Next Article in Journal
Investigation on Properties of Pervious Concrete Containing Co-Sintering Lightweight Aggregate from Dredged Sediment and Rice Husks
Previous Article in Journal
Ontology for BIM-Based Robotic Navigation and Inspection Tasks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transfer-Learning Prediction Model for Low-Cycle Fatigue Life of Bimetallic Steel Bars

1
School of Civil Engineering, Chongqing University, Chongqing 400045, China
2
Key Laboratory of New Technology for Construction of Cities in Mountain Area, Ministry of Education, Chongqing University, Chongqing 400045, China
3
School of Management Science and Real Estate, Chongqing University, Chongqing 400045, China
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(8), 2275; https://doi.org/10.3390/buildings14082275
Submission received: 18 June 2024 / Revised: 11 July 2024 / Accepted: 19 July 2024 / Published: 23 July 2024
(This article belongs to the Section Building Materials, and Repair & Renovation)

Abstract

:
The prediction of the low-cycle fatigue life of bimetallic steel bars (BSBs) is essential to promote the engineering application of BSBs. However, research on the low-cycle fatigue properties of BSB is limited, and fatigue experiments are time-consuming. Moreover, considering that sufficient data are needed for model training, the lack of data hinders the leverage of typical data-driven machine learning, which is widely used in fatigue life prediction. To address this issue, a transfer learning framework was suggested to accurately predict the low-cycle fatigue life of BSBs with limited data. To achieve this goal, 54 data points obtained from low-cycle fatigue tests on BSBs and 264 data points of other metallic bars were collected. Source models based on artificial neural networks (ANNs) were first constructed using the collected source dataset. Then, the learned knowledge stored in the source models was transferred to the transfer models. After that, transfer models were further fine-tuned and then tested using the target dataset of BSBs. The ANN models, which were of the same structure as the transfer models but only trained with the target dataset without transferring deep features from the source models, were set as baseline models. Compared with baseline models, the constructed transfer models could be used to accurately predict the fatigue life of BSBs. Moreover, the influence of hidden layers of ANNs on accuracy was examined by comparing one-layer and two-layer transfer models. Furthermore, the influence of key parameters on fatigue life of metallic bars was evaluated by feature analysis.

1. Introduction

Longitudinal steel bars in reinforced concrete (RC) members are often damaged by low-cycle fatigue loading during earthquakes [1,2,3,4]. Thus, it is important to determine the low-cycle fatigue life of steel bars to evaluate the hysteretic behavior of RC structures. Several researchers have performed comprehensive investigations on the low-cycle fatigue properties of steel bars. Recent research indicated that the low-cycle fatigue properties of steel bars were influenced by many factors, such as the slenderness ratio (λ), mechanical properties of steel, and fatigue strain amplitude (εa). Aldabagh and Alam [5] investigated the influence of εa and λ on the low-cycle fatigue performance of ASTMA1035 Grade 690 bars. Apostolopoulos [6] conducted low-cycle fatigue tests on corroded reinforcing steel bars (S500s). The experimental results illustrated that the corroded steel bars exhibited a gradual reduction in both their load bearing ability and available energy. Hawileh et al. [7] studied the low-cycle fatigue behavior of ASTM A706 and A615 Grade 60 steel bars with different εa values and established an equation relating dissipated energy to εa and fatigue life. Kashani et al. [8] quantified the combined effects of inelastic buckling and chloride-induced corrosion damage on the low-cycle fatigue life of embedded reinforcing bars in concrete. By utilizing the fatigue life data of existing structures, the fatigue life of newly constructed structures can be predicted, providing a scientific basis for structural design and ensuring the safety and economy of the structure. It should be noted that the dimensions and mechanical properties of steel bar specimens were different in different experimental investigations. Furthermore, the applied fatigue load in previous studies also varied. Therefore, it is difficult to directly compare different experimental results. Although many empirical regression formulas have been proposed by different scholars to clarify the S–N curve of steel bars under low-cycle fatigue loads, the generality of these prediction formulas is questionable. Traditional prediction methods are based on numerical regression, which makes it difficult to capture the complex features hidden in the data, necessitating that they be calculated separately for each situation. The lack of a general quantitative method for determining the low-cycle fatigue properties of steel bars also led to difficulties in engineering design and evaluation. Therefore, it is necessary to establish a unified method to effectively analyze the low-cycle fatigue properties of different types of steel bars. Some investigations were conducted in this study to address this issue.
Bimetallic steel bars (BSBs) are novel steel reinforcements that exhibit outstanding corrosion resistance and good mechanical properties. BSBs consist of stainless steel cladding, carbon steel substrate, and a metallurgical bonding layer between the stainless-steel cladding and carbon steel substrate, which is produced through the hot-rolling process (Figure 1) and ensures cooperative deformation of the stainless-steel cladding and carbon steel substrate in the BSB [9,10]. Research conducted by Shi et al. [11,12] indicated that the metallurgical bonding layer from the hot-rolling process had outstanding connection performance, especially for shear resistance. Stainless-steel cladding enhances the corrosion resistance of BSB by isolating the corrosion factor from the carbon steel substrate, which blocks the process of carbon steel corrosion [13]. Wang et al. [14] performed a series of investigations into the monotonic mechanical properties of BSB. The experimental results indicated that the BSB exhibited good strength properties, and its stress–strain curve was influenced by the cladding ratio. The ductile properties of the bimetallic steel bars (BSBs) met the requirements of RC structures. Li et al. [3] conducted an experimental study on the fatigue properties of BSBs, where the fatigue life and fracture behavior were investigated. The BSB also performs well in terms of residual mechanical properties after exposure to fire [15,16,17]. Based on the experimental results in [18], the bonding properties of BSBs in seawater sea sand concrete were similar to those of the traditional carbon steel bar, where the predictive bond-slip model was suggested. The above investigations indicate that the BSB, which is a substitute for traditional carbon steel bars, can be reasonably applied to RC structures to enhance their durability. Furthermore, a comprehensive economic analysis of BSB was conducted by Liu et al. [19], which indicated that for RC structures in a corrosive environment, the BSB had the lowest life-cycle cost compared with stainless steel bars and carbon steel bars. In general, recent studies on BSB have demonstrated satisfactory engineering application potential.
As stated previously, longitudinal steel bars in concrete members are commonly damaged by low-cycle fatigue loading during earthquakes. Hence, it is necessary to conduct a comprehensive investigation of the low-cycle fatigue behavior of BSB to promote its engineering application. Although there have been studies on the low-cycle fatigue properties of BSB [15,20,21,22], the key parameters considered in previous investigations were very limited, and the dataset is insufficient for a comprehensive predictive method for the low-cycle fatigue properties of BSB. The material properties of bimetallic steel bars are more complex than those of single metals, and their fatigue performance is not only affected by the characteristics of each metal but also by the interaction between the two metal interfaces. Therefore, it is necessary to consider more factors when predicting BSB fatigue life, which increases the difficulty of prediction. Conventional regression prediction methods cannot meet these requirements, so it is necessary to use better models to predict BSB fatigue performance.
Owing to its capability to describe the nonlinear relationship between high-dimensional data, machine learning (ML) has been applied in many fields of structural engineering [23,24,25,26]. Machine learning can consider multiple independent variables, thereby achieving higher accuracy in predicting material properties [27]. Recently, machine learning has been widely selected to quantify fatigue life of various materials [28,29,30,31,32,33,34]. For instance, various ML models have been used to predict the creep, fatigue, and creep fatigue of stainless steel [35]. Yang et al. [36] proposed a novel ML-based method to handle multiaxial fatigue life prediction by utilizing features extracted from a long short-term memory network as the input of a neural network. Kamiyama et al. [37] predicted the low-cycle fatigue crack development of metal thin films using a deep convolutional neural network with microscopic images as the input. By constraining the biases and weights of the neural network, Chen and Liu [38] developed a physics-guided neural network for the estimation of fatigue S–N curves. Guo et al. [39] trained a model using the adaptive boosting algorithm to explore the quantitative influence of corrosion degree, stress ranges, loading ratio, loading frequency and other factors on the fatigue life of corroded high-strength steel wires. The literature mentioned above indicates that ML achieves satisfactory applicability and accuracy for fatigue life prediction. However, the aforementioned studies were all based on the assumption of a sufficient dataset, because ML is mostly a data-driven method. Therefore, despite the high-dimensional regression capacity of typical data-driven ML models, the time-consuming fatigue experiments that are unfavorable for data acquisition hinder the application of typical ML models in fatigue life prediction.
Transfer learning has proven to be one of the most effective approaches, where the small-sample learning needs to be addressed [40,41], and can be used to improve a model from one domain by transferring information from a related domain. The two common steps in transfer learning are deep feature extraction and fine-tuning [42]. In the deep feature extraction, a source model is constructed with the source dataset, and activation values of hidden layers are stored as features. In fine-tuning, the source model is fine-tuned with the target dataset for a similar task by fixing the first several layers and fine-tuning the final layers. It has been widely used in many applications, such as predicting the remaining useful life of machines [43], machine fault diagnosis [44], and plant classification [42]. By combining LSTM and transfer learning, Wei et al. [45] suggested a method to reveal S–N curves. A comprehensive review of the literature on transfer learning can be found in [40,41].
In this paper, a framework based on machine learning and transfer learning is proposed, aimed at achieving efficient prediction of low-cycle fatigue life of BSBs with limited data. This framework is also suitable for fatigue behavior research on other materials. The basic idea is to construct a source model using artificial neural networks (ANNs) based on a comprehensive source dataset collected from existing research on the low-cycle fatigue life of other metallic materials. Then, by leveraging the prior knowledge learned from the source dataset, transfer models fine-tuned using the target dataset of BSBs with very limited data can predict the low-cycle fatigue life of BSBs. Thus, the problem of predicting the low-cycle fatigue life of BSBs with insufficient BSB samples will be efficiently solved. Based on our knowledge, the transfer learning approach to low-cycle fatigue life prediction still remains very limited.
The main contributions of this paper are as follows:
(1) A novel framework based on neural networks and transfer learning is proposed, which significantly reduces the cost and simplifies the quantification of the low-cycle fatigue life of BSBs.
(2) An ML model is trained to efficiently predict the low-cycle fatigue life of various types of metallic bars with high accuracy, which could provide a practical analysis method in engineering.
(3) The relative importance of several key parameters to the low-cycle fatigue life of steel bars is evaluated.
Experimental results from existing research are introduced in Section 2. The framework and model training process are proposed in Section 3. The prediction results are discussed in Section 4. The efficiency of transfer learning is discussed, and the parametric influence is analyzed in Section 5. The main conclusions are presented in Section 6.

2. Experimental Results from Fatigue Experiment

2.1. Test Approach

Although concrete and stirrups were designed to provide an embedding effect on the longitudinal steel bar in RC structures, the concrete between stirrups often cracked and separated (Figure 2) under an earthquake because of the poor tensile properties of concrete. Then, without the embedded effect of concrete, the slenderness ratios of the longitudinal steel bars between the transverse stirrups were relatively large. Thus, the longitudinal steel bars were prone to overall buckling during earthquakes and exhibited noteworthy transverse deformations, as shown in Figure 2. To accurately reveal the low-cycle fatigue properties of steel bars under hysteretic loading, it is necessary to properly simulate the above boundary conditions and steel bar deformations in a laboratory environment. The experimental methods and MTS system shown in Figure 3 were widely employed in previous research [5,7,8,22] to clarify the low-cycle fatigue performance of steel bars. The test specimen shown in Figure 3 consists of a clamping section and a test section. To simulate the embedding effect of concrete and transverse stirrups on the longitudinal steel bar, the clamping section’s in-plane translation and rotation degrees of freedom were limited [21,22,46,47,48,49]. During fatigue loading, the vertical deformation of the testing section, the fatigue load, and the loading cycles were recorded and employed to determine the strain, stress, and fatigue life of specimens, respectively.

2.2. Variables

Cycles to failure (Nf) was selected as the performance index to quantify the low-cycle fatigue properties of steel bars in this study, which has been employed by many researchers in their studies [3,5,22,50,51]. The experimental results indicated that Nf was influenced by many parameters. The mechanical properties of steel bars might cause differences in the accumulation of fatigue damage [52], and the influences of mechanical properties on Nf were reflected by the yield strength fy, ultimate strength fu, and ratio of fy/fu. With respect to the geometric parameters, the diameter d and slenderness ratio λ = l/d significantly affected the failure mode of the steel bars, where l is the length of the test segment. When d is relatively small and λ is relatively large, the specimen under fatigue loading is prone to inelastic buckling, as shown in Figure 4. Large plastic deformation occurred at the inflection points of inelastic buckling deformation, which accelerated the accumulation of fatigue damage and reduced the fatigue life. The effects of fatigue load on Nf are reflected by the fatigue strain amplitude εa. With an increase in εa, more obvious plastic deformation occurred in the steel bars, which accelerated the accumulation of fatigue damage. In addition, the increase in εa led to a greater compressive stress and compression deformation in the steel bars, which aggravated the generation of inelastic buckling. With the increase in εa, the fatigue life Nf of the steel bars decreased gradually. Therefore, the yield strength fy, ultimate strength fu, ratio fy/fu, slenderness ratio λ, diameter d, and fatigue strain amplitude εa are included in the following research. It should be noted that the corrosion effect, which changes the properties of steel bars, was not considered in this study.

2.3. Dataset of Experimental Data

The objective of this study is to quantify the low-cycle fatigue life of BSBs by leveraging existing experimental results on different metallic bars. To achieve this objective, a source dataset consisting of seven types of other metallic bars (264 samples in total) was collected from existing research [3,5,22,50,51], as summarized in Table 1. To avoid a decrease in model accuracy caused by redundancy or noise, the complexity of the dataset should be minimized as much as possible [53,54]. In this study, 7 variables from the dataset were selected based on test results, including fy, fu, fy/fu, d, εa, λ, and Nf. The collected target dataset of BSBs is also listed in Table 1. As stated above, studies on the low-cycle fatigue behavior of BSBs are very limited. Therefore, only 54 samples were collected. It is worth mentioning that among the 54 samples, every three formed a group with the same geometric parameters and loading conditions. The average test result for each group was adopted in the model training. Hence, there were only 18 efficient data samples for BSB, which was insufficient for directly constructing an ML model. The transfer framework in this study, however, will be able to capture the fatigue behavior of BSBs using the dataset of other metallic bars.
The details on the experimental results are listed in Appendix A, Table A1. The statistical distributions of all variables are shown in Figure 5, which has been normalized within the range between zero and one to avoid the scale effect when training the ML models.

3. Framework

The goal of this framework is to realize low-cycle fatigue life prediction of BSBs by leveraging the source dataset of other metallic bars. As shown in Figure 6, the framework consists of two parts. A source model to predict the fatigue life of other metallic bars is constructed based on the collected source dataset with sufficient data. Subsequently, the learned knowledge in the source model is transferred to the transfer model to reveal the fatigue life of BSBs by retraining only a few model parameters with limited BSB data. To predict low-cycle fatigue life, a typical machine learning model, an artificial neural network (ANN), was used.

3.1. Performance Metrics

To assess the constructed model's accuracy, the predicted results are studied in terms of three performance metrics: the coefficient of determination ( R 2 ), root mean squared error (RMSE), and mean absolute error (MAE), which are expressed as
R 2 = 1 i = 1 N y ^ i y i 2 / i = 1 N y i y ¯ 2
R M S E = i = 1 N ( y ^ i y i ) 2 N
M A E = i = 1 N | y ^ i y i | N
where y i denotes the test result for the normalized low-cycle fatigue life of the i t h sample, y ^ i is the fatigue life predicted by the ML models, y ¯ denotes the average value of the test results in the training or testing set, and N denotes the total number of samples in the corresponding set.

3.2. Construction of Source Model

3.2.1. Neural Network Architecture

The ANN model processes input data through a series of layers, with the most critical component being the hidden layer, which is responsible for extracting features from the data to improve model performance. The typical structure of an ANN is illustrated in Figure 7 using an ANN with two hidden layers. The network starts with an input layer. Then, there are two sequential hidden layers. In the end, there is a regression output layer. The behavior of the   1 s t and 2 n d hidden layer can be determined as
y i = f i ( k = 1 6 w k i x k + b i )
z i = f i ( k = 1 n w k i y k + b i )
Then, the output is
o = f i ( k = 1 n w k i z k + b i )
where x is the input feature; w , b , and f denote the weights, bias, and activation functions, respectively; and n is the number of neurons in the hidden layer. The activation function is utilized to ensure that all values passed to the next layer are within the range between zero and one. The initial weights and biases were assigned randomly. The gradient descent algorithm was then used to minimize the loss function in order to update the weights and biases in the training process.

3.2.2. Model Training and Hyperparameter Tuning

The performance of ML models is highly dependent on hyperparameters, and appropriate hyperparameters help improve the model prediction accuracy. In this paper, the hyperparameters were selected based on the grid search methods and K-fold cross-validation [23,24]. Grid search is a methodical approach to tuning the hyperparameters of a machine learning model. The K-fold cross-validation method can fully utilize the dataset by splitting the training set into K folds, and each fold serves for training and validation sequentially. The model performance on different data subsets was evaluated by repeating the training and testing process, so as to ensure that the grouping does not affect the generalization ability of the model. Then, the optimal hyperparameters were determined according to the best average performance score (such as R2) of the training. The grid search method extensively searches for all hyperparameters within specified ranges.
The dataset for the other metallic bars was first randomly split into training (85%) and testing (15%) sets. The training set was then split ten-fold. The model was trained 10 times for the given hyperparameters, wherein in each iteration, nine out of the 10 folds were used for training and the remaining one was selected for validation. The performance scores of all 10 folds were averaged to evaluate the performance of the hyperparameters. After the grid search tuning phase, the best performing model was the one with the highest score. Five hyperparameters which significantly affected the model’s performance were tuned, i.e., the number of neurons in the hidden layers, activation function, learning rate, batch size, and epochs. Both one-layer and two-layer ANNs were investigated. The inspected and optimal values for each ANN are listed in Table 2. To ensure the generalization ability of the model, the epoch was determined to be 300 based on the analyses with different epochs. In addition, the number of neurons in the two hidden layers was set the same for the one-layer ANN to reduce the computational cost. After optimal hyperparameters were obtained, the final models were trained using the entire training set.

3.3. Construction of Transfer Model

As stated above, the transfer learning approach is based on feature extraction and fine-tuning. A source model was first trained to extract meaningful features from a large dataset to achieve a specific task. Then, in fine-tuning, the learned knowledge was transferred by fixing the first several layers of the source model and fine-tuning the final layers of the model to learn the properties of a new smaller dataset for similar tasks. In this paper, the one-layer and two-layer transfer models for the low-cycle fatigue life prediction for BSBs were trained as follows: (1) the hidden layers of the corresponding source models were fixed and were not adjusted in fine training; (2) the weights and biases of the output layer (16 parameters in total) were trained with the training set of BSBs. To fully validate the accuracy of the model, the BSBs dataset was split into a training set (30%) and a testing set (70%). Usually, in typical ML model construction, the size of the training set is larger than that of the testing set to fully extract features. But in the transfer framework, the feature extraction has been completed in the construction of source models. In the construction of transfer models, only a few parameters need to be fine-tuned. Moreover, the reliability of transfer models should be fully validated. Thus, a ratio of 30:70 was adopted in this work. Baseline models are the corresponding one-layer and two-layer ANNs trained directly with the dataset of BSBs without transfer from source models.

4. Results

4.1. Prediction Results for Source Models

Figure 8a,b show the variations in the predicted low-cycle fatigue life of other metallic bars by the one-layer and two-layer ANN source models against the test results, respectively, wherein the results of the training and testing sets are both indicated. From Figure 8, it can be seen that both the one-layer and two-layer ANNs predict the low-cycle fatigue life reasonably well. The three metrics defined in Section 3.1 are furthermore utilized to evaluate the accuracy of predictions.
Table 3 shows the R 2 , RMSE, and MAE values for the two models. From the validation results on the testing set, these two models were capable of predicting the fatigue life with an R 2 value higher than 0.90. The one-layer model gave a slightly better coefficient of determination and smaller values of RMSE and MAE on the testing set, compared with the two-layer model. The above phenomenon is due to the noise introduced by the double-layer model with too many parameters when the dataset is small. The RMSE of the test set for the one-layer model was 38.321, which was due to the dimensionality of the dataset. By far, the ability of the source models to extract the features of low-cycle fatigue life has been validated, which will be leveraged by the transfer model to predict the fatigue life of BSBs based on the limited target dataset.

4.2. Prediction Results for Transfer Models

Figure 9 compares the predicted results for the low-cycle fatigue life of BSBs to the test results for both the one-layer and two-layer transfer models. For the testing set, these two models achieved considerable accuracy, as almost all of the dots were spotted on the line of y = x. Furthermore, the coefficients of determination were both 1.00, as shown in Table 3. For the testing set, the one-layer transfer model attained a coefficient of determination of 0.861, which indicates that the one-layered transfer model can predict the low-cycle fatigue life of BSB with satisfactory accuracy. However, the accuracy of the two-layer transfer model deteriorated significantly on the testing set compared to that of the training set. The value of R 2 decreases from 1.00 to 0.392, which is smaller than the results for the one-layer transfer model. The results in Figure 9b show a relatively higher scatter on the testing set than the results in Figure 9a. Thus, the one-layer ANN model possesses a higher representative ability under the current BSBs dataset.

5. Discussion

5.1. Efficiency of the Transfer Framework

The objective of this paper is to validate the benefit of leveraging the dataset of other metallic bars to predict the low-cycle fatigue life of BSBs, which will remarkably reduce the cost by decreasing the need for data samples of BSBs. To highlight the efficiency and accuracy of the constructed transfer models, as stated above, one- and two-layer baseline models based solely on the dataset of BSBs without transferring the features learned in the source models were also constructed. The low-cycle fatigue life of BSBs predicted by the one-layer and two-layer baseline models against the test results is shown in Figure 10a,b, respectively. Compared with the corresponding results for the transfer models in Figure 9, the results for both the one-layer and two-layer baseline models display significantly larger dispersion. It is noteworthy that the coefficients of determination of these two baseline models were negative, as listed in Table 3, which indicates that the BSB dataset is far from sufficient for fatigue life prediction. However, by leveraging the dataset of other metallic bars, with the same dataset of BSBs, a satisfactory coefficient of determination of 0.861 was achieved using the one-layered transfer model. Thus, the advantage of the transfer model was validated.

5.2. Influence of Key Parameters

Based on the one-layer transfer model, the influences of key parameters on the low-cycle fatigue life of BSBs were studied. Figure 11 shows the evolution of the low-cycle fatigue life of BSBs with variations in the fatigue strain amplitude εa and slenderness ratio λ. It can be seen that as εa and λ increase, the fatigue life decreases significantly. The colored dots in Figure 11 denote test results for the BSBs, which again indicate the accuracy of the transfer model. It should be stated that the low-cycle fatigue tests in the existing literature [15,20,21,22,55] only considered the value of εa within the range of 0.00 and 0.03. By utilizing the constructed model, the fatigue life with εa in the range of 0.03 and 0.05 can be predicted, as shown in Figure 11.

5.3. The Relative Importance of Features on the Fatigue Life

As stated above, the low-cycle fatigue life of metallic bars is influenced by several parameters, and the influence is complicated owing to the coupling effect among different parameters. Figure 12 depicts the influence of εa on Nf based on the collected dataset. In general, with an increase in εa, Nf of different steel bars decreased gradually. However, even for the same εa, there were clear differences in Nf among different steel bars. Regarding the coupling effect, it is difficult to evaluate the relative importance of each parameter using conventional methods. By contrast, ML provides an easier way to evaluate the relative importance of each parameter. Based on ML regression models, the impact of the input parameters on the fatigue life was evaluated using the SHapley Additive exPlanations (SHAP) framework [23,56,57,58,59]. The SHAP value, which is determined by comparing model predictions with and without each feature, is central to the method.
Figure 13 displays the relative importance of the input features on the low-cycle fatigue life of metallic bars. According to the SHAP value, the fatigue strain amplitude εa has the greatest impact on fatigue life, and as the value of εa increases, Nf decreases. The other two parameters that have relatively more significant impact are fy and λ. Similarly, increases in fy and λ tend to decrease Nf.
The interaction between two features, that is, how one feature affects the impact of another feature on the low-cycle fatigue life, can be revealed from the SHAP feature dependence plot. The dependence of λ on εa is shown in Figure 14. The effect of εa on the contribution of λ depends on the value of λ. When λ is relatively large, increasing εa leads to an increase in the SHAP value for λ. However, when λ is comparatively small, increasing εa generally decreases its SHAP value.

5.4. Limitations and Prospects

In order to facilitate the operation and data analysis of the low-cycle fatigue test, some hypotheses were set up. One of the important assumptions is that the load is a tensile and compressive equal amplitude load. Another important assumption is that the reinforcement is consolidated at both ends. In the concrete structure, the vertical reinforcement is constrained by stirrups and concrete, and its two ends can be basically fixed. In some extreme situations, the restraining effect of concrete and stirrups may fail due to concrete crushing or stirrup fracture. Therefore, a variety of boundary conditions can be considered in future studies to fully grasp the low-cycle fatigue properties of BSBs.
The prediction model proposed in this study was based on 54 BSB data points and 264 other steel bar data points. Due to the limited research on low-cycle fatigue of BSBs, there were only 54 BSB data points in the dataset. It is worth noting that strict model testing and validation are crucial to ensure the accuracy and reliability of transfer learning frameworks in fatigue life prediction, which requires sufficient datasets. Therefore, more experimental research should be conducted in the future to enrich the BSB dataset. In addition, a framework based on machine learning and transfer learning is proposed in this study, aiming at achieving efficient prediction on low-cycle fatigue life of BSBs with limited data. When the dataset is too large, the computational cost of transfer learning should not be ignored. Many methods can be attempted in future studies to reduce computational costs, such as removing unimportant neurons or connections to reduce the size and complexity of the model, thereby lowering computational costs. In addition to transfer learning frameworks, traditional regression models and other machine learning models also play important roles in the fields of machine learning and artificial intelligence, each with its unique advantages and disadvantages. The choice of model depends on the specific type of problem, the characteristics of the data, and the requirements of the model. Therefore, it is still necessary to make choices based on the specific situation when solving similar issues.
The applicability of the results in this study may be limited when applied to other types of metal materials or fatigue loading conditions not included in the dataset. Especially for metal materials with significantly different material properties from ordinary steel bars, such as nickel-based alloys, titanium alloys, aluminum alloys, etc. In order to adapt the model to other materials and fatigue loading conditions, it is necessary to supplement it with more fatigue test datasets.

6. Conclusions

A transfer learning framework was proposed in this study, in which a dataset of 264 other metallic bars was leveraged to predict the low-cycle fatigue life of BSBs with a very limited experimental dataset (48 samples). The main conclusions of this study are as follows:
(1) The transfer models accurately predicted the low-cycle fatigue life of BSBs using a very small number of samples. The one-layer transfer model achieved a coefficient of determination of 0.861, which was significantly better than the performance of baseline models that did not transfer knowledge learned from the dataset of other metallic bars.
(2) The constructed source model was capable of accurately predicting the low-cycle fatigue life of seven types of metallic bars collected, thus achieving a unified prediction among various metallic materials.
(3) The influence of the hidden layers of the ANN was considered by comparing the prediction performance of one-layer and two-layer transfer models. The results indicated that the one-layer model outperformed the two-layer model on the current BSB dataset.
(4) Feature analysis revealed that the three parameters of relatively high importance for the low-cycle fatigue life of metallic bars are the fatigue strain amplitude εa, yield strength fy, and slenderness ratio λ.
The proposed transfer framework was validated as an efficient method to predict the low-cycle fatigue life of BSBs quickly and at low cost. In future studies, more experiments will be considered to improve the predictive performance of the model.

Author Contributions

Investigation, F.W. and W.D.; writing—original draft preparation, X.X.; writing—review and editing, N.W. and J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Key Research and Development Program of China (grant number 2021YFC3100300) and the National Natural Science Foundation of China (grant number 52308143, 52108115).

Data Availability Statement

Data available on request.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Experimental results from current investigations.
Table A1. Experimental results from current investigations.
NumberTypeλd (mm)εafy (MPa)fu (MPa)Nf
T-6-0.006304-HRB400 stainless-clad bimetallic steel bar [22]6180.60%486.15 652.52 410
T-6-0.0066180.60%486.15 652.52 387
T-6-0.0066180.60%486.15 652.52 415
T-6-0.0156181.50%486.15 652.52 50
T-6-0.0156181.50%486.15 652.52 42
T-6-0.0156181.50%486.15 652.52 44
T-6-0.0246182.40%486.15 652.52 17
T-6-0.0246182.40%486.15 652.52 18
T-6-0.0246182.40%486.15 652.52 16
T-3-0.013181.00%486.15 652.52 2390
T-3-0.013181.00%486.15 652.52 2001
T-3-0.013181.00%486.15 652.52 2637
T-6-0.016181.00%486.15 652.52 370
T-6-0.016181.00%486.15 652.52 297
T-6-0.016181.00%486.15 652.52 325
T-9-0.019181.00%486.15 652.52 164
T-9-0.019181.00%486.15 652.52 144
T-9-0.019181.00%486.15 652.52 136
T-12-0.0112181.00%486.15 652.52 76
T-12-0.0112181.00%486.15 652.52 73
T-12-0.0112181.00%486.15 652.52 72
T-15-0.0115181.00%486.15 652.52 76
T-15-0.0115181.00%486.15 652.52 71
T-15-0.0115181.00%486.15 652.52 80
T-3-0.023182.00%486.15 652.52 269
T-3-0.023182.00%486.15 652.52 300
T-3-0.023182.00%486.15 652.52 339
T-6-0.026182.00%486.15 652.52 47
T-6-0.026182.00%486.15 652.52 46
T-6-0.026182.00%486.15 652.52 48
T-9-0.029182.00%486.15 652.52 31
T-9-0.029182.00%486.15 652.52 23
T-9-0.029182.00%486.15 652.52 22
T-12-0.0212182.00%486.15 652.52 28
T-12-0.0212182.00%486.15 652.52 25
T-12-0.0212182.00%486.15 652.52 23
T-15-0.0215182.00%486.15 652.52 26
T-15-0.0215182.00%486.15 652.52 34
T-15-0.0215182.00%486.15 652.52 29
T-3-0.033183.00%486.15 652.52 71
T-3-0.033183.00%486.15 652.52 70
T-3-0.033183.00%486.15 652.52 76
T-6-0.036183.00%486.15 652.52 14
T-6-0.036183.00%486.15 652.52 17
T-6-0.036183.00%486.15 652.52 17
T-9-0.039183.00%486.15 652.52 13
T-9-0.039183.00%486.15 652.52 13
T-9-0.039183.00%486.15 652.52 13
T-12-0.0312183.00%486.15 652.52 12
T-12-0.0312183.00%486.15 652.52 14
T-12-0.0312183.00%486.15 652.52 17
T-15-0.0315183.00%486.15 652.52 18
T-15-0.0315183.00%486.15 652.52 15
T-15-0.0315183.00%486.15 652.52 14
12.7-6db-0.01 (1)ASTM A1035 Grade 690 [5]612.71.00%795.51054.6549
12.7-6db-0.01 (2)612.71.00%795.51054.6544
12.7-6db-0.02 (1)612.72.00%795.51054.654
12.7-6db-0.02 (2)612.72.00%795.51054.654
12.7-6db-0.03 (1)612.73.00%795.51054.651
12.7-6db-0.03 (2)612.73.00%795.51054.652
12.7-6db-0.04 (1)612.74.00%795.51054.651
12.7-6db-0.04 (2)612.74.00%795.51054.651
12.7-9db-0.01 (1)912.71.00%795.51054.6530
12.7-9db-0.01 (2)912.71.00%795.51054.6528
12.7-9db-0.02 (1)912.72.00%795.51054.654
12.7-9db-0.02 (2)912.72.00%795.51054.653
12.7-9db-0.03 (1)912.73.00%795.51054.652
12.7-9db-0.03 (2)912.73.00%795.51054.652
12.7-9db-0.04 (1)912.74.00%795.51054.652
12.7-9db-0.04 (2)912.74.00%795.51054.652
12.7-12db-0.01 (1)1212.71.00%795.51054.6515
12.7-12db-0.01 (2)1212.71.00%795.51054.6517
12.7-12db-0.02 (1)1212.72.00%795.51054.654
12.7-12db-0.02 (2)1212.72.00%795.51054.654
12.7-12db-0.03 (1)1212.73.00%795.51054.653
12.7-12db-0.03 (2)1212.73.00%795.51054.652
12.7-12db-0.04 (1)1212.74.00%795.51054.652
12.7-12db-0.04 (2)1212.74.00%795.51054.652
12.7-15db-0.01 (1)1512.71.00%795.51054.6515
12.7-15db-0.01 (2)1512.71.00%795.51054.6514
12.7-15db-0.02 (1)1512.72.00%795.51054.655
12.7-15db-0.02 (2)1512.72.00%795.51054.655
12.7-15db-0.03 (1)1512.73.00%795.51054.653
12.7-15db-0.03 (2)1512.73.00%795.51054.653
12.7-15db-0.04 (1)1512.74.00%795.51054.653
12.7-15db-0.04 (2)1512.74.00%795.51054.653
15.88-6db-0.01 (1)615.881.00%7741017.348
15.88-6db-0.01 (2)615.881.00%7741017.331
15.88-6db-0.02 (1)615.882.00%7741017.33
15.88-6db-0.02 (2)615.882.00%7741017.35
15.88-6db-0.03 (1)615.883.00%7741017.32
15.88-6db-0.03 (2)615.883.00%7741017.31
15.88-6db-0.04 (1)615.884.00%7741017.31
15.88-6db-0.04 (2)615.884.00%7741017.31
15.88-9db-0.01 (1)915.881.00%7741017.327
15.88-9db-0.01 (2)915.881.00%7741017.323
15.88-9db-0.02 (1)915.882.00%7741017.33
15.88-9db-0.02 (2)915.882.00%7741017.34
15.88-9db-0.03 (1)915.883.00%7741017.32
15.88-9db-0.03 (2)915.883.00%7741017.32
15.88-9db-0.04 (1)915.884.00%7741017.32
15.88-9db-0.04 (2)915.884.00%7741017.31
15.88-12db-0.01 (1)1215.881.00%7741017.315
15.88-12db-0.01 (2)1215.881.00%7741017.317
15.88-12db-0.02 (1)1215.882.00%7741017.36
15.88-12db-0.02 (2)1215.882.00%7741017.34
15.88-12db-0.03 (1)1215.883.00%7741017.33
15.88-12db-0.03 (2)1215.883.00%7741017.32
15.88-12db-0.04 (1)1215.884.00%7741017.32
15.88-12db-0.04 (2)1215.884.00%7741017.32
15.88-15db-0.01 (1)1515.881.00%7741017.314
15.88-15db-0.01 (2)1515.881.00%7741017.314
15.88-15db-0.02 (1)1515.882.00%7741017.34
15.88-15db-0.02 (2)1515.882.00%7741017.35
15.88-15db-0.03 (1)1515.883.00%7741017.33
15.88-15db-0.03 (2)1515.883.00%7741017.33
15.88-15db-0.04 (1)1515.884.00%7741017.32
15.88-15db-0.04 (2)1515.884.00%7741017.33
B460-5d-0.01B460 smooth bar [50]5121.00%474.5510.5641226
B460-10d-0.0110121.00%474.5510.564243
B460-12d-0.0112121.00%474.5510.564192
B460-15d-0.0115121.00%474.5510.564140
B460-5d-0.0155121.50%474.5510.564367
B460-8d-0.0158121.50%474.5510.564115
B460-10d-0.01510121.50%474.5510.564100
B460-12d-0.01512121.50%474.5510.56494
B460-15d-0.01515121.50%474.5510.56485
B460-5d-0.025122.00%474.5510.564184
B460-8d-0.028122.00%474.5510.56469
B460-10d-0.0210122.00%474.5510.56464
B460-12d-0.0212122.00%474.5510.56458
B460-15d-0.0215122.00%474.5510.56452
B460-5d-0.035123.00%474.5510.56469
B460-8d-0.038123.00%474.5510.56438
B460-10d-0.0310123.00%474.5510.56437
B460-12d-0.0312123.00%474.5510.56432
B460-15d-0.0315123.00%474.5510.56432
B460-5d-0.045124.00%474.5510.56440
B460-8d-0.048124.00%474.5510.56423
B460-10d-0.0410124.00%474.5510.56422
B460-12d-0.0412124.00%474.5510.56422
B460-15d-0.0415124.00%474.5510.56424
B460-5d-0.055125.00%474.5510.56426
B460-8d-0.058125.00%474.5510.56417
B460-10d-0.0510125.00%474.5510.56417
B460-12d-0.0512125.00%474.5510.56417
B460-15d-0.0515125.00%474.5510.56418
FSC14S2.0R0.1(1)HRB400E/316L stainless steel clad bar [3]6141.00%466.5661175
FSC14S2.0R0.1(2)6141.00%466.5661159
FSC14S4.0R0.1(1)6142.00%466.566124
FSC14S4.0R0.1(2)6142.00%466.566123
FSC14S6.0R0.1(1)6143.00%466.56616
FSC14S6.0R0.1(2)6143.00%466.56616
FSC14S8.0R0.1(1)6144.00%466.56613
FSC14S8.0R0.1(2)6144.00%466.56613
FSC18S2.0R0.1(1)6181.00%474.6645181
FSC18S2.0R0.1(2)6181.00%474.6645183
FSC18S4.0R0.1(1)6182.00%474.664522
FSC18S4.0R0.1(2)6182.00%474.664518
FSC18S6.0R0.1(1)6183.00%474.66459
FSC18S6.0R0.1(2)6183.00%474.66457
FSC18S8.0R0.1(1)6184.00%474.66456
FSC18S8.0R0.1(2)6184.00%474.66455
FSC25S2.0R0.1(1)6251.00%495.5662.1292
FSC25S2.0R0.1(2)6251.00%495.5662.1319
FSC25S4.0R0.1(1)6252.00%495.5662.131
FSC25S4.0R0.1(2)6252.00%495.5662.135
FSC25S6.0R0.1(1)6253.00%495.5662.113
FSC25S6.0R0.1(2)6253.00%495.5662.118
FSC25S8.0R0.1(1)6254.00%495.5662.16
FSC25S8.0R0.1(2)6254.00%495.5662.16
B500B-5d-0.01B500B bar [50]5161.00%535.67 633.75 609
B500B-8d-0.018161.00%535.67 633.75 171
B500B-10d-0.0110161.00%535.67 633.75 92
B500B-12d-0.0112161.00%535.67 633.75 68
B500B-15d-0.0115161.00%535.67 633.75 62
B500B-5d-0.0155161.50%535.67 633.75 162
B500B-8d-0.0158161.50%535.67 633.75 57
B500B-10d-0.01510161.50%535.67 633.75 33
B500B-12d-0.01512161.50%535.67 633.75 32
B500B-15d-0.01515161.50%535.67 633.75 31
B500B-5d-0.025162.00%535.67 633.75 66
B500B-8d-0.028162.00%535.67 633.75 23
B500B-10d-0.0210162.00%535.67 633.75 22
B500B-12d-0.0212162.00%535.67 633.75 25
B500B-15d-0.0215162.00%535.67 633.75 19
B500B-5d-0.0255162.50%535.67 633.75 32
B500B-8d-0.0258162.50%535.67 633.75 15
B500B-10d-0.02510162.50%535.67 633.75 14
B500B-12d-0.02512162.50%535.67 633.75 14
B500B-15d-0.02515162.50%535.67 633.75 16
B500B-5d-0.035163.00%535.67 633.75 24
B500B-8d-0.038163.00%535.67 633.75 11
B500B-10d-0.0310163.00%535.67 633.75 11
B500B-12d-0.0312163.00%535.67 633.75 11
B500B-15d-0.0315163.00%535.67 633.75 13
B500B-5d-0.045164.00%535.67 633.75 13
B500B-8d-0.048164.00%535.67 633.75 7
B500B-10d-0.0410164.00%535.67 633.75 7
B500B-12d-0.0412164.00%535.67 633.75 9
B500B-15d-0.0415164.00%535.67 633.75 9
B500B-5d-0.015121.00%544.33 640.67 467
B500B-8d-0.018121.00%544.33 640.67 167
B500B-10d-0.0110121.00%544.33 640.67 111
B500B-12d-0.0112121.00%544.33 640.67 74
B500B-15d-0.0115121.00%544.33 640.67 72
B500B-5d-0.0155121.50%544.33 640.67 166
B500B-8d-0.0158121.50%544.33 640.67 64
B500B-10d-0.01510121.50%544.33 640.67 50
B500B-12d-0.01512121.50%544.33 640.67 41
B500B-15d-0.01515121.50%544.33 640.67 36
B500B-5d-0.025122.00%544.33 640.67 71
B500B-8d-0.028122.00%544.33 640.67 32
B500B-10d-0.0210122.00%544.33 640.67 26
B500B-12d-0.0212122.00%544.33 640.67 23
B500B-15d-0.0215122.00%544.33 640.67 27
B500B-5d-0.035123.00%544.33 640.67 26
B500B-8d-0.038123.00%544.33 640.67 15
B500B-10d-0.0310123.00%544.33 640.67 13
B500B-12d-0.0312123.00%544.33 640.67 19
B500B-15d-0.0315123.00%544.33 640.67 17
B500B-5d-0.045124.00%544.33 640.67 17
B500B-8d-0.048124.00%544.33 640.67 12
B500B-10d-0.0410124.00%544.33 640.67 9
B500B-12d-0.0412124.00%544.33 640.67 14
B500B-15d-0.0415124.00%544.33 640.67 12
B500B-5d-0.055125.00%544.33 640.67 10
B500B-8d-0.058125.00%544.33 640.67 8
B500B-10d-0.0510125.00%544.33 640.67 8
B500B-12d-0.0512125.00%544.33 640.67 8
B500B-15d-0.0515125.00%544.33 640.67 9
B500B-16-TEMP Prod.1.1B500B bar [51]6162.50%596.6671.419
B500B-16-TEMP Prod.1.16164.00%596.6671.48
B500B-16-TEMP Prod.1.18162.50%596.6671.415
B500B-16-TEMP Prod.1.18164.00%596.6671.410
B500B-16-TEMP Prod.1.26162.50%572668.319
B500B-16-TEMP Prod.1.26164.00%572668.38
B500B-16-TEMP Prod.1.28162.50%572668.315
B500B-16-TEMP Prod.1.28164.00%572668.36
B500B-16-TEMP Prod.1.36162.50%513.1616.319
B500B-16-TEMP Prod.1.36164.00%513.1616.39
B500B-16-TEMP Prod.1.38162.50%513.1616.313
B500B-16-TEMP Prod.1.38164.00%513.1616.38
B500B-16-TEMP Prod.26162.50%516.963520
B500B-16-TEMP Prod.26164.00%516.963511
B500B-16-TEMP Prod.28162.50%516.963514
B500B-16-TEMP Prod.28164.00%516.96357
B500B-20-TEMP Prod.26202.50%515.3621.820
B500B-20-TEMP Prod.26204.00%515.3621.89
B500B-20-TEMP Prod.28202.50%515.3621.816
B500B-20-TEMP Prod.28204.00%515.3621.87
B500B-8-STR Prod.1682.50%565.6619.320
B500B-8-STR Prod.1684.00%565.6619.311
B500B-8-STR Prod.1882.50%565.6619.320
B500B-8-STR Prod.1884.00%565.6619.310
B500B-8-TEMP Prod.2682.50%584.7671.520
B500B-8-TEMP Prod.2684.00%584.7671.511
B500B-8-TEMP Prod.2882.50%584.7671.517
B500B-8-TEMP Prod.2884.00%584.7671.59
B500B-12-STR Prod.16122.50%538.462725
B500B-12-STR Prod.16124.00%538.46278
B500B-12-STR Prod.18122.50%538.462715
B500B-12-STR Prod.18124.00%538.46277
B500A-8-CW Prod.2B500A bar [51]682.50%526.4546.820
B500A-8-CW Prod.2684.00%526.4546.814
B500A-8-CW Prod.2882.50%526.4546.819
B500A-8-CW Prod.2884.00%526.4546.812
B500A-12-CW Prod.26122.50%567.758920
B500A-12-CW Prod.26124.00%567.75898
B500A-12-CW Prod.28122.50%567.758917
B500A-12-CW Prod.28124.00%567.75898
B400C-8-TEMP Prod.1B400C bar [51]682.50%442.9567.320
B400C-8-TEMP Prod.1684.00%442.9567.312
B400C-8-TEMP Prod.1882.50%442.9567.320
B400C-8-TEMP Prod.1884.00%442.9567.312
B400C-16-MA Prod.26162.50%434.5565.320
B400C-16-MA Prod.26164.00%434.5565.312
B400C-16-MA Prod.28162.50%434.5565.317
B400C-16-MA Prod.28164.00%434.5565.38
B400C-20-MA Prod.26202.50%416563.320
B400C-20-MA Prod.26204.00%416563.39
B400C-20-MA Prod.28202.50%416563.318
B400C-20-MA Prod.28204.00%416563.39
B400C-20-TEMP Prod.26202.50%436.2557.220
B400C-20-TEMP Prod.26204.00%436.2557.27
B400C-20-TEMP Prod.28202.50%436.2557.27
B450C-8-STR Prod.1B450C bar [51]682.50%530.2624.820
B450C-8-STR Prod.1684.00%530.2624.815
B450C-8-STR Prod.1882.50%530.2624.820
B450C-8-STR Prod.1884.00%530.2624.816
B450C-12-STR Prod.16122.50%530.2619.628
B450C-12-STR Prod.16124.00%530.2619.69
B450C-12-STR Prod.18122.50%530.2619.618
B450C-12-STR Prod.18124.00%530.2619.68
B450C-12-STR Prod.26122.50%513.8599.720
B450C-12-STR Prod.26124.00%513.8599.714
B450C-12-STR Prod.28122.50%513.8599.720
B450C-12-STR Prod.28124.00%513.8599.712
B450C-16-TEMP Prod.1.16162.50%537.3640.519
B450C-16-TEMP Prod.1.16164.00%537.3640.59
B450C-16-TEMP Prod.1.18162.50%537.3640.513
B450C-16-TEMP Prod.1.18164.00%537.3640.511
B450C-16-TEMP Prod.1.26162.50%446.7542.718
B450C-16-TEMP Prod.1.26164.00%446.7542.718
B450C-16-TEMP Prod.1.28162.50%446.7542.718
B450C-16-TEMP Prod.1.28164.00%446.7542.79
B450C-16-TEMP Prod.1.36162.50%517.8615.419
B450C-16-TEMP Prod.1.36164.00%517.8615.414
B450C-16-TEMP Prod.1.38162.50%517.8615.415
B450C-16-TEMP Prod.1.38164.00%517.8615.49
B450C-16-TEMP Prod.26162.50%479.3601.118
B450C-16-TEMP Prod.26164.00%479.3601.18
B450C-16-TEMP Prod.28162.50%479.3601.118
B450C-16-TEMP Prod.28164.00%479.3601.17
B450C-20-TEMP Prod.26202.50%492.9591.419
B450C-20-TEMP Prod.26204.00%492.9591.47
B450C-20-TEMP Prod.28202.50%492.9591.419
B450C-20-TEMP Prod.28204.00%492.9591.47

References

  1. Wang, N.; Zhou, F.; Xu, H.; Xu, Z. Seismic behaviour of flat connections between steel-plate composite (SC) walls and reinforced concrete (RC) walls. J. Constr. Steel Res. 2020, 168, 105929. [Google Scholar] [CrossRef]
  2. Wang, N.; Zhou, F.; Li, Z.; Xu, Z.; Xu, H. Experimental and numerical investigation on the L-joint composed of steel-plate composite (SC) walls under seismic loading. Eng. Struct. 2021, 227, 111360. [Google Scholar] [CrossRef]
  3. Li, W.; Wang, Q.; Qu, H.; Xiang, Y.; Li, Z. Mechanical properties of HRB400E/316L stainless steel clad rebar under low-cycle fatigue. Structures 2022, 38, 292–305. [Google Scholar] [CrossRef]
  4. Deng, Z.; Huang, S.; Wang, Y.; Xue, H. Experimental research on fatigue behavior of prestressed ultra-high performance concrete beams with high-strength steel bars. Structures 2022, 43, 1778–1789. [Google Scholar] [CrossRef]
  5. Aldabagh, S.; Alam, M.S. Low-cycle fatigue performance of high-strength steel reinforcing bars considering the effect of inelastic buckling. Eng. Struct. 2021, 235, 112114. [Google Scholar] [CrossRef]
  6. Apostolopoulos, C.A. Mechanical behavior of corroded reinforcing steel bars S500s tempcore under low cycle fatigue. Constr. Build. Mater. 2007, 21, 1447–1456. [Google Scholar] [CrossRef]
  7. Hawileh, R.; Rahman, A.; Tabatabai, H. Evaluation of the Low-Cycle Fatigue Life in ASTM A706 and A615 Grade 60 Steel Reinforcing Bars. J. Mater. Civ. Eng. 2010, 22, 65–76. [Google Scholar] [CrossRef]
  8. Kashani, M.M.; Alagheband, P.; Khan, R.; Davis, S. Impact of corrosion on low-cycle fatigue degradation of reinforcing bars with the effect of inelastic buckling. Int. J. Fatigue 2015, 77, 174–185. [Google Scholar] [CrossRef]
  9. Hua, J.; Wang, F.; Huang, L.; Wang, N.; Xue, X. Experimental study on mechanical properties of corroded stainless-clad bimetallic steel bars. Constr. Build. Mater. 2021, 287, 123019. [Google Scholar] [CrossRef]
  10. Hua, J.; Wang, F.; Wang, N.; Huang, L.; Hai, L.; Li, Y.; Zhu, X.; Xue, X. Experimental and numerical investigations on corroded stainless-clad bimetallic steel bar with artificial damage. J. Build. Eng. 2021, 44, 102779. [Google Scholar] [CrossRef]
  11. Shi, Y.; Luo, Z.; Zhou, X.; Xue, X.; Li, J. Post-fire mechanical properties of titanium–clad bimetallic steel in different cooling approaches. J. Constr. Steel Res. 2022, 191, 107169. [Google Scholar] [CrossRef]
  12. Shi, Y.; Luo, Z.; Zhou, X.; Xue, X.; Xiang, Y. Post-fire performance of bonding interface in explosion-welded stainless-clad bimetallic steel. J. Constr. Steel Res. 2022, 193, 107285. [Google Scholar] [CrossRef]
  13. Xue, X.; Hua, J.; Wang, F.; Wang, N.; Li, S. Mechanical Property Model of Q620 High-Strength Steel with Corrosion Effects. Buildings 2022, 12, 1651. [Google Scholar] [CrossRef]
  14. He, Z.; Shen, A.; Wang, W.; Zuo, X.; Wu, J. Evaluation and optimization of various treatment methods for enhancing the properties of brick-concrete recycled coarse aggregate. J. Adhes. Sci. Technol. 2022, 36, 1060–1080. [Google Scholar] [CrossRef]
  15. Hua, J.; Wang, F.; Yang, Z.; Xue, X.; Huang, L.; Chen, Z. Low-cycle fatigue properties of bimetallic steel bars after exposure to elevated temperature. J. Constr. Steel Res. 2021, 187, 106959. [Google Scholar] [CrossRef]
  16. Hua, J.; Wang, F.; Xiang, Y.; Yang, Z.; Xue, X.; Huang, L.; Wang, N. Mechanical properties of stainless-clad bimetallic steel bars exposed to elevated temperatures. Fire Saf. J. 2022, 127, 103521. [Google Scholar] [CrossRef]
  17. Hua, J.; Yang, Z.; Wang, F.; Xue, X.; Wang, N.; Huang, L. Relation between the metallographic structure and mechanical properties of a bimetallic steel bar after fire. J. Mater. Civ. Eng. 2022, 34, 04022193. [Google Scholar] [CrossRef]
  18. Hua, J.; Yang, Z.; Xue, X.; Huang, L.; Wang, N.; Chen, Z. Bond properties of bimetallic steel bar in seawater sea-sand concrete at different ages. Constr. Build. Mater. 2022, 323, 126539. [Google Scholar] [CrossRef]
  19. Liu, G.; Hua, J.; Wang, N.; Deng, W.; Xue, X. Material Alternatives for Concrete Structures on Remote Islands: Based on Life-cycle cost Analysis. Adv. Civ. Eng. 2022, 2022, 7329408. [Google Scholar] [CrossRef]
  20. Hua, J.; Fan, H.; Yan, W.; Wang, N.; Xue, X.; Huang, L. Seismic resistance of the corroded bimetallic steel bar under different strain amplitudes. Constr. Build. Mater. 2022, 319, 126088. [Google Scholar] [CrossRef]
  21. Hua, J.; Wang, F.; Xue, X.; Fan, H.; Yan, W. Fatigue properties of bimetallic steel bar: An experimental and numerical study. Eng. Fail. Anal. 2022, 136, 106212. [Google Scholar] [CrossRef]
  22. Wang, F.; Hua, J.; Xue, X.; Ding, Z.; Lyu, Y.; Liu, Q. Low-cycle fatigue performance of bimetallic steel bar considering the effect of inelastic buckling. Constr. Build. Mater. 2022, 351, 128787. [Google Scholar] [CrossRef]
  23. Xiao, L.; Li, Q.Y.; Li, H.; Ren, Q. Loading capacity prediction and optimization of cold-formed steel built-up section columns based on machine learning methods. Thin-Walled Struct. 2022, 180, 109826. [Google Scholar] [CrossRef]
  24. Xiao, L.; Hua, J.; Li, H.; Xue, X.; Wang, N.; Wang, F. Quantitative analysis on post–fire–resistant performance of high–strength steel plate girders using LSTM. J. Constr. Steel Res. 2022, 199, 107588. [Google Scholar] [CrossRef]
  25. Feng, D.-C.; Li, Y.; Shafieezadeh, A.; Taciroglu, E. Advances in Data-Driven Risk-Based Performance Assessment of Structures and Infrastructure Systems. J. Struct. Eng. 2023, 149, 2022–2023. [Google Scholar] [CrossRef]
  26. Cardellicchio, A.; Ruggieri, S.; Nettis, A.; Renò, V.; Uva, G. Physical interpretation of machine learning-based recognition of defects for the risk management of existing bridge heritage. Eng. Fail. Anal. 2023, 149, 107237. [Google Scholar] [CrossRef]
  27. Jafari, A.; Ma, L.Y.; Shahmansouri, A.A.; Dugnani, R. Quantitative fractography for brittle fracture via multilayer perceptron neural network. Eng. Fract. Mech. 2023, 291, 109545. [Google Scholar] [CrossRef]
  28. Bao, H.; Wu, S.; Wu, Z.; Kang, G.; Peng, X.; Withers, P.J. A machine-learning fatigue life prediction approach of additively manufactured metals. Eng. Fract. Mech. 2021, 242, 107508. [Google Scholar] [CrossRef]
  29. Yang, J.; Kang, G.; Liu, Y.; Chen, K.; Kan, Q. Life prediction for rate-dependent low-cycle fatigue of PA6 polymer considering ratchetting: Semi-empirical model and neural network based approach. Int. J. Fatigue 2020, 136, 105619. [Google Scholar] [CrossRef]
  30. Srinivasan, V.S.; Valsan, M.; Rao, K.B.S.; Mannan, S.L.; Raj, B. Low cycle fatigue and creep–fatigue interaction behavior of 316L(N) stainless steel and life prediction by artificial neural network approach. Int. J. Fatigue 2003, 25, 1327–1338. [Google Scholar] [CrossRef]
  31. Vassilopoulos, A.P.; Georgopoulos, E.F.; Dionysopoulos, V. Artificial neural networks in spectrum fatigue life prediction of composite materials. Int. J. Fatigue 2007, 29, 20–29. [Google Scholar] [CrossRef]
  32. Zhan, Z.; Li, H. Machine learning based fatigue life prediction with effects of additive manufacturing process parameters for printed SS 316L. Int. J. Fatigue 2021, 142, 105941. [Google Scholar] [CrossRef]
  33. Zhan, Z.; Li, H. A novel approach based on the elastoplastic fatigue damage and machine learning models for life prediction of aerospace alloy parts fabricated by additive manufacturing. Int. J. Fatigue 2021, 145, 106089. [Google Scholar] [CrossRef]
  34. He, L.; Wang, Z.; Akebono, H.; Sugeta, A. Machine learning-based predictions of fatigue life and fatigue limit for steels. J. Mater. Sci. Technol. 2021, 90, 9–19. [Google Scholar] [CrossRef]
  35. Zhang, X.-C.; Gong, J.-G.; Xuan, F.-Z. A deep learning based life prediction method for components under creep, fatigue and creep-fatigue conditions. Int. J. Fatigue 2021, 148, 106236. [Google Scholar] [CrossRef]
  36. Yang, J.; Kang, G.; Liu, Y.; Kan, Q. A novel method of multiaxial fatigue life prediction based on deep learning. Int. J. Fatigue 2021, 151, 106356. [Google Scholar] [CrossRef]
  37. Kamiyama, M.; Shimizu, K.; Akiniwa, Y. Prediction of low-cycle fatigue crack development of sputtered Cu thin film using deep convolutional neural network. Int. J. Fatigue 2022, 162, 106998. [Google Scholar] [CrossRef]
  38. Chen, J.; Liu, Y. Probabilistic physics-guided machine learning for fatigue data analysis. Expert Syst. Appl. 2021, 168, 114316. [Google Scholar] [CrossRef]
  39. Guo, X.; Liu, X.; Long, G.; Zhao, Y.; Yuan, Y. Data-driven prediction of the fatigue performance of corroded high-strength steel wires. Eng. Fail. Anal. 2023, 146, 107108. [Google Scholar] [CrossRef]
  40. Pan, S.J.; Yang, Q. A Survey on Transfer Learning. IEEE Trans. Knowl. Data Eng. 2010, 22, 1345–1359. [Google Scholar] [CrossRef]
  41. Weiss, K.; Khoshgoftaar, T.M.; Wang, D.D. A Survey of Transfer Learning; Springer International Publishing: Berlin/Heidelberg, Germany, 2016. [Google Scholar] [CrossRef]
  42. Kaya, A.; Keceli, A.S.; Catal, C.; Yalic, H.Y.; Temucin, H.; Tekinerdogan, B. Analysis of transfer learning for deep neural network based plant classification models. Comput. Electron. Agric. 2019, 158, 20–29. [Google Scholar] [CrossRef]
  43. He, R.; Tian, Z.G.; Zuo, M.J. A transferable neural network method for remaining useful life prediction. Mech. Syst. Signal Process. 2023, 183, 109608. [Google Scholar] [CrossRef]
  44. Yin, T.; Lu, N.; Guo, G.S.; Lei, Y.G.; Wang, S.H.; Guan, X.H. Knowledge and data dual-driven transfer network for industrial robot fault diagnosis. Mech. Syst. Signal Process. 2023, 182, 109597. [Google Scholar] [CrossRef]
  45. Wei, X.; Zhang, C.; Han, S.; Jia, Z.; Wang, C.; Xu, W. High cycle fatigue S-N curve prediction of steels based on transfer learning guided long short term memory network. Int. J. Fatigue 2022, 163, 107050. [Google Scholar] [CrossRef]
  46. Kashani, M.M.; Crewe, A.J.; Alexander, N.A. Nonlinear stress-strain behaviour of corrosion-damaged reinforcing bars including inelastic buckling. Eng. Struct. 2013, 48, 417–429. [Google Scholar] [CrossRef]
  47. Tripathi, M.; Dhakal, R.P.; Dashti, F.; Massone, L.M. Low-cycle fatigue behaviour of reinforcing bars including the effect of inelastic buckling. Constr. Build. Mater. 2018, 190, 1226–1235. [Google Scholar] [CrossRef]
  48. Hu, M.; Han, Q.; Xu, K.; Du, X. Corrosion influences on monotonic properties of ultra-high-strength reinforcing steels. Constr. Build. Mater. 2019, 198, 82–97. [Google Scholar] [CrossRef]
  49. Hua, J.; Wang, F.; Xue, X.; Ding, Z.; Chen, Z. Residual monotonic mechanical properties of bimetallic steel bar with fatigue damage. J. Build. Eng. 2022, 55, 104703. [Google Scholar] [CrossRef]
  50. Kashani, M.M.; Barmi, A.K.; Malinova, V.S. Influence of inelastic buckling on low-cycle fatigue degradation of reinforcing bars. Constr. Build. Mater. 2015, 94, 644–655. [Google Scholar] [CrossRef]
  51. Caprili, S.; Salvatore, W. Cyclic behaviour of uncorroded and corroded steel reinforcing bars. Constr. Build. Mater. 2015, 76, 168–186. [Google Scholar] [CrossRef]
  52. Hua, J.; Yang, Z.; Zhou, F.; Hai, L.; Wang, N.; Wang, F. Effects of exposure temperature on low–cycle fatigue properties of Q690 high–strength steel. J. Constr. Steel Res. 2022, 190, 107159. [Google Scholar] [CrossRef]
  53. Bolón-Canedo, V.; Remeseiro, B. Feature selection in image analysis: A survey. Artif. Intell. Rev. 2020, 53, 2905–2931. [Google Scholar] [CrossRef]
  54. Kabir, H.; Garg, N. Machine learning enabled orthogonal camera goniometry for accurate and robust contact angle measurements. Sci. Rep. 2023, 13, 1497. [Google Scholar] [CrossRef] [PubMed]
  55. Hua, J.; Wang, F.; Xue, X. Study on fatigue properties of post-fire bimetallic steel bar with different cooling methods. Structures 2022, 40, 633–645. [Google Scholar] [CrossRef]
  56. Mangalathu, S.; Karthikeyan, K.; Feng, D.C.; Jeon, J.S. Machine-learning interpretability techniques for seismic performance assessment of infrastructure systems. Eng. Struct. 2022, 250, 112883. [Google Scholar] [CrossRef]
  57. Lundberg, S.M.; Lee, S.I. A Unified Approach to Interpreting Model Predictions. In Proceedings of the Advances in Neural Information Processing Systems 30 (NIPS 2017), Long Beach, CA, USA, 4–9 December 2017; Volume 30. [Google Scholar]
  58. Shen, Y.; Wu, L.; Liang, S. Explainable machine learning-based model for failure mode identification of RC flat slabs without transverse reinforcement. Eng. Fail. Anal. 2022, 141, 106647. [Google Scholar] [CrossRef]
  59. Gao, X.; Lin, C. Prediction model of the failure mode of beam-column joints using machine learning methods. Eng. Fail. Anal. 2021, 120, 105072. [Google Scholar] [CrossRef]
Figure 1. Details of a BSB specimen.
Figure 1. Details of a BSB specimen.
Buildings 14 02275 g001
Figure 2. Deformation mode of longitudinal steel bar in RC column base.
Figure 2. Deformation mode of longitudinal steel bar in RC column base.
Buildings 14 02275 g002
Figure 3. MTS system with a steel bar installed, and test specimens of steel bars used in low-cycle fatigue tests.
Figure 3. MTS system with a steel bar installed, and test specimens of steel bars used in low-cycle fatigue tests.
Buildings 14 02275 g003
Figure 4. Inelastic buckling phenomenon.
Figure 4. Inelastic buckling phenomenon.
Buildings 14 02275 g004
Figure 5. Statistical distributions of variables: (a) fy (MPa), (b) fu (MPa), (c) fy/fu, (d) d (mm), (e) εa, (f) λ, and (g) Nf. Note: fy denotes the yield strength (MPa) of specimens; fu denotes the ultimate strength (MPa) of specimens; d denotes the diameter (mm) of specimens; εa denotes the strain amplitude; λ denotes the slenderness ratio of specimens; Nf denotes the fatigue life of specimens.
Figure 5. Statistical distributions of variables: (a) fy (MPa), (b) fu (MPa), (c) fy/fu, (d) d (mm), (e) εa, (f) λ, and (g) Nf. Note: fy denotes the yield strength (MPa) of specimens; fu denotes the ultimate strength (MPa) of specimens; d denotes the diameter (mm) of specimens; εa denotes the strain amplitude; λ denotes the slenderness ratio of specimens; Nf denotes the fatigue life of specimens.
Buildings 14 02275 g005
Figure 6. Framework of transfer learning to predict low-cycle fatigue life.
Figure 6. Framework of transfer learning to predict low-cycle fatigue life.
Buildings 14 02275 g006
Figure 7. Structure of an ANN with two hidden layers for the prediction of low-cycle (LC) fatigue life.
Figure 7. Structure of an ANN with two hidden layers for the prediction of low-cycle (LC) fatigue life.
Buildings 14 02275 g007
Figure 8. Low-cycle fatigue life of other metallic bars from source models: (a) one-layer and (b) two-layer ANNs against the test results.
Figure 8. Low-cycle fatigue life of other metallic bars from source models: (a) one-layer and (b) two-layer ANNs against the test results.
Buildings 14 02275 g008
Figure 9. Low-cycle fatigue life of BSBs predicted by transfer models: (a) one-layer and (b) two-layer ANNs against the test results.
Figure 9. Low-cycle fatigue life of BSBs predicted by transfer models: (a) one-layer and (b) two-layer ANNs against the test results.
Buildings 14 02275 g009
Figure 10. Low-cycle fatigue life of BSBs predicted by baseline models: (a) one-layer and (b) two-layer ANNs against the test results.
Figure 10. Low-cycle fatigue life of BSBs predicted by baseline models: (a) one-layer and (b) two-layer ANNs against the test results.
Buildings 14 02275 g010
Figure 11. Influence of the fatigue strain ratio and slenderness ratio on the low-cycle fatigue life of BSBs.
Figure 11. Influence of the fatigue strain ratio and slenderness ratio on the low-cycle fatigue life of BSBs.
Buildings 14 02275 g011
Figure 12. Relation between fatigue strain amplitude and cycles to failure of different steel bars [3,5,22,50,51].
Figure 12. Relation between fatigue strain amplitude and cycles to failure of different steel bars [3,5,22,50,51].
Buildings 14 02275 g012
Figure 13. Relative importance of the input features on the low-cycle fatigue life of metallic bars. The value of each input feature was separately normalized in the range of zero and one. Note: fy denotes the yield strength (MPa) of specimens; fu denotes the ultimate strength (MPa) of specimens; d denotes the diameter (mm) of specimens; εa denotes the strain amplitude; λ denotes the slenderness ratio of specimens; Nf denotes the fatigue life of specimens.
Figure 13. Relative importance of the input features on the low-cycle fatigue life of metallic bars. The value of each input feature was separately normalized in the range of zero and one. Note: fy denotes the yield strength (MPa) of specimens; fu denotes the ultimate strength (MPa) of specimens; d denotes the diameter (mm) of specimens; εa denotes the strain amplitude; λ denotes the slenderness ratio of specimens; Nf denotes the fatigue life of specimens.
Buildings 14 02275 g013
Figure 14. Dependence of slenderness ratio on fatigue strain amplitude.
Figure 14. Dependence of slenderness ratio on fatigue strain amplitude.
Buildings 14 02275 g014
Table 1. Experimental results from existing research [3,5,22,50,51].
Table 1. Experimental results from existing research [3,5,22,50,51].
GroupTypeNumber of
Samples
Nf
1Other metallic bars (264 samples in total) ASTM A1035 Grade 690641~49
2B460 smooth bar2917~1226
3HRB400E/316L stainless steel clad bar243~319
4B500B bar926~171
5B500A bar88~20
6B400C bar157~20
7B450C bar327~20
8BSB304-HRB400 stainless-clad BSB5412~2637
Table 2. Hyperparameters in source model.
Table 2. Hyperparameters in source model.
HyperparameterInspected ValuesOptimal Value for One-Layer ANNOptimal Value for Two-Layer ANN
Number of neurons in hidden layers7, 9, 11, 13, and 151515
Activation functionRelu, Tanh and SigmoidReluRelu
Learning rate0.0001, 0.001, and 0.010.010.001
Batch size16 and 321616
Epochs100, 200 and 300300300
Table 3. Performance metrics of one-layer and two-layer source models, transfer models and corresponding baseline models.
Table 3. Performance metrics of one-layer and two-layer source models, transfer models and corresponding baseline models.
ModelSetOne-Layer Model Two-Layer Model
R 2 RMSEMAE R 2 RMSEMAE
Source model Training0.959 21.130 6.784 0.979 15.101 5.060
Testing0.911 32.231 11.700 0.906 33.023 13.882
Transfer modelTraining1.000 5.121 2.290 1.000 0.278 0.125
Testing0.861 38.321 19.820 0.392 80.143 39.832
Baseline modelTraining0.919 253.032 163.764 0.831 365.041 214.744
Testing−0.976 144.467 99.245 −0.113 108.389 74.993
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

Xue, X.; Wang, F.; Wang, N.; Hua, J.; Deng, W. Transfer-Learning Prediction Model for Low-Cycle Fatigue Life of Bimetallic Steel Bars. Buildings 2024, 14, 2275. https://doi.org/10.3390/buildings14082275

AMA Style

Xue X, Wang F, Wang N, Hua J, Deng W. Transfer-Learning Prediction Model for Low-Cycle Fatigue Life of Bimetallic Steel Bars. Buildings. 2024; 14(8):2275. https://doi.org/10.3390/buildings14082275

Chicago/Turabian Style

Xue, Xuanyi, Fei Wang, Neng Wang, Jianmin Hua, and Wenjie Deng. 2024. "Transfer-Learning Prediction Model for Low-Cycle Fatigue Life of Bimetallic Steel Bars" Buildings 14, no. 8: 2275. https://doi.org/10.3390/buildings14082275

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