C)Constant Current Charging Current (A)Charge Cut-Off Voltage (V)Constant Discharge Current (A)Discharge Cut-Off Voltage (V)Capacity (Ah)
5241.54.22.02.72.0
6241.54.22.02.52.0
7241.54.22.02.22.0
18241.54.22.02.52.0
Table 2. Results of the correlation analysis between the IVTs and battery capacity.
Table 2. Results of the correlation analysis between the IVTs and battery capacity.
No.Pearson’s Correlation Coefficient
50.99995
60.99980
70.99987
180.99977
Table 3. The results of PSO for optimizing the SVR parameters.
Table 3. The results of PSO for optimizing the SVR parameters.
No.Start PointCγ
56978.170.04017
8514.560.09385
66932.100.06572
857.920.06847
76934.880.05593
8544.260.02135
185496.500.01042
67115.410.57710
Table 4. Comparison of the results of the different prediction methods.
Table 4. Comparison of the results of the different prediction methods.
No.Start PointMethodActual RULPredicted RULErrorMAERMSER2
569DNN56--0.11750.1313−0.6972
LSTM46100.02820.03300.8924
SVR6480.03120.03250.8961
PSO–SVR5600.00830.01370.9816
85DNN40--0.08520.0957−0.4212
LSTM4000.01250.01850.9470
SVR4330.01570.01830.9479
PSO–SVR4000.00760.01330.9726
669DNN4063230.06750.08020.4622
LSTM3280.04030.04920.7977
SVR3910.03060.03690.8865
PSO–SVR3910.01940.02690.9397
85DNN2453290.08150.0996−0.0090
LSTM1950.02330.03440.08798
SVR2220.04670.05450.6975
PSO–SVR2510.01410.02190.9510
769DNN78--0.09320.1046−0.7833
LSTM7440.02130.02680.8833
SVR88100.01760.01980.9362
PSO–SVR7800.00800.01390.9687
85DNN62--0.05880.0679−0.1013
LSTM6110.01300.02020.9021
SVR6310.01230.01680.9329
PSO–SVR6310.01180.01640.9360
1854DNN43--0.14540.1577−3.0975
LSTM--0.12320.1322−1.8809
SVR4740.01680.02190.9211
PSO–SVR4740.01690.02190.9211
67DNN30--0.03480.03910.3564
LSTM3440.01950.02510.7340
SVR51210.03010.03350.5282
PSO–SVR3110.01770.02300.7774
Table 5. Comparison of the results of PSO–SVR to that of other prediction methods.
Table 5. Comparison of the results of PSO–SVR to that of other prediction methods.
No.Start PointMethodActual RULPredicted RULError
568ALO–SVR [29]43430
PSO–SVR43441
84ALO–SVR [29]27270
PSO–SVR27270
668ALO–SVR [29]31354
PSO–SVR32320
84ALO–SVR [29]15194
PSO–SVR16171
768ALO–SVR [29]789719
PSO–SVR79790
84ALO–SVR [29]627412
PSO–SVR63641
1853ALO–SVR [29]30388
PSO–SVR30388
66ALO–SVR [29]17192
PSO–SVR17170
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zou, L.; Wen, B.; Wei, Y.; Zhang, Y.; Yang, J.; Zhang, H. Online Prediction of Remaining Useful Life for Li-Ion Batteries Based on Discharge Voltage Data. Energies 2022, 15, 2237. https://doi.org/10.3390/en15062237

AMA Style

Zou L, Wen B, Wei Y, Zhang Y, Yang J, Zhang H. Online Prediction of Remaining Useful Life for Li-Ion Batteries Based on Discharge Voltage Data. Energies. 2022; 15(6):2237. https://doi.org/10.3390/en15062237

Chicago/Turabian Style

Zou, Lin, Baoyi Wen, Yiying Wei, Yong Zhang, Jie Yang, and Hui Zhang. 2022. "Online Prediction of Remaining Useful Life for Li-Ion Batteries Based on Discharge Voltage Data" Energies 15, no. 6: 2237. https://doi.org/10.3390/en15062237

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