Next Article in Journal
Tuning the Refractive Index Sensitivity of LSPR Transducers Based on Nanocomposite Thin Films Composed of Noble Metal Nanoparticles Dispersed in TiO2
Previous Article in Journal
Rice Husk-Based Adsorbents for Removal of Metals from Aqueous Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prediction of the Fatigue Strength of Steel Based on Interpretable Machine Learning

1
Institute of Structural Steel, Central Iron and Steel Research Institute, Beijing 100081, China
2
Material Digital R&D Center, China Iron and Steel Research Institute Group, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(23), 7354; https://doi.org/10.3390/ma16237354
Submission received: 6 November 2023 / Revised: 23 November 2023 / Accepted: 23 November 2023 / Published: 26 November 2023
(This article belongs to the Section Metals and Alloys)

Abstract

:
Most failures in steel materials are due to fatigue damage, so it is of great significance to analyze the key features of fatigue strength (FS) in order to improve fatigue performance. This study collected data on the fatigue strength of steel materials and established a predictive model for FS based on machine learning (ML). Three feature-construction strategies were proposed based on the dataset, and compared on four typical ML algorithms. The combination of Strategy Ⅲ (composition, heat-treatment, and atomic features) and the GBT algorithm showed the best performance. Subsequently, input features were selected step by step using methods such as the analysis of variance (ANOVA), embedded method, recursive method, and exhaustive method. The key features affecting FS were found to be TT, mE, APID, and Mo. Based on these key features and Bayesian optimization, an ML model was established, which showed a good performance. Finally, Shapley additive explanations (SHAP) and symbolic regression (SR) are introduced to improve the interpretability of the prediction model. It had been discovered through SHAP analysis that TT and Mo had the most significant impact on FS. Specifically, it was observed that 160 < TT < 500 and Mo > 0.15 was beneficial for increasing the value of FS. SR was used to establish a significant mathematical relationship between these key features and FS.

1. Introduction

Metal fatigue fracture refers to the sudden brittle fracture of a metal material under the action of alternating stress after a certain number of cycles. In total, 50% to 90% of component failures are due to fatigue damage, and there is no obvious deformation of the component before fatigue failure, making it difficult to detect. This poses significant risks to actual production. Therefore, it is crucial to predict the fatigue performance of materials in advance. Traditional fatigue performance is obtained through material fatigue tests, but a single fatigue test cycle is time-consuming, resulting in a significant waste of material development costs and time. As a result, researchers have been exploring a more rapid and efficient way to predict fatigue strength (FS) [1,2]. Researchers have found that there is a certain relationship between fatigue strength and other mechanical properties of materials. For example, Wöhler et al. [3] found a linear relationship between the fatigue strength and tensile strength of materials. However, this linear relationship no longer holds as tensile strength increases further [4]. Subsequent researchers proposed other modification models [5], but they are only effective for material strength at lower levels. Additionally, material strength is influenced by factors such as composition [6], microstructure [7], and processing [8]. Establishing a predictive model between fatigue strength and other properties cannot reveal the most essential characteristics, thus having certain limitations.
With the advancement of computer science, machine learning, as an emerging technology, has been widely applied in the field of materials science. Machine learning possesses powerful fitting and predictive capabilities, allowing for the establishment of mapping relationships between material influencing factors (such as composition and processing) and target variables (such as microstructure and properties). This enables the prediction of material composition, microstructure, processing, and performance, as well as the discovery of new materials. Zhao et al. [9] developed a predictive model for the mechanical properties of cold-rolled steel using extreme random trees, achieving high-precision predictions of yield strength, tensile strength, and elongation. Lee et al. [10] utilized machine learning to design high-strength medium manganese steel, successfully developing a new type of steel with higher tensile strength than existing steel grades and almost no loss in elongation. These examples demonstrate the potential of machine learning in the materials field, enabling more accurate predictions of material properties and the discovery of novel materials with a superior performance.
This study proposes three feature-construction strategies and establishes a high-precision prediction model for FS based on machine learning. Through a series of feature-selection methods, key features influencing FS are identified. the interpretability of the prediction model is enhanced by SHAP (Shapley additional explaining) and symbolic regression (SR), and the influence mechanism of key features on FS is revealed. By adopting these strategies and utilizing machine-learning techniques, the study effectively constructs the prediction model for FS, which provides valuable insights into the influence of key features on FS.

2. Materials and Methods

Typically, the performance of materials is controlled by the combination of composition, microstructure, and processing. However, the variations and combinations of these factors result in a vast search space, making it impractical to study them individually through manual analysis. Machine learning, on the other hand, explores the hidden relationships between features and target variables, establishes quantitative models between them, and identifies the influence of features on the target variables, thereby accelerating the efficiency of materials’ research and development [11,12,13,14]. In this study, a prediction model for FS is established considering the factors of composition, processing, and microstructure. Key features are identified through feature selection, and finally, the specific expression of the prediction model is obtained using symbolic regression, clarifying the influence of each feature on FS and providing reasonable optimization suggestions. The specific workflow is illustrated in Figure 1. Initially, relevant literature and experimental data are collected and cleaned to obtain an appropriate dataset. The dataset includes information about the composition, heat treatment, and inclusion distribution of bearing steel. Three feature-construction strategies are designed based on this information. Strategy I involves inputting the composition features of steel, Strategy II involves inputting both composition and heat-treatment features, and Strategy III incorporates the atomic features of elements on top of Strategy II. The features generated by these three strategies are separately fed into different machine-learning models, and multiple metrics are used to evaluate the performance of the models in order to identify the best-performing one. Based on this, feature selection is performed on the input features. Correlation analysis, recursive feature elimination, and exhaustive methods are employed to identify the key features that have an impact on FS. Finally, SR and SHAP are used for key features to increase the interpretability of the model.

2.1. Data Collection and Processing

The data of this study comes from the dataset of steel fatigue strength of Japan National Institute of Materials (NIMS) [15]. The specific dataset has been attached to the Supplementary Materials in S1. The dataset consists of 26 variables, and the specific meanings and distributions of the data are shown in Table 1. The specific data are in the Supplementary Materials. The variables in the dataset include 10 composition parameters, 12 heat-treatment parameters, a rolling parameter, 3 inclusion parameters, and the target performance variable (FS). During the data-cleaning process, data with numerical anomalies were removed, and data with missing values were ignored. Ultimately, 437 sets of data were obtained for subsequent feature selection and model evaluation. To eliminate the disparity in magnitudes between different features and achieve dimensionless data, normalization was applied to the dataset. The formula for the normalization process is shown in Equation (1) [16,17].
X = X min ( X ) max ( X ) min ( X )
where X* represents the normalized feature, and min(X) and max(X) represent the minimum and maximum values of the original feature, respectively.

2.2. Machine-Learning Algorithm

Machine-learning algorithms can be categorized into classification algorithms and regression algorithms, depending on their applicability. Regression algorithms can be further classified into linear algorithms, nonlinear algorithms, and ensemble algorithms. Different algorithms are suitable for different problems, and there is no single algorithm that performs well on universal problems. Therefore, in this study, artificial neural network (ANN), elastic net regression (EN), gradient boosting machine (GBT), and bagging regression (BGR) algorithms are used to predict FS. The principle of the algorithm is introduced in the Supplementary Materials in S2. Cross-validation is employed to evaluate the performance of the models.
To evaluate the performance of different models, several commonly used statistical metrics for regression are employed, including the coefficient of determination (R2) and mean absolute percentage error (MAPE). R2 measures the proportion of the variance in the dependent variable that is predictable from the independent variables. It ranges from 0 to 1, where a value closer to 1 indicates a better fit between the predicted and actual values. An R2 value of 1 indicates that the model perfectly explains the variability of the target variable. The formula for calculating R2 is shown in Equation (2). MAPE measures the deviation between the predicted values and the actual values. Smaller values indicate a better predictive performance of the model. The specific formulas for calculating MAPE are shown in Equation (3) [18].
R 2 = 1 i = 1 n ( f i y i ) 2 i = 1 n ( y i y i ) 2
M A P E = 1 n i = 1 n f i y i y i × 100 %
where n represents the number of data, yi represents the actual value, fi represents the predicted value, and ȳi represents the mean of the actual values.
In order to effectively evaluate machine-learning models, a 5-fold cross-validation approach is used to calculate the evaluation metrics during algorithm selection and subsequent feature selection. Specifically, the original data is divided into 5 subsets, with 4 subsets used for training the model and 1 subset used for validation. This process is repeated 5 times to calculate the evaluation metrics, and the average value of the 5 iterations is taken as the final performance of the model [19]. For the final model construction, a holdout method is employed where the dataset is divided into an 80% training set and a 20% test set. The model is trained using the training set and then evaluated on the test set.

2.3. Feature Selection

The selection of input features has a significant impact on the performance of machine-learning models. Sometimes, there may be redundant features that have an ineffective or even harmful effect on the models. Therefore, it is necessary to analyze the relationship between input features and the target variable to select the key features that influence the target variable. Common methods for feature selection in machine learning include filter methods, embedded methods, and wrapper methods [20]. In this study, filter methods and embedded methods are primarily used. Filter methods select the most predictive or informative subset of features from the original feature set, aiming to reduce dimensionality and enhance model performance. The basic idea of filter methods is to evaluate and rank features based on their statistical characteristics or correlations and select the top-ranked features as the final feature subset. In this research, the analysis of variance (ANOVA) test [21,22] is mainly adopted. ANOVA is based on the analysis of variance, which decomposes the total variance of a population into between-group variance and within-group variance. By comparing the magnitudes of these two variances, ANOVA determines whether there are significant differences in means among different groups. The between-group variance reflects the degree of difference among different groups, while the within-group variance reflects the variability within each group. If the between-group variance is significantly greater than the within-group variance, indicating that the proportion of between-group variance to total variance is relatively large, it can be concluded that there are significant mean differences among different groups. ANOVA utilizes the F-statistic for hypothesis testing. The F-statistic is the ratio of the between-group variance to the within-group variance. Based on the results of hypothesis testing, the F-value and p-value are obtained, where the p-value indicates the significance level of the difference. Generally, when the p-value is less than 0.05 [23], it suggests a significant linear relationship between the feature and the target label, while a p-value greater than 0.05 indicates no significant linear relationship between the feature and the label, and thus the feature can be discarded. Embedded methods [24,25,26], on the other hand, automatically select the optimal feature subset during the training process of the machine-learning model to improve the model performance and generalization ability. The main idea of embedded methods is to combine the feature-selection process with the model-training process. By evaluating the importance or weight of features, embedded methods embed the feature selection into the model training. In each iteration, the embedded method updates the feature subset based on the importance of features according to a predefined threshold, until a pre-defined stopping criterion or the desired number of selected features is reached.
After the selection through filter methods and embedded methods, it is often necessary to further reduce the dimensionality of the selected features. In this case, recursive and exhaustive methods [27,28,29] are commonly used for dimensionality reduction. Recursive and exhaustive methods involve calculating the performance of the model by systematically combining features, selecting the best feature combinations. However, these methods are computationally expensive due to the exhaustive search process. Therefore, they are usually employed when further feature selection is needed after the initial feature-selection steps.

3. Results and Discussion

3.1. Comparison of Machine-Learning Algorithms

Figure 2 shows a comparison of the performance of machine-learning models under different feature-construction strategies. Figure 2a shows the performance comparison of different models under Strategy I. Among them, the GBT model exhibits the highest accuracy (R2 = 0.92, MAPE = 8.15%). However, the model accuracy is not sufficiently high, indicating that predicting FS based solely on composition is insufficient. Thus, in Strategy Ⅱ, the heat-treatment parameters are introduced. The performance of different machine-learning models under Strategy Ⅱ is shown in Figure 2b. Again, the GBT model has the highest accuracy (R2 = 0.98, MAPE = 3.14%). Compared to Strategy I, the model accuracy is significantly improved, indicating the important influence of heat-treatment parameters on the FS of steel. To further improve the model accuracy, Strategy Ⅲ introduces the atomic features of elements, as shown in Table 2. Figure 2c illustrates the performance of different machine-learning models corresponding to Strategy Ⅲ, where the GBT model achieves the highest prediction accuracy (R2 = 0.98, MAPE = 3.05%). Compared to Strategy Ⅱ, the model accuracy is further improved, indicating the effectiveness of introducing atomic features. In addition, although ANN and GBT have a similar prediction accuracy, ANN has many parameters, which leads to a long running time, so GBT is selected as the prediction algorithm. As the GBT model exhibits the highest accuracy under all three strategies, Figure 2d shows a comparison of evaluation metrics under three strategies. Different evaluation metrics all indicate that Strategy III achieves the highest accuracy.

3.2. Key Feature Screening

After determining the feature-construction strategy and machine-learning algorithm, the next step is to select the features to reduce the input dimensionality and model complexity while maintaining model accuracy. Figure 3 illustrates the feature-selection process. Figure 3a shows the correlation heatmap of the original input features, indicating a high correlation between certain features, with correlation coefficients reaching 0.95 or even 1. This suggests the need for feature correlation filtering. By conducting ANOVA tests, features with p-values greater than 0.05 are removed, reducing the feature dimensionality from 41 to 35. Next, the embedded method is used in conjunction with the GDB model to automatically select features, with MAPE chosen as the evaluation metric. Figure 3b shows the variation in MAPE with the input threshold of the embedded method. When the threshold is greater than 0, it indicates that features have been removed, and in this case, the model’s MAPE decreases, indicating the existence of redundant original input features that are detrimental to the model. When the threshold is set to 0.001032, the model’s performance improves compared to the original features (MAPE = 2.97%). Further increasing the threshold would lead to a decrease in model performance, so 0.001032 is chosen as the input threshold for the embedded method, reducing the feature dimensionality from 35 to 23. Figure 3c illustrates the change in the model MAPE with the recursive feature elimination method. As the number of input features increases, the MAPE decreases, indicating an improvement in model performance. When the number of features exceeds 8, the MAPE remains relatively unchanged, indicating that further increasing the number of features has no impact on model performance (MAPE = 2.90%). Figure 3d shows the impact of the number of features on the model R2 and MAPE through exhaustive feature selection. As the number of features increases, R2 increases and MAPE decreases, indicating an improvement in model accuracy. The highest accuracy is achieved when the number of features is four (R2 = 0.98, MAPE = 2.55%), corresponding to the feature combination TT, mE, APID, and Mo.
TT refers to the tempering temperature of steel, which mainly affects the transformation of steel microstructure during tempering and consequently influences the steel’s properties [30,31,32]. Numerous studies have already demonstrated this point. mE refers to the average electronegativity. Both reflect the ease of electron loss by atoms and indicate the strength of the metallic bonds between atoms. The mechanical strength of metals largely depends on the strength of the metallic bonds [33]. The greater the ability of an atom to provide electrons, the larger its contribution to the density of free electrons and the stronger the metallic bond. Therefore, the electronegativity can be used to estimate the strength of metallic bonds [34]. APID refers to the proportion of inclusions in steel that are deformed through plastic working. Inclusions have different compositions compared to the steel matrix. When subjected to external forces, stress concentrations occur around the inclusions. Therefore, inclusions often serve as origins of fatigue fractures and have a significant impact on the fatigue strength of steel [35,36]. The addition of Mo can effectively enhance precipitation strengthening and phase-equilibrium strengthening in steel [6], thereby influencing its properties.

3.3. Optimal Model Establishment

Next, an optimized machine-learning model is built using the GBT algorithm on the selected key features. During the model-building process, it is necessary to adjust the hyperparameters of the GBT algorithm to achieve optimal performance. The GBT algorithm has numerous hyperparameters, and traditional grid search or random search methods involve exhaustive searching in the parameter space, which is computationally expensive and inefficient. Bayesian optimization [37,38], on the other hand, intelligently selects parameter combinations for evaluation by modeling the relationship between parameters and the objective function, thus finding the optimal parameters faster. Bayesian optimization first requires defining the search space for hyperparameters. In this study, the parameter space is defined as follows: ‘n_estimators’ = [50, 2000], ‘learning_rate’ = [0.001, 1], ‘max_depth’ = [2, 20], ‘subsample’ = [0.5, 1], ‘min_samples_split’ = [2, 50], ‘min_samples_leaf’ = [1, 20]. The specific meaning of hyperparameters is listed in the Supplementary Materials in S2. The MAPE from a five-fold cross-validation is chosen as the objective function. Gaussian process regression is used to model the mapping relationship between the parameters and the objective function. Based on the current parameter combination, the objective function value is calculated. Using Bayesian theorem, the posterior distribution of parameters is computed based on the prior distribution and the objective function values. The next parameter combination to evaluate is selected based on the posterior distribution and selection strategy. This process is repeated until the specified number of iterations is reached.
Figure 4 shows the change in MAPE with the number of iterations during the Bayesian optimization process. As the parameters are iteratively updated, MAPE continuously decreases, indicating the beneficial impact of Bayesian optimization on model performance. When the number of iterations reaches 230, MAPE no longer decreases (MAPE = 2.50%), and the iteration converges. The corresponding hyperparameter combination at this point is ‘n_estimators’ = 2000, ‘learning_rate’ = 0.0155, ‘max_depth’ = 13, ‘subsample’ = 0.659, ‘min_samples_split’ = 42, ‘min_samples_leaf’ = 1.
An optimized machine-learning model is built based on the hyperparameters obtained through Bayesian optimization. Figure 5 shows the comparison between the predicted and experimental values of the model on the training set and the test set. The data points on the diagonal line indicate a good predictive performance of the model (train set: R2=0.99, MAPE = 0.70%; test set: R2 = 0.98, MAPE = 2.58%).

3.4. Interpretable Analysis

3.4.1. SHAP

To enhance the interpretability of the machine-learning model, SHAP values [39,40] are introduced. SHAP is a method that calculates the contribution of each feature by considering permutations and combinations of features to determine their impact on the prediction. SHAP employs a “feature stacking” approach, starting with cases that include only one feature and gradually adding features while computing their contributions to the results. When incorporating a new feature, it computes prediction results for all relevant feature combinations and measures the resulting changes. The SHAP value represents the average contribution of each feature to the variation in the result. To ensure equitable allocation, SHAP values consider different possibilities of feature permutations and combinations and utilize Shapley’s core to handle feature interactions. The Shapley core is a distribution scheme that guarantees the fairness principles of cooperative game theory when assigning feature contributions. By calculating the SHAP values for each feature, we can ascertain their relative importance in predicting the outcome. This interpretability enhances the comprehension of the model’s decision-making process and illuminates relationships and influences among the features.
Figure 6 presents the interpretability analysis of the optimized machine-learning model using SHAP values. Figure 6a shows the distribution of the absolute values of the averaged SHAP values for each feature, where TT has the highest impact on the model’s output. Figure 6b illustrates the distribution of SHAP values for each sample, with blue representing low-data samples and red representing high-data samples. The x-axis represents the magnitude of the SHAP values, with positive and negative values indicating a positive or negative impact on the target variable, respectively. Taking the Mo feature as an example, when the Mo data value is high and red, the corresponding SHAP value distribution is on the positive half of the x-axis, indicating a positive effect on FS. Conversely, when the Mo data value is low and blue, the corresponding SHAP value distribution is on the negative half of the x-axis, indicating a detrimental effect on FS. Therefore, to improve the FS of the steel, it is recommended to increase the Mo content. Figure 7 displays the distribution of the SHAP values for TT and Mo. When 160 < TT < 500, the SHAP value is positive, indicating a positive impact on FS. Similarly, when Mo > 0.15, the SHAP value is positive, also contributing to the increase in FS. Conversely, when TT and Mo are in other ranges, the SHAP value is negative, impeding the increase in FS. It should be noted that the increase in Mo content is not unlimited. Excessive Mo content may have adverse effects on the properties of steel [41]. However, the dataset used in this analysis does not include such data, so it is unable to analyze the critical Mo content.

3.4.2. SR

Due to the black-box nature of machine-learning models, even with the introduction of SHAP for interpretability analysis, it is still not possible to obtain specific mathematical expressions for the relationship between features and the target variable. To enhance the interpretability of the model, this study introduces SR [42,43,44], which establishes mathematical expressions between inputs and outputs. Gplearn [45,46,47], a Python open-source library based on genetic programming (GP), is utilized for SR. The process is as follows: Initially, a population of mathematical expression individuals is randomly generated, typically consisting of basic operators and functions. In this study, the basic operators include addition, subtraction, multiplication, and division, aiming to simplify the expressions as much as possible. Next, fitness evaluation is performed. The input features of the training data are plugged into the mathematical expressions to obtain predicted outputs. The fitness score of each individual is calculated by comparing the predicted outputs with the actual outputs, measuring their fit to the data. Based on the fitness scores, the selection operation is applied to choose individuals from the population for the next generation. In the selected parents, two individuals are randomly chosen for the crossover operation. The crossover operation occurs at the chromosome level of the genome, where parts of the expressions between the two individuals are exchanged to generate new individuals. Mutation may be applied to the crossover individuals to introduce new genetic variations and diversity. The mutation operation randomly modifies the mathematical expressions of individuals, such as changing constant values or replacing certain functions. The process of selection, crossover, and mutation is repeated, generating new individuals and calculating their fitness. This evolutionary iteration process gradually improves the individuals in the population to better fit the training data until the maximum number of iterations is reached, and the evolution process stops. The individual with the best fitness is selected as the final symbolic regression model.
The hyperparameters set for SR in this study are shown in Table 3. After 200 iterations, the expression obtained is very complicated, and the simplified formula is obtained after manual selection, as shown in Equation (4). Figure 8 compares the predicted values of FS based on Equation (4) with the experimental values in the training and testing sets. At this stage, the R2 values are both greater than 0.87 for the test and train sets, indicating the good prediction accuracy of SR. According to Equation (4), to obtain a better FS, the Mo content (Mo) and average valence electron number should be increased, while the quantity of inclusions in the steel (APID) should be decreased. Additionally, the tempering temperature (TT) should be appropriately increased.
The established SR expression can visually describe the relationship between steel features and FS, providing guidance for designing new high-FS steel grades. Compared to traditional fatigue performance research and development methods, predictive models established through machine learning and symbolic regression offer improved research and development efficiency, facilitating the exploration of higher-performance metal alloys.
F S = ( c 0 × M o + c 1 × m E + T T × ( c 2 × A P I D c 3 × T T + M o ( c 4 × T T + c 5 ) × c 6 + c 7 × m E c 8 × T T + c 9 ) × c 10 + c 11 )
where c0 = 45.355, c1 = 47.371, c2 = 3.6436, c3 = 1.5942, c4 = 1.9598, c5 = 24.132, c6 = −1.726, c7 = −0.27831, c8 = 0.19812, c9 = 12.105, c10 = −61.577, c11 = 1108.7.

4. Conclusions

(1)
Among the three feature-construction strategies and four machine-learning algorithms, the combination of Strategy Ⅲ and the GBT algorithm yields the best predictive accuracy.
(2)
By using the ANOVA test, embedded method, recursive feature elimination, and exhaustive search, the key feature combination influencing FS is determined to be: TT, mE, APID, and Mo.
(3)
Taking the key features as inputs, the final predictive model is established by adjusting the hyperparameters of the GBT algorithm using Bayesian optimization. At this stage, the model exhibits a good predictive accuracy (test set: R2 = 0.98, MAPE = 2.58%).
(4)
SHAP values are used for interpretability analysis of the machine-learning model, providing insights into the contribution of each feature to the target value. When 160 < TT < 500 and Mo > 0.15, it is beneficial for increasing the value of FS. Symbolic regression is utilized to establish a mathematical expression between the key features and FS, effectively explaining the underlying mechanism of feature impact on FS.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16237354/s1.

Author Contributions

Data curation, C.L. and X.W.; formal analysis, C.L. and W.C.; investigation, C.L., X.W., J.Y. and W.C.; Conceptualization, H.S.; methodology, H.S. and J.Y.; software, H.S.; validation, C.L. and J.Y.; funding acquisition, H.S. supervision, H.S.; writing—original draft preparation, C.L.; writing—review and editing, H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (Grant No. 2021YFB3501502, Grant No. 2021YFB3701201).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The authors are very grateful to the reviewers and editors for their valuable suggestions, which have helped to improve the paper substantially.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Levitin, V. Fatigue of Metals. In Interatomic Bonding in Solids: Fundamentals, Simulation, and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2014. [Google Scholar]
  2. Lee, Y. Fatigue Testing and Analysis: Theory and Practice; Butterworth-Heinemann: Oxford, UK, 2005. [Google Scholar]
  3. Tóth, L.; Yarema, S.Y. Formation of the science of fatigue of metals. Part 1. Mater. Sci. 2006, 42, 673–680. [Google Scholar] [CrossRef]
  4. Pang, J.C.; Li, S.X.; Wang, Z.G.; Zhang, Z.F. General relation between tensile strength and fatigue strength of metallic materials. Mater. Sci. Eng. A 2013, 564, 331–341. [Google Scholar] [CrossRef]
  5. Murakami, Y. Metal Fatigue: Effects of Small Defects and Nonmetallic Inclusions; Academic Press: Cambridge, MA, USA, 2019. [Google Scholar]
  6. Rezende, A.B.; Fonseca, S.T.; Miranda, R.S.; Fernandes, F.M.; Grijalba, F.A.F.; Farina, P.F.S.; Mei, P.R. Effect of niobium and molybdenum addition on the wear resistance and the rolling contact fatigue of railway wheels. Wear 2021, 466, 203571. [Google Scholar] [CrossRef]
  7. Ebara, R. The influence of metallurgical factors on corrosion fatigue strength of stainless steels. Procedia Struct. Integr. 2016, 2, 517–524. [Google Scholar] [CrossRef]
  8. Pyttel, B.; Schwerdt, D.; Berger, C. Fatigue strength and failure mechanisms in the VHCF-region for quenched and tempered steel 42CrMoS4 and consequences to fatigue design. Procedia Eng. 2010, 2, 1327–1336. [Google Scholar] [CrossRef]
  9. Zhao, Y.-B.; Song, Y.; Li, F.-F.; Yan, X.-L. Prediction of mechanical properties of cold rolled strip based on improved extreme random tree. J. Iron Steel Res. Int. 2023, 30, 293–304. [Google Scholar] [CrossRef]
  10. Lee, J.; Kim, M.; Lee, Y. Design of high strength medium-Mn steel using machine learning. Mater. Sci. Eng. A 2022, 843, 143148. [Google Scholar] [CrossRef]
  11. Eser, A.; Broeckmann, C.; Simsir, C. Multiscale modeling of tempering of AISI H13 hot-work tool steel–Part 1: Prediction of microstructure evolution and coupling with mechanical properties. Comput. Mater. Sci. 2016, 113, 280–291. [Google Scholar] [CrossRef]
  12. Guo, S.; Yu, J.; Liu, X.; Wang, C.; Jiang, Q. A predicting model for properties of steel using the industrial big data based on machine learning. Comput. Mater. Sci. 2019, 160, 95–104. [Google Scholar] [CrossRef]
  13. Jung, J.; Yoon, J.I.; Park, H.K.; Kim, J.Y.; Kim, H.S. Bayesian approach in predicting mechanical properties of materials: Application to dual phase steels. Mater. Sci. Eng. A 2019, 743, 382–390. [Google Scholar] [CrossRef]
  14. Fehlemann, N.; Aguilera, A.L.S.; Sandfeld, S.; Bexter, F.; Neite, M.; Lenz, D.; Könemann, M.; Münstermann, S. Identification of martensite bands in dual phase steels–a deep learning object detection approach using Faster R-CNN. Steel Res. Int. 2023, 94, 2200836. [Google Scholar] [CrossRef]
  15. National Institute of Materials Science. Available online: http://smds.nims.go.jp/fatigue/index_en.html (accessed on 15 January 2023).
  16. Hong, D.; Kwon, S.; Yim, C. Exploration of Machine Learning to Predict Hot Ductility of Cast Steel from Chemical Composition and Thermal Conditions. Met. Mater. Int. 2021, 27, 298–305. [Google Scholar] [CrossRef]
  17. Liu, Y.; Wu, J.; Wang, Z.; Lu, X.-G.; Avdeev, M.; Shi, S.; Wang, C.; Yu, T. Predicting creep rupture life of Ni-based single crystal superalloys using divide-and-conquer approach based machine learning. Acta Mater. 2020, 195, 454–467. [Google Scholar] [CrossRef]
  18. Chen, Y.; Wang, S.; Xiong, J.; Wu, G.; Gao, J.; Wu, Y.; Ma, G.; Wu, H.-H.; Mao, X. Identifying facile material descriptors for Charpy impact toughness in low-alloy steel via machine learning. J. Mater. Sci. Technol. 2023, 132, 213–222. [Google Scholar] [CrossRef]
  19. Cui, C.; Cao, G.; Li, X.; Gao, Z.; Liu, J.; Liu, Z. A strategy combining machine learning and physical metallurgical principles to predict mechanical properties for hot rolled Ti micro-alloyed steels. J. Mater. Process. Technol. 2023, 311, 117810. [Google Scholar] [CrossRef]
  20. Stańczyk, U.; Jain, L.C. Feature Selection for Data and Pattern Recognition: An Introduction. In Feature Selection for Data and Pattern Recognition; Stańczyk, U., Jain, L.C., Eds.; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1–7. [Google Scholar]
  21. St»hle, L.; Wold, S. Analysis of variance (ANOVA). Chemom. Intell. Lab. Syst. 1989, 6, 259–272. [Google Scholar] [CrossRef]
  22. Shaw, R.G.; Mitchell-Olds, T. Anova for Unbalanced Data: An Overview. Ecology 1993, 74, 1638–1645. [Google Scholar] [CrossRef]
  23. Agrawal, A.; Choudhary, A. An online tool for predicting fatigue strength of steel alloys based on ensemble data mining. Int. J. Fatigue 2018, 113, 389–400. [Google Scholar] [CrossRef]
  24. Lal, T.N.; Chapelle, O.; Weston, J.; Elisseeff, A. Embedded Methods. In Feature Extraction: Foundations and Applications; Guyon, I., Nikravesh, M., Gunn, S., Zadeh, L.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2006; pp. 137–165. [Google Scholar]
  25. Chandrashekar, G.; Sahin, F. A survey on feature selection methods. Comput. Electr. Eng. 2014, 40, 16–28. [Google Scholar] [CrossRef]
  26. Maldonado, S.; López, J. Dealing with high-dimensional class-imbalanced datasets: Embedded feature selection for SVM classification. Appl. Soft Comput. 2018, 67, 94–105. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Wen, C.; Wang, C.; Antonov, S.; Xue, D.; Bai, Y.; Su, Y. Phase prediction in high entropy alloys with a rational selection of materials descriptors and machine learning models. Acta Mater. 2020, 185, 528–539. [Google Scholar] [CrossRef]
  28. Zhang, H.; Fu, H.; He, X.; Wang, C.; Jiang, L.; Chen, L.; Xie, J. Dramatically Enhanced Combination of Ultimate Tensile Strength and Electric Conductivity of Alloys via Machine Learning Screening. Acta Mater. 2020, 200, 803–810. [Google Scholar] [CrossRef]
  29. Wen, C.; Zhang, Y.; Wang, C.; Xue, D.; Bai, Y.; Antonov, S.; Dai, L.; Lookman, T.; Su, Y. Machine learning assisted design of high entropy alloys with desired property. Acta Mater. 2019, 170, 109–117. [Google Scholar] [CrossRef]
  30. Qu, S.; Zhang, Y.; Lai, F.; Li, X. Effect of Tempering Temperatures on Tensile Properties and Rotary Bending Fatigue Behaviors of 17Cr2Ni2MoVNb Steel. Metals 2018, 8, 507. [Google Scholar] [CrossRef]
  31. Salemi, A.; Abdollah-zadeh, A. The effect of tempering temperature on the mechanical properties and fracture morphology of a NiCrMoV steel. Mater. Charact. 2008, 59, 484–487. [Google Scholar] [CrossRef]
  32. Sorokin, G.M.; Bobrov, S.N. Effect of tempering temperature on the fatigue strength of high-strength tool steel. Met. Sci. Heat Treat. 1974, 16, 755–757. [Google Scholar] [CrossRef]
  33. Pearson, R.G. Absolute electronegativity and absolute hardness of Lewis acids and bases. J. Am. Chem. Soc. 1985, 107, 6801–6806. [Google Scholar] [CrossRef]
  34. Chapetti, M.D.; Miyata, H.; Tagawa, T.; Miyata, T.; Fujioka, M. Fatigue strength of ultra-fine grained steels. Mater. Sci. Eng. A 2004, 381, 331–336. [Google Scholar] [CrossRef]
  35. Lipiński, T.; Wach, A. Influence of inclusions on bending fatigue strength coefficient the medium carbon steel melted in an electric furnace. Prod. Eng. Arch. 2020, 26, 88–91. [Google Scholar] [CrossRef]
  36. Yao, J.; Qu, X.; He, X.; Zhang, L. Effect of inclusion size on the high cycle fatigue strength and failure mode of a high V alloyed powder metallurgy tool steel. Int. J. Miner. Met. Mater. 2012, 19, 608–614. [Google Scholar] [CrossRef]
  37. Zhang, W.; Wu, C.; Zhong, H.; Li, Y.; Wang, L. Prediction of undrained shear strength using extreme gradient boosting and random forest based on Bayesian optimization. Geosci. Front. 2021, 12, 469–477. [Google Scholar] [CrossRef]
  38. Hernández-Lobato, J.M.; Gelbart, M.A.; Adams, R.P.; Hoffman, M.W.; Ghahramani, Z. A general framework for constrained Bayesian optimization using information-based search. J. Mach. Learn. Res. 2016, 17, 1–53. [Google Scholar]
  39. Yan, Z.; Li, L.; Cheng, L.; Chen, X.; Wu, K. New insight in predicting martensite start temperature in steels. J. Mater. Sci. 2022, 57, 11392–11410. [Google Scholar] [CrossRef]
  40. Yan, F.; Song, K.; Liu, Y.; Chen, S.; Chen, J. Predictions and mechanism analyses of the fatigue strength of steel based on machine learning. J. Mater. Sci. 2020, 55, 15334–15349. [Google Scholar] [CrossRef]
  41. Hashimoto, S.; Ikeda, S.; Sugimoto, K.-I.; Miyake, S. Effects of Nb and Mo addition to 0.2% C-1.5% Si-1.5% Mn steel on mechanical properties of hot rolled TRIP-aided steel sheets. ISIJ Int. 2004, 44, 1590–1598. [Google Scholar] [CrossRef]
  42. Weng, B.; Song, Z.; Zhu, R.; Yan, Q.; Sun, Q.; Grice, C.G.; Yan, Y.; Yin, W.-J. Simple descriptor derived from symbolic regression accelerating the discovery of new perovskite catalysts. Nat. Commun. 2020, 11, 3513. [Google Scholar] [CrossRef] [PubMed]
  43. Loftis, C.; Yuan, K.; Zhao, Y.; Hu, M.; Hu, J. Lattice Thermal Conductivity Prediction Using Symbolic Regression and Machine Learning. J. Phys. Chem. A 2021, 125, 435–450. [Google Scholar] [CrossRef]
  44. Yuan, F.; Mueller, T. Identifying models of dielectric breakdown strength from high-throughput data via genetic programming. Sci. Rep. 2017, 7, 17594. [Google Scholar] [CrossRef]
  45. Kinnear, K.E.; Angeline, P.J.; Spector, L. Advances in Genetic Programming; MIT Press: Cambridge, MA, USA, 1994. [Google Scholar]
  46. Zhang, R.; Lensen, A.; Sun, Y. Speeding up Genetic Programming Based Symbolic Regression Using GPUs. In Proceedings of the PRICAI 2022: Trends in Artificial Intelligence, Shanghai, China, 10–13 November 2022; Khanna, S., Cao, J., Bai, Q., Xu, G., Eds.; Springer Nature Switzerland: Cham, Switzerland; pp. 519–533. [Google Scholar]
  47. Ferreira, J.; Pedemonte, M.; Torres, A.I. A Genetic Programming Approach for Construction of Surrogate Models. In Computer Aided Chemical Engineering; Muñoz, S.G., Laird, C.D., Realff, M.J., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; Volume 47, pp. 451–456. [Google Scholar]
Figure 1. The workflow chart of fatigue strength prediction of steel based on machine learning.
Figure 1. The workflow chart of fatigue strength prediction of steel based on machine learning.
Materials 16 07354 g001
Figure 2. The Comparison of machine-learning algorithms under different strategies: (a) Strategy I, (b) Strategy II, (c) Strategy III, (d) different strategies in the GBT algorithm.
Figure 2. The Comparison of machine-learning algorithms under different strategies: (a) Strategy I, (b) Strategy II, (c) Strategy III, (d) different strategies in the GBT algorithm.
Materials 16 07354 g002
Figure 3. The feature-selection process: (a) the correlation between original features, (b) the influence of threshold input by embedding method on MAPE, (c) the influence of feature number on MAPE in recursive method, (d) the influence of the number of features on R2 and MAPE in the exhaustive method.
Figure 3. The feature-selection process: (a) the correlation between original features, (b) the influence of threshold input by embedding method on MAPE, (c) the influence of feature number on MAPE in recursive method, (d) the influence of the number of features on R2 and MAPE in the exhaustive method.
Materials 16 07354 g003
Figure 4. The influence of iteration times on MAPE in Bayesian optimization.
Figure 4. The influence of iteration times on MAPE in Bayesian optimization.
Materials 16 07354 g004
Figure 5. The comparison between the predicted values of the optimized machine-learning model on the training set and the test set and the experimental values of the training set and the test.
Figure 5. The comparison between the predicted values of the optimized machine-learning model on the training set and the test set and the experimental values of the training set and the test.
Materials 16 07354 g005
Figure 6. Interpretability of the SHAP value to the machine-learning model: (a) mean |SHAP| value (b) SHAP value of each sample.
Figure 6. Interpretability of the SHAP value to the machine-learning model: (a) mean |SHAP| value (b) SHAP value of each sample.
Materials 16 07354 g006
Figure 7. The distribution of SHAP value: (a) TT (b) Mo.
Figure 7. The distribution of SHAP value: (a) TT (b) Mo.
Materials 16 07354 g007
Figure 8. Comparison between the predicted value of SR on the training set and test set and the experimental value of the training set and test set.
Figure 8. Comparison between the predicted value of SR on the training set and test set and the experimental value of the training set and test set.
Materials 16 07354 g008
Table 1. The 26 variables of the used FS data.
Table 1. The 26 variables of the used FS data.
FeaturesDescriptionMaxMinMean
InputsCwt.% of C0.630.170.388
Siwt.% of Si2.050.160.299
Mnwt.% of Mn1.600.370.823
Pwt.% of P0.0310.0020.0157
Swt.% of S0.030.0030.0146
Niwt.% of Ni2.780.010.517
Crwt.% of Cr1.170.010.570
Cuwt.% of Cu0.260.010.0677
Mowt.% of Mo0.2400.0697
Fewt.% of Fe99.07295.10897.233
NTNormalizing Temperature (T)930825872.299
THTThrough Hardening Temperature (T)86530737.643
THtThrough Hardening Time (s)30025.949
CRTHCooling Rate for Through Hardening (T/s)24010.654
CTCarburization Temperature (T)93030128.855
CtCarburization Time (s)540040.502
DTDiffusion Temperature (T)90330123.699
DtDiffusion time (s)7004.843
QMTQuenching Media Temperature (T)1403035.491
TTTempering Temperature (T)68030536.842
TtTempering Time(s)120065.080
CRTCooling Rate for Tempering (T/s)24020.814
RRReduction Ratio (%)5530240923.63
APIDArea Proportion of Inclusions Deformed by Plastic Work (%)0.1300.0472
APIOArea Proportion of Inclusions Occurring in Discontinuous Array (%)0.0500.00339
APIIArea Proportion of Isolated Inclusions (%)0.05800.00771
OutputFSFatigue Strength (MPa)1190225552.90
Table 2. Atomic features.
Table 2. Atomic features.
Abb.Description
mAWmean atomic weight
mCmean column
mRmean row
rNrange number
mNmean number
rARrange atomic radius
mARmean atomic radius
rErange electronegativity
mEmean electronegativity
asveavg s valence electrons
apveavg p valence electrons
adveavg d valence electrons
fsvefrac s valence electrons
fpvefrac p valence electrons
fdvefrac d valence electrons
Table 3. Hyperparameter in symbolic regression.
Table 3. Hyperparameter in symbolic regression.
HyperparameterValue
population_size3000
generations200
tournament_size20
stopping_criteria0.01
p_crossover0.7
p_subtree_mutation0.1
p_hoist_mutation0.05
p_point_mutation0.1
parsimony_coefficient0.01
const_range(−1,1)
feature_names[‘TT’, ‘mE’, ‘APID’, ‘Mo’]
max_samples0.9
verbose1
function_set(‘add’, ’sub’,’mul’,’div’)
metricmean absolute error
random_state50
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

Liu, C.; Wang, X.; Cai, W.; Yang, J.; Su, H. Prediction of the Fatigue Strength of Steel Based on Interpretable Machine Learning. Materials 2023, 16, 7354. https://doi.org/10.3390/ma16237354

AMA Style

Liu C, Wang X, Cai W, Yang J, Su H. Prediction of the Fatigue Strength of Steel Based on Interpretable Machine Learning. Materials. 2023; 16(23):7354. https://doi.org/10.3390/ma16237354

Chicago/Turabian Style

Liu, Chengcheng, Xuandong Wang, Weidong Cai, Jiahui Yang, and Hang Su. 2023. "Prediction of the Fatigue Strength of Steel Based on Interpretable Machine Learning" Materials 16, no. 23: 7354. https://doi.org/10.3390/ma16237354

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