Next Article in Journal
Flexural Behavior of Galvanized Iron Based Cold-Formed Steel Back-to-Back Built-Up Beams at Elevated Temperatures
Previous Article in Journal
Research on Prediction of Excavation Parameters for Deep Buried Tunnel Boring Machine Based on Convolutional Neural Network-Long Short-Term Memory Model
Previous Article in Special Issue
Optimization and Modeling of 7-Day Ultra-High-Performance Concrete Comprising Desert Sand and Supplementary Cementitious Materials Using Response Surface Methodology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Data-Driven Prediction of Electrical Resistivity of Graphene Oxide/Cement Composites Considering the Effects of Specimen Size and Measurement Method

1
College of Civil Engineering, Nanjing Tech University, Nanjing 211816, China
2
Intelligent Civil Engineering Materials and Structures Laboratory, Nanjing Tech University, Nanjing 211816, China
*
Author to whom correspondence should be addressed.
Buildings 2024, 14(8), 2455; https://doi.org/10.3390/buildings14082455
Submission received: 5 July 2024 / Revised: 3 August 2024 / Accepted: 6 August 2024 / Published: 8 August 2024
(This article belongs to the Special Issue Advances in Modeling and Characterization of Cementitious Composites)

Abstract

:
The prediction of electrical resistivity of graphene oxide (GO) reinforced cement composites (GORCCs) is essential to promote the application of the composites in civil engineering. Traditional experiments find it challenging to capture the effect of various features on the electrical resistivity of the GORCCs. In this work, machine learning (ML) techniques are employed to explore the complex nonlinear relationships between different influencing factors and the electrical resistivity of the GORCCs. A total of 171 datasets are utilized for training and testing the ML models. It is demonstrated that the applied ML models are effective and efficient. Apart from the water/cement ratio, correlation analysis shows that the electrical resistivity of the GORCCs is highly dependent on the specimen size and measurement method. Feature importance analysis shows that the dispersion of GO has a significant influence on the electrical resistivity. The extreme gradient boosting (XGB) model and the artificial neural network (ANN) model with 3 hidden layers are proven to have better predictions, as evidenced by higher R2 and lower root mean square error (RMSE). This work is envisioned to provide an effective and efficient way to identify the complex relationship between the material properties of the GORCCs and the various influencing factors.

1. Introduction

Cement composites have become the world’s most widely used building materials due to their versatility, durability, and cost-effectiveness [1]. They are integrated into buildings, bridges, roads, and other infrastructure projects [2]. However, traditional cement composites with structural functions have fallen short of meeting the contemporary demands for developing construction applications with multifunction [3,4,5]. Researchers have explored different methods to enhance the material properties of cement composites (i.e., compressive strength, flexural strength, electrical properties, thermal properties, etc.) by incorporating various nano-fillers into the cement matrix [6,7,8,9,10,11]. Among the numerous nano-fillers, graphene oxide (GO) [12,13,14], as a graphene derivative, has become a promising reinforcement candidate in the cement matrix due to its improved dispersion, excellent mechanical and physical properties, and low manufacturing cost [13,15,16,17].
Numerous works have been carried out on GO reinforced cement composites (GORCCs). Li et al. [18] added GO to cement and measured the electrical resistivity of the GORCCs by the two-electrode method. They found that the electrical resistivity of the cement pastes initially increased, reaching a peak at 0.02 wt% GO, and then decreased as the GO content further increased to 0.04 wt%. The electrical resistivity of the cement composites with well-dispersed GO was much lower than that of the cement matrix. Sartipi et al. [19] studied the electrical properties of the GORCCs, and the volumetric conductivity of the samples was tested at 7, 14, and 28 days. It was shown that the addition of GO increased the electrical conductivity of the samples. Qureshi et al. [17] found that the electrical resistivity of the cement composites with GO and rGO increased with the increase in hydration time. They concluded that the added nanofillers participated in the cement hydration process, which led to increased C-S-H gel pores in the composite, reduced mean free paths for electron transmission, and increased electrical resistivity. Guo et al. [20] found that the electrical resistivity of the GORCCs varied when the moisture in the cement matrix changed. Their investigation revealed that the electrical resistivity reached its minimum value when the GO content was 0.05 wt% and the water/cement ratio was 0.35. It was observed by Hu et al. [21] that when the GO content was 0.03 wt%, the electrical conductivity of the cement composites was nearly 50% higher than that of the plain sample. They found that an appropriate GO content resulted in a denser and more regular microstructure in the cement, which contributed to the enhanced electrical conductivity. Recently, Yan et al. [22] found that GO nanosheets and rGO could be highly aligned in cement pastes under an external magnetic field. The mechanical and electrical properties were found to be enhanced compared to the cement paste with a random distribution of the fillers. Xiong et al. [23] invented a novel power ultrasound (PUS) method for optimizing the dispersion of GO in cement paste. The synergistic effects of PUS-assisted mixing and PCE resulted in improved dispersion of the GO fillers.
Although numerous experiments have been conducted on the GORCCs, it is challenging to identify an explicit expression between the electrical properties and the many influencing factors such as w/c, content of GO, curing temperature, etc. [24]. In addition, during fabrication, the dispersion and measurement methods may also have potential impacts on the final results [15,18]. Recently, machine learning (ML) has attracted considerable attention in the field of composites due to its powerful data analysis capabilities as well as high efficiency without the need for conducting massive experiments [25,26,27]. Machine learning can consider various characteristics of different experiments in its model by collecting a large amount of data. In addition, machine learning is able to reflect the extent to which individual features affect the final result, as well as the complex nonlinear relationships between the involved features. Huang et al. [28] predicted the compressive and flexural strength of carbon nanotube (CNT) reinforced cement composites via ML methods, i.e., artificial neural network (ANN) and support vector machine (SVM). They found that ANN performed better on small datasets, whereas SVM was more suitable for large datasets. Montazerian et al. [29] investigated the thermal conductivity of graphene reinforced cement composites through a hybrid modeling approach employing ML and computational micromechanics models. Their findings indicated that the ANN demonstrated superior prediction accuracy and generalization ability compared to SVM and other models employed in their study. Bang et al. [30] used different ML techniques to evaluate the piezoelectric properties of CNT reinforced cement composites. They concluded that the type of loading, water/cement ratio (w/c), and CNT content were the influencing factors for the electrical resistance of CNT reinforced cement composites. Kekez et al. [31] investigated the application of ANN to predict the compressive and flexural strength of CNT reinforced cement composites. The regression coefficient of the optimal ANN was 0.94. Li et al. [32] employed linear regression (LR), support vector regression (SVR), random forest (RF), and extreme gradient boosting (XGB) to predict the compressive strength of cement paste, mortar, and concrete. The findings revealed that the optimal length and diameter of CNT were 20 μm and 25 nm, respectively, and the appropriate range of CNT content should be within 0.1 wt%. Alyousef et al. [33] utilized the GEP, ANN, and ANFIS models to predict the electrical properties of graphene reinforced cement composites. They found that the incorporation of 0.1% GNP content resulted in a notable 16.58% increase in strength and a 14% decrease in fractional change in resistivity. In addition, the R-value of the three models employed was 0.974, 0.963, and 0.954, respectively.
Although various ML models have been widely used in the field of composites, limited studies applied ML techniques to estimate the electrical properties of the GORCCs, which are intricately related to many features/influencing factors [34,35,36]. Under such circumstances, it is challenging and time-consuming to conduct massive experiments and establish an explicit expression to describe the nuanced relationships between the electrical properties and the many features. Additionally, the existing investigations neglect the effects of the specimen size as a crucial factor despite its evident influences. Based on the identified research gap, this work is the first attempt to employ ML techniques to investigate the electrical resistivity of the GORCCs considering the sample size as an input feature. Moreover, from the perspective of manufacturing, the dimensions of GO typically conform to a Gaussian distribution. Therefore, this study adopts a Gaussian function to characterize the distribution of the GO dimensions instead of using fixed values. Three different ML models are employed to investigate the complex nonlinear relationship between the electrical resistivity of the GORCCs and the coupling effects of the influencing factors. To train the ML model, this study collects 171 sets of experimental data for the electrical resistivity of the GORCCs. For discontinuous features, such as dispersion methods, cement type, etc., the one-hot encoding method is employed for feature input. Particularly, two parameters, i.e., Ks and Kh, are introduced to represent the dimensions of the specimens. Correlation analysis is conducted to examine the intrinsic relationships between the electrical resistivity of the GORCCs and the individual features, which are quantified through the Pearson correlation coefficient. Four evaluation metrics, i.e., R2, mean absolute error (MAE), root mean square error (RMSE), and mean absolute percentage error (MAPE), are adopted to evaluate the prediction performance of the ML models.

2. Data Collection and Pre-Processing

Data collection and pre-processing are crucial for the accurate prediction of the ML models. In this study, a total of 171 experimental datasets were collected from the previous literature. In this work, the criteria for data selection include: (1) The source of experimental data should be published papers or technical reports in authoritative journals, and the results must be rigorously tested against specified requirements. (2) The data collected in the study should ensure diversity and be able to adequately reflect the effect of influencing factors on the GORCCs. The study must consider the w/c, type of cement, curing temperature, curing age, method of dispersion, method of testing, sample size, content of GO, and attributes of GO (i.e., diameter, number of layers, and thickness). (3) The data collected in this study can be normalized into a uniform format to facilitate subsequent data processing and the training of the ML models. Details of the 171 samples can be found in Appendix A Table A1.
Several traditional data pre-processing methods were used in this study. For example, for features such as cement type, dispersion method, measurement method, etc., one-hot encoding was used, where all types of features are converted to binary 0 and 1 for feature input. For cement types, 1 and 2 represent Type I 42.5 Ordinary Portland Cement and Type II 42.5 Ordinary Portland Cement, respectively. These two types of cement are the most commonly used in practical construction and in research as well. The selected curing temperature range was from 18 °C to 23 °C, which represents the typical laboratory conditions for curing cement composites. For the dispersion method, 1 represents mixing after ultrasonication, and 2 represents magnetic mixing followed by ultrasonication. For the measurement method, 1 and 2 represent the four-electrode method and the two-electrode method, respectively. The dimensions of the specimen were normalized according to 50 mm (length) × 50 mm (width) × 50 mm (height), where two dimensionless parameters, Ks and Kh, were introduced. Ks is the ratio of the bottom area of the specimen to the area of 50 mm × 50 mm, and Kh is the ratio of the height of the specimen to the height of the standard specimen, i.e., 50 mm. For the dimensions of GO, from the perspective of manufacturing, the dimensions of GO adhere to a Gaussian distribution. In this study, Gaussian random numbers within the upper and lower limits of the interval were employed to characterize the dimensions of GO, including diameter, thickness, and number of layers. In the data collected in this work, the dispersant used in most studies was PCE. Therefore, the dispersant was not considered as a separate feature in this work. Table 1 presents a description of the collected data. The 171 collected data points were partitioned into 0.8 and 0.2, resulting in 136 training sets and 35 test sets, respectively.

3. Methodology

Different ML models have been developed and used in various fields [37]. Among the models, ANN has demonstrated an excellent ability to deal with complex nonlinear problems [38]. RF also exhibits capabilities in handling nonlinear problems [39,40] and mitigates overfitting issues by eliminating datasets featuring noise and outliers [41]. Similar to the RF model, another robust ensemble learning algorithm is XGB. Compared to RF, XGB employs a gradient boosting algorithm, which trains the decision tree iteratively. For each iteration, the model is adjusted based on the results of the previous iteration. XGB introduces a regularization term in the objective function, which is used to optimize the residual using gradient boosting [42,43]. The three models mentioned above are employed for the prediction of the electrical resistivity of the GORCCs in this work.

3.1. Random Forest

Random forest (RF) is an ensemble learning algorithm composed of decision trees. A decision tree is a model in which each node represents a feature, each branch denotes a possible value of this feature, and each leaf represents an output variable. RF combines several different decision trees, each of which generates a result. Then RF generates the final result by “voting”. For regression problems, this “voting” is the averaging of the values. The architecture of the RF model is shown in Figure 1. The loss function for the regression tree is
L oss = i C 1 ( j , s ) a i y 1 2 + i C 2 ( j , s ) a i y 2 2
where ai represents the actual value, y1 and y2 are the predicted values, C1 and C2 denote the two split subtrees, and j and s stand for splitting variable and point, respectively.
The randomness of RF is mainly reflected in two aspects [44], i.e., randomness of data sampling and feature selection. RF creates different training datasets by the bootstrapping method. This is accomplished by randomly selecting samples from the original dataset with replacement. The introduction of randomness helps reduce the variance of the model, preventing a particular feature from dominating the entire model due to the increased variability of each tree.

3.2. Extreme Gradient Boosting

Similar to the RF algorithm, extreme gradient boosting (XGB) is an ensemble learning algorithm [45], which demonstrates a strong learning ability when dealing with complex nonlinear problems. Additionally, this ensemble learning approach enables the model to be robust when subjected to noise and outliers, which improves the stability of the model.
Compared to RF, the XGB model employs the idea of gradient boosting. The model is deployed by gradually introducing new trees, each of which attempts to correct the bias (the model’s prediction error relative to the true value) of the previous tree [46]. The final model is a weighted sum of these trees (for regression problems) or a voting result (for classification problems). The flowchart of the XGB model is illustrated in Figure 2. The training objective of XGB is to minimize an objective function consisting of a loss function and a regularization term. The general form of this objective function is given as
O b j e c t i v e = i = 1 n L y i , y ^ i + k = 1 K Ω f k
where n is the number of samples, L y i , y ^ i is the loss function measuring the prediction error of the model for sample i, K is the number of trees, fk is the kth tree, and Ω(fk) is the regularization term of the tree.
To enhance the performance of XGB and ensure the robustness of the model, XGB introduces the concept of learning rate, which is used to control the weight of each tree. This concept enables the model to be more stable by reducing the impact of each tree in every iteration.

3.3. Artificial Neural Network

Artificial neural network (ANN) is a computational model inspired by the biological neural network, which emulates the information processing mechanism of the human brain [47]. ANN consists of three distinct layers, including input, hidden, and output. The structure of ANN is shown in Figure 3. Neurons are interconnected through assigned weights, reflecting the degree of influence of neurons in the previous layer on the ones in the next layer [48].
Weight coefficients and activation functions are essential for the ANN model [49,50]. The function of the weight coefficients is to facilitate the propagation of neural signals, and the propagation process can be expressed by
f i = i = 1 n w i j · x i + b
where fi represents the weighted sum of the outputs of the ith layer in the forward propagation, wij represents the corresponding weighting coefficients of the neurons between the previous jth layer and the ith layer, xi represents the inputs of the neurons in layer i, and b represents the bias. The activation function introduces nonlinear characteristics that empower ANN to learn and represent more intricate relationships. The commonly used functions include Sigmoid, Tanh, ReLU (Rectified Linear Unit), etc. In this work, the ReLU function is employed. The expression for the activation function is shown in Equation (4).
O i = h f i
where Oi represents the output value of layer i, and h represents the activation function.

3.4. Evaluation Metrics

This work evaluates the performance of the model using four parameters, which are R2, MAE, RMSE, and MAPE. The definitions of these four parameters are
R 2 = 1 i = 1 n y i y ^ i 2 i = 1 n y i y ¯ i 2
M A E = 1 n i = 1 n y i y ^ i
R M S E = 1 n i = 1 n y i y ^ i 2
and
M A P E = 1 n i = 1 n y i y ^ i y i × 100 %
where n represents the number of samples, yi represents the actual value, y ^ i represents the predicted value, and y ¯ i represents the average of the actual values.

3.5. K-Fold Cross-Validation

In this study, k-fold cross-validation is utilized to assess the performance of a model on a limited dataset and to improve the generalization of the model. The steps of the k-fold cross-validation method are as follows:
(1)
Divide the original dataset into k folds of mutually exclusive subsets with equal size.
(2)
For each round, one fold is selected as the test set and the remaining folds are used as the training set. The model is trained, and its performance is evaluated by the test set. Repeat this process until each fold has been used as a test set.
(3)
The results of each round of the model evaluation are summarized, and the final output is the average of the results of the k rounds.
It has been proven that 10-fold cross-validation is an effective approach, which balances performance and time consumption [51]. Therefore, this study applies the 10-fold cross-validation method to improve the performance of the model and avoid the overfitting issue. A general schematic diagram of the 10-fold cross-validation method is shown in Figure 4.

4. Results and Discussion

Based on the collected data, the electrical resistivity of the GORCCs was predicted and analyzed by employing RF, XGB, and ANN, respectively. As the prediction performance of the ANN models is intricately related to the configuration of the hidden layer, four ANN models with different hidden layers were applied in this work.

4.1. Correlation Analysis

In this study, correlation analysis was conducted during feature engineering. Figure 5 demonstrates the Pearson correlation coefficient between continuous features and the electrical resistivity of the GORCCs. A strong negative correlation was observed between the electrical resistivity and the w/c, with a value of −0.7. The experimental samples were maintained in an environment with humidity exceeding 90% and kept in water at about 23 ℃ after being unmolded. During the cement curing process, the pores within the cement were filled with capillary water, adsorbed water, etc. The porosity in the cement increased with the increase of the w/c, resulting in greater absorption of water with dissolved ions (i.e., hydroxide and calcium ions) in these pores. Ionic conduction dominated at this stage, leading to an increase in the electrical conductivity of the cement specimens. Another strong negative correlation was observed between curing temperature and the electrical resistivity, with a value of −0.68, while a positive correlation was observed between curing age and the electrical resistivity, with a value of 0.17. These correlations emphasize the importance of the curing environment, which has a significant influence on the electrical resistivity of the GORCCs. The strong positive correlation of Ks, with a value of 0.86, indicates the important effect of specimen size. The positive correlation between specimen size and the electrical resistivity can be attributed to the fact that smaller specimens had a higher probability of forming effective conductive networks because of their reduced volume, enhancing the electrical conductivity. In contrast, larger specimens had a lower probability of forming such conductive networks throughout the entire sample, leading to higher electrical resistivity. In addition, larger specimens might be subjected to non-uniform curing conditions from outside to inside, which might induce different microstructure, resulting in an increase in the resistivity of the sample. Therefore, a positive correlation was shown between specimen size and the resistivity. The Pearson correlation between the dimension of GO and the electrical resistivity of the GORCCs is weak. Effective dispersion is crucial for forming conductive networks within the composite. Therefore, the uniform distribution of GO nanosheets in the specimen had a greater impact on the final results compared to the size of GO. The negative correlation between the GO content and the electrical resistivity shows that the incorporation of GO reduces the electrical resistivity of the GORCCs. When well dispersed within the cement matrix, GO sheets form interconnected conductive pathways, facilitating electron transport across the composite and thereby reducing overall electrical resistivity. Additionally, the incorporation of GO improves the microstructure of the cement composite by filling voids and reducing porosity. A denser microstructure with fewer defects enhances the material’s electrical conductivity. Figure 6 further reveals the correlation between the electrical resistivity and the w/c, Ks, and Kh, respectively. The electrical resistivity tends to decrease with the increase of the w/c. Figure 6b,c suggest that the electrical resistivity increases with the increase of Ks and Kh. In addition to the positive correlation between specimen size and the resistivity, another possible reason is that the presence of cracks increases as the specimen size increases. According to previous studies, cracks may interrupt the conductive paths formed by the conductive nanoparticles, leading to increased electrical resistivity [52].

4.2. Electrical Resistivity by RF and XGB

Both RF and XGB belong to the ensemble learning algorithms. Table 2 presents the performance of the models used in this work. It can be seen that the XGB model demonstrates superior prediction performance, with MAE, RMSE, and MAPE being 0.76 Ω·m, 3.18 Ω·m, and 1.04%, respectively. In contrast, the RF model shows inferior prediction performance, with MAE, RMSE, and MAPE being 3.95 Ω·m, 10.33 Ω·m, and 5.75%, respectively.
Figure 7 illustrates the relationship between the actual and predicted values for the RF and XGB models. The distribution of data points for both models shows close proximity to the line y = x, suggesting excellent performances of the models. However, Figure 7a shows that there is a data point noticeably deviating from the line y = x. The actual and predicted values of this point are 24.97 Ω·m and 36.78 Ω·m, respectively. In contrast, in Figure 7b, the prediction for all data points demonstrates closer proximity to the actual values, with reduced average and maximum error values compared to the predictions obtained from RF. The results indicate that the XGB model performs better than the RF model in predicting the electrical resistivity of the GORCCs.
Figure 8 displays the distribution of data points representing actual and predicted values for both the RF and XGB models. The data points with errors smaller than 10 Ω·m exceed 90% for both models. The maximum, minimum, and average errors for the RF model are 39.81 Ω·m, 0.002 Ω·m, and 3.95 Ω·m, respectively. In contrast, these three corresponding errors for the XGB model are 18.75 Ω·m, 0.002 Ω·m, and 0.76 Ω·m, respectively, which indicates that the maximum and average errors are significantly reduced compared to the ones of the RF model. The value of R2 is 0.99 for both models. As shown in Equation (5), when the predicted value is relatively close to the actual value, and the actual value significantly deviates from the average of all actual values, the value of R2 approaches 1.
Figure 9 shows a feature importance analysis of the predicted results by the RF and the XGB models. The metrics in the figures are used to quantify the degree of contribution of each feature to the final result. A higher value indicates a more significant effect on the electrical resistivity of the GORCCs. Figure 9a,b show feature importance using the RF technique and the F-score of the XGB model, respectively. It is evident in Figure 9a that the measurement method, specimen size (Ks), and cement type have significant influence on the electrical resistivity, with the values being 0.232, 0.217, and 0.213, respectively. This suggests that it is of great importance to consider the effect of the selection of measurement method, specimen size, and cement type on the electrical resistivity of the GORCC samples. In addition, both curing age and the content of GO have a moderate impact on the electrical resistivity. In Figure 9b, the thickness of GO emerges as the most influential factor on the electrical resistivity, as indicated by the high value of the F-score value. The second most influential factor is the measurement method, whose F-score reaches 1045. It is found that the four-probe method has higher accuracy compared to the two-electrode method. This can be explained by the fact that the accuracy of the two-electrode method is affected by the contact resistance between the electrodes and the specimen. In contrast, the four-probe method can compensate for the inaccuracy caused by the contact resistance by using separate pairs of electrodes for current injection and voltage measurement. The content of GO, with an F-score exceeding 1000, also exhibits apparent effects on the electrical resistivity. When well dispersed within the cement matrix, GO sheets form conductive pathways that facilitate electron transport and reduce the electrical resistivity by filling voids and reducing porosity. In conclusion, both the RF and XGB models place emphasis on the measurement methods, curing age, and content of GO, which indicates that more attention should be paid to these aspects when designing experiments.

4.3. Electrical Resistivity by ANN

ANN provides an efficient and accurate computational method for solving complex nonlinear problems [53]. However, the structure of ANN has a substantial impact on the performance and prediction of the model. In order to determine the optimal structure of ANN, this study adopted ANN models with different numbers of hidden layers to predict the electrical resistivity of the GORCCs. The four distinct ANN models are as follows: the first features a single hidden layer with 32 neurons; the second comprises two hidden layers with 64 and 32 neurons, respectively; the third involves three hidden layers with 64, 32, and 16 neurons, respectively; and the fourth incorporates four hidden layers with 64, 32, 16, and 8 neurons, respectively. Among them, the ReLU function was utilized as the activation function, which is expressed as
f ( x ) = max 0 , x
where x is the input value. The ReLU function takes the value of x in the region x > 0 and 0 in the region x ≤ 0.
Table 3 demonstrates the evaluation metrics of the employed ANN models. It is evident that as the number of hidden layers increases, the performance of the model initially improves and then declines. This observation suggests that an increase in the number of hidden layers does not necessarily lead to better performance. The ANN model with 3 hidden layers exhibits the best performance in predicting the electrical resistivity of the GORCCs, with MAE, RMSE, and MAPE being 1.11 Ω·m, 3.07 Ω·m, and 2.36%, respectively.
Figure 10 illustrates the scatter distributions of actual and predicted values for the four ANN models on the test set. It can be observed from Figure 10a that a number of data points deviate from the line y = x, which indicates the poor performance of the ANN 15-32-1 model. The maximum error value reaches 97.44 Ω·m, which is larger than the average error value of 14.88 Ω·m. In comparison, the accuracy of the predictions of the remaining three ANN models has improved significantly. Among them, the ANN 15-64-32-16-1 model demonstrates the best performance, with 80% of data points having an error value below 1 Ω·m. It is evident from Figure 10c that all the data points align closely with the line y = x, which means that the predicted values are relatively close to the actual values. The maximum error is 17.33 Ω·m, with the average error being 1.11 Ω·m only.
Figure 11 demonstrates the errors in predictions from the four ANN models. It is found that the maximum error in Figure 11a is the largest, surpassing 90 Ω·m. Furthermore, there are 8 data points with an error exceeding 10 Ω·m, which indicates the worst performance among the four models. In Figure 11b,d, the number of data points with error values exceeding 10 Ω·m declines, although the value of the maximum error data point is still around 50 Ω·m. In contrast, the errors in Figure 11c are significantly reduced. There is only one data point with an error exceeding 10 Ω·m, while approximately 91.4% of the data points exhibit an error of less than 2 Ω·m. This indicates a significantly improved prediction, highlighting the superior performance of the ANN model with three hidden layers.
In addition, the function of the ANN model with three hidden layers can be expressed as
O = F 1 16 w ¯ q i 3 F 1 32 w ¯ r q 2 F 1 64 w ¯ p r 1 i ¯ p + b ¯ r 1 + b ¯ q 2 + b ¯ i 3
where w ¯ m n l represents the weights connecting the mth neurons in the lth layer and the nth neurons in the (l + 1)th layer, b ¯ n l and i ¯ n are the added bias and the net input in the lth layer for the nth neuron in the (l + 1)th layer, and F represents the activation function, i.e., the ReLU function.

4.4. Comparison of Performance Evaluation for Different Models

Figure 12 demonstrates the modified Taylor diagram of different ML models as involved in this work, where ANN-1 to ANN-4 represent the ANN models with 1, 2, 3, and 4 hidden layers, respectively. The Taylor diagram provides a visual representation of the relationship between actual and predicted values, facilitating the comparison of the evaluation metrics. The radial line in the Taylor diagram represents the correlation coefficient, indicating the relationship between the experimental results and the predictions. A higher correlation coefficient means better performance of the model. The horizontal and vertical axes of the Taylor diagram correspond to the normalized standard deviation, representing the ratio of the standard deviation of the predicted value to the standard deviation of the actual value. From the diagram, it is found that the ANN model with one hidden layer demonstrates the worst prediction performance, featuring lower correlation coefficients and higher normalized RMSE values. As observed in the enlarged figure, the XGB and ANN-3 models exhibit superior performance in predicting the electrical resistivity. The normalized standard deviation for both models is close to 1, indicating an equivalent dispersion between the actual value and the predicted data. Furthermore, the XGB and ANN-3 models demonstrate higher correlation coefficients and lower normalized RMSE. The performance of the ANN-4 model exhibits a clear overall decline. This further suggests that more hidden layers are not always favorable for better performance. Although increasing the number of hidden layers and neurons normally may increase the learning ability of the network, allowing it to better fit complex data patterns, an overly complex network structure may lead to overfitting [54], resulting in poor performance of the ANN model.

5. Conclusions

In order to investigate the complex nonlinear relationship between the coupling of multiple factors and the electrical resistivity of the GORCCs, three different ML algorithms are employed in this study, including RF, XGB, and ANN. A total set of 171 data points is collected and utilized after data cleaning. In this work, the dimensions of GO, such as diameter, thickness, and number of layers of GO, are considered as features following a Gaussian distribution. Particularly, apart from considering the w/c, cement type, etc. as input features, this work is the first attempt to incorporate the effects of specimen size on the electrical resistivity of the GORCCs. Among them, cement type, dispersion method, and measurement method are processed by the one-hot encoding method for feature engineering. Furthermore, in order to describe the correlation between individual features and the electrical resistivity of the GORCCs, the Pearson correlation coefficient matrix is calculated in this study. This work conducts the error analysis of the ANN models with different hidden layers. All the above-mentioned models are evaluated by four metrics, including R2, MAE, RMSE, and MAPE. The results are visualized in the form of a Taylor diagram. The major conclusions of this work can be summarized as follows:
(1)
The three ML models are evidenced to be capable of predicting the resistivity of the GORCCs, which suggests that ML is promising for practical applications in the field of identifying the nonlinear and complex relationship between the material properties and various influencing factors.
(2)
The structure of the ANN has a significant impact on the prediction performance. It is found that the ANN model with hidden layers may increase the learning ability of the network, allowing it to better fit complex data patterns. However, more hidden layers may lead to overfitting and are not always favorable for improving the prediction performance. The ANN model with the structure of 15-64-32-16-1 is proven to have the best performance in this work. This can provide useful guidelines when conducting ML modeling for predicting the material properties.
(3)
The strong correlation between specimen size and the resistivity suggests that the consideration of the specimen size is important for accurate capture of the material properties. In addition, it is found that the electrical resistivity of the GORCCs is also highly dependent on the measurement method. This observation is of great importance when selecting the measurement method to obtain accurate and objective experimental data.

Author Contributions

Conceptualization, C.F.; data curation, R.C.; formal analysis, R.C.; funding acquisition, C.F.; investigation, Z.H. and Y.F.; methodology, R.C.; supervision, C.F.; validation, J.Y.; visualization, J.Z.; writing—original draft, R.C.; writing—review and editing, C.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Innovative and Entrepreneurial Talents Program of Jiangsu Province, China.

Data Availability Statement

Data will be made available upon request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have influenced the work reported in this paper.

Appendix A

Table A1. Details of the collected data.
Table A1. Details of the collected data.
No.Type of
Cement
Dispersion
Method
Measurement
Method
w/cCuring
Temperature
Curing
Age
KsKhDiameter
(μm)
Thickness
(nm)
LayersGO
Content
Electrical ResistivityReference
11110.40 20280.660.30.22.4920.000 183.95 [15]
21110.40 20280.660.30.81.4330.005 206.12
31110.40 20280.660.32.51.5620.010 230.08
41110.40 20280.660.30.72.3910.015 262.88
51110.40 20280.660.31.81.0810.020 302.39
61110.40 20280.660.30.32.6910.025 296.78
71110.40 20280.660.30.42.7420.030 293.97
81110.40 20280.660.30.21.3810.035 294.57
91110.40 20280.660.32.41.7510.040 291.26
101110.45 231 111.32.2610.000 7.58 [55]
111110.45 232 111.91.1110.000 10.41
121110.45 233 112.11.7920.000 13.08
131110.45 235 1111.4930.000 15.33
141110.45 237 1111.610.000 17.24
151110.45 2310 110.62.5120.000 17.88
161110.45 2314 110.92.5810.000 18.68
171110.45 2318 110.92.3310.000 19.24
181110.45 2320 111.52.3220.000 19.57
191110.45 2325 112.92.7320.000 20.22
201110.45 2328 111.42.6520.000 20.59
211110.45 231 110.51.2110.010 9.73
221110.45 232 110.81.9310.010 13.06
231110.45 233 111.31.8710.010 15.73
241110.45 235 110.62.8310.010 16.90
251110.45 237 110.52.5510.010 18.10
261110.45 2310 111.51.1710.010 18.83
271110.45 2314 110.91.6110.010 19.80
281110.45 2318 112.52.4410.010 20.79
291110.45 2320 112.11.8130.010 21.32
301110.45 2325 111.92.8420.010 22.59
311110.45 2328 110.41.8410.010 23.29
321110.45 231 112.61.5810.020 7.64
331110.45 232 111.31.7110.020 10.54
341110.45 233 110.71.9730.020 12.91
351110.45 235 111.31.5510.020 14.82
361110.45 237 112.42.1710.020 16.94
371110.45 2310 110.31.5210.020 17.93
381110.45 2314 1122.810.020 19.14
391110.45 2318 110.92.8820.020 19.76
401110.45 2320 110.11.2810.020 20.27
411110.45 2325 110.62.2110.020 21.09
421110.45 2328 110.21.2330.020 21.84
431110.45 231 111.82.8410.040 7.96
1441110.45 232 110.81.4710.040 9.32
451110.45 233 110.62.4730.040 10.83
461110.45 235 112.91.4630.040 13.67
471110.45 237 112.52.9510.040 16.30
481110.45 2310 110.62.9130.040 17.48
491110.45 2314 110.82.0610.040 18.96
501110.45 2318 111.92.6810.040 19.36
511110.45 2320 112.71.8110.040 19.60
521110.45 2325 110.31.7420.040 20.73
531110.45 2328 111.82.3830.040 21.64
541110.45 231 111.91.4420.080 7.34
551110.45 232 111.81.2120.080 9.39
561110.45 233 110.61.5730.080 11.23
571110.45 235 112.21.3620.080 13.38
581110.45 237 110.11.9730.080 15.28
591110.45 2310 110.41.3620.080 16.90
601110.45 2314 112.12.8620.080 18.75
611110.45 2318 1111.8820.080 19.15
621110.45 2320 110.32.2710.080 19.41
631110.45 2325 111.71.3530.080 20.28
641110.45 2328 112.31.5810.080 20.79
651110.45 231 111.12.7820.160 7.46
661110.45 232 110.61.0310.160 9.42
671110.45 233 110.82.4610.160 10.95
681110.45 235 1112.0120.160 13.69
691110.45 237 110.42.3920.160 16.09
701110.45 2310 110.42.630.160 16.98
711110.45 2314 112.82.4520.160 18.13
721110.45 2318 111.91.7310.160 18.59
731110.45 2320 111.72.5810.160 18.84
741110.45 2325 110.72.3210.160 19.54
751110.45 2328 112.11.3130.160 19.98
761110.40 20280.660.30.61.9120.007 90.52 [18]
771110.40 20280.660.30.31.4120.008 85.67
781110.40 20280.660.31.42.0930.009 78.49
791110.40 20280.660.31.82.0720.010 69.65
801110.40 20280.660.32.41.7330.011 61.86
811110.40 20280.660.30.12.7330.012 53.02
821110.40 20280.660.30.42.9120.013 43.26
831110.40 20280.660.31.21.8220.014 41.60
841110.40 20280.660.32.71.6410.015 39.64
851110.40 20280.660.331.2410.016 39.02
861110.40 20280.660.30.21.3130.017 36.08
871110.40 20280.660.32.82.3730.018 34.42
881110.40 20280.660.32.62.8220.019 32.46
891110.40 20280.660.31.82.2310.020 30.19
901110.40 20280.660.321.0210.021 28.59
911110.40 20280.660.31.61.4820.022 25.95
921110.40 20280.660.30.82.4430.023 24.97
931110.40 20280.660.30.91.1410.024 24.36
941110.40 20280.660.31.32.9820.025 21.72
951110.40 20280.660.30.9130.026 19.75
961120.43 2071.991.412.21.5820.020 344.46 [56]
972120.43 2071.991.410.41.7510.040 284.56
982120.43 2071.991.410.12.5410.080 263.79
992120.43 2071.991.412.91.0430.160 216.08
1002120.43 20141.991.410.92.7420.020 381.30
1012120.43 20141.991.412.81.3230.040 314.87
1022120.43 20141.991.412.31.4220.080 287.94
1032120.43 20141.991.410.31.7510.160 244.95
1042120.43 20281.991.410.92.8320.020 410.28
1052120.43 20281.991.4131.710.040 341.68
1062120.43 20281.991.412.12.9520.080 314.15
1072120.43 20281.991.411.41.2610.160 299.41
1081210.45 231 110.61.4130.020 6.19 [17]
1091210.45 233 111.41.0910.020 10.63
1101210.45 235 111.22.5910.020 11.73
1111210.45 237 110.51.1530.020 12.69
1121210.45 2310 112.22.3920.020 13.66
1131210.45 2312 111.91.7210.020 14.22
1141210.45 2314 112.71.9630.020 14.81
1151210.45 2317 111.81.3710.020 15.60
1161210.45 2320 110.51.8430.020 16.48
1171210.45 2323 112.12.3430.020 17.21
1181210.45 2325 112.22.5420.020 17.78
1191210.45 2328 111.12.8330.020 18.56
1201210.45 231 110.11.420.040 6.77
1211210.45 233 112.11.5320.040 10.27
1221210.45 235 112.61.1320.040 12.49
1231210.45 237 111.41.5810.040 14.39
1241210.45 2310 111.92.3230.040 14.70
1251210.45 2312 110.42.6220.040 14.93
1261210.45 2314 110.62.4210.040 15.10
1271210.45 2317 111.62.8130.040 15.78
1281210.45 2320 111.91.3330.040 16.44
1291210.45 2323 110.31.3920.040 17.08
1301210.45 2325 110.91.6630.040 17.61
1311210.45 2328 111.41.830.040 18.16
1321210.45 231 112.51.110.060 6.62
1331210.45 233 110.11.5810.060 10.43
1341210.45 235 112.92.7630.060 11.58
1351210.45 237 110.72.6910.060 12.80
1361210.45 2310 110.12.2120.060 13.57
1371210.45 2312 112.62.7730.060 14.07
1381210.45 2314 110.4320.060 14.59
1391210.45 2317 112.11.4910.060 15.38
1401210.45 2320 111.81.7810.060 16.15
1411210.45 2323 110.11.6520.060 16.86
1421210.45 2325 110.81.8110.060 17.40
1431210.45 2328 110.81.6620.060 18.15
1442120.40 221 2.560.80.11.0310.040 210.87 [57]
1452120.40 192 2.560.82.11.0310.040 282.33
1462120.40 183 2.560.81.41.7620.040 386.96
1472120.40 214 2.560.81.11.8230.040 386.96
1482120.40 195 2.560.82.11.5430.040 378.68
1492120.40 196 2.560.81.92.520.040 382.80
1502120.40 207 2.560.82.92.1120.040 382.80
1512120.40 198 2.560.82.12.0720.040 382.80
1522120.40 219 2.560.81.72.4110.040 386.96
1532120.40 2210 2.560.82.61.8530.040 382.80
1542120.40 2111 2.560.82.71.8510.040 391.16
1552120.40 1812 2.560.80.61.3520.040 395.41
1562120.40 1813 2.560.80.81.4520.040 398.99
1572120.40 2214 2.560.82.32.130.040 403.32
1582120.40 1815 2.560.80.32.8810.040 412.14
1592120.40 2216 2.560.832.4920.040 429.57
1602120.40 1917 2.560.82.41.5610.040 443.73
1612120.40 1818 2.560.80.12.6410.040 462.50
1622120.40 2119 2.560.82.32.5630.040 482.06
1632120.40 2020 2.560.80.22.3810.040 497.95
1642120.40 2121 2.560.82.22.9620.040 530.35
1652120.40 2022 2.560.80.41.910.040 535.15
1662120.40 1823 2.560.81.51.610.040 576.17
1672120.40 1824 2.560.82.12.4630.040 576.17
1682120.40 2225 2.560.82.51.0310.040 595.16
1692120.40 2126 2.560.80.52.0830.040 620.34
1702120.40 1927 2.560.81.81.4830.040 660.71
1712120.40 2028 2.560.80.41.9430.040 653.60

References

  1. Yang, M.; Chen, L.; Lai, J.; Osman, A.I.; Farghali, M.; Rooney, D.W.; Yap, P.-S. Advancing environmental sustainability in construction through innovative low-carbon, high-performance cement-based composites: A review. Mater. Today Sustain. 2024, 26, 100712. [Google Scholar] [CrossRef]
  2. Huang, J.; Liu, X.; Long, Y.; Li, W.; Wu, R. An Investigation into the Variables Influencing the Structural Bamboo Architecture Using Filled Concrete and Cement Mortar. Buildings 2024, 14, 2029. [Google Scholar] [CrossRef]
  3. Singh, N.B.; Kalra, M.; Saxena, S.K. Nanoscience of Cement and Concrete. Mater. Today Proc. 2017, 4, 5478–5487. [Google Scholar] [CrossRef]
  4. Wang, X.; Dong, S.; Li, Z.; Han, B.; Ou, J. Nanomechanical Characteristics of Interfacial Transition Zone in Nano-Engineered Concrete. Engineering 2022, 17, 99–109. [Google Scholar] [CrossRef]
  5. Zhang, P.; Wang, M.; Han, X.; Zheng, Y. A review on properties of cement-based composites doped with graphene. J. Build. Eng. 2023, 70, 106367. [Google Scholar] [CrossRef]
  6. Norhasri, M.S.M.; Hamidah, M.S.; Fadzil, A.M. Applications of using nano material in concrete: A review. Constr. Build. Mater. 2017, 133, 91–97. [Google Scholar] [CrossRef]
  7. Joseph, L.; Kalyana Chakravarthi, E.; Sarath Kumar, P.; Jayanarayanan, K.; Mini, K.M. Nano filler incorporated epoxy based natural hybrid fiber confinement of concrete systems: Effect of fiber layers and nano filler addition. Structures 2023, 51, 320–331. [Google Scholar] [CrossRef]
  8. Du, H.; Pang, S.D. High performance cement composites with colloidal nano-silica. Constr. Build. Mater. 2019, 224, 317–325. [Google Scholar] [CrossRef]
  9. Ardanuy, M.; Claramunt, J.; Toledo Filho, R.D. Cellulosic fiber reinforced cement-based composites: A review of recent research. Constr. Build. Mater. 2015, 79, 115–128. [Google Scholar] [CrossRef]
  10. Yan, S.; Cheng, Y.; Wang, W.; Jin, L.; Ding, Z. Gypsum-Enhanced Red Mud Composites: A Study on Strength, Durability, and Leaching Characteristics. Buildings 2024, 14, 1979. [Google Scholar] [CrossRef]
  11. Ramezani, A.M.; Khajehdezfuly, A.; Poorveis, D. Structural Lightweight Concrete Containing Basalt Stone Powder. Buildings 2024, 14, 1904. [Google Scholar] [CrossRef]
  12. Lee, S.; Jin, J.-U.; Hahn, J.R.; Ryu, S.; You, N.-H. Highly water-dispersible methylpyridinium salt functionalized reduced graphene oxide and poly(vinyl alcohol) composites. Compos. Part B Eng. 2024, 271, 111142. [Google Scholar] [CrossRef]
  13. Naseem, Z.; Shamsaei, E.; Sagoe-Crentsil, K.; Duan, W. Rheological enhancement of fresh polymer-modified cement composites via surface-modified graphene oxide. Cem. Concr. Compos. 2024, 147, 105413. [Google Scholar] [CrossRef]
  14. Naseem, Z.; Shamsaei, E.; Sagoe-Crentsil, K.; Duan, W. Antifoaming effect of graphene oxide nanosheets in polymer-modified cement composites for enhanced microstructure and mechanical performance. Cem. Concr. Res. 2022, 158, 106843. [Google Scholar] [CrossRef]
  15. Suo, Y.; Guo, R.; Xia, H.; Yang, Y.; Zhou, B.; Zhao, Z. A review of graphene oxide/cement composites: Performance, functionality, mechanisms, and prospects. J. Build. Eng. 2022, 53, 104502. [Google Scholar] [CrossRef]
  16. Pan, Z.; He, L.; Qiu, L.; Korayem, A.H.; Li, G.; Zhu, J.W.; Collins, F.; Li, D.; Duan, W.H.; Wang, M.C. Mechanical properties and microstructure of a graphene oxide–cement composite. Cem. Concr. Compos. 2015, 58, 140–147. [Google Scholar] [CrossRef]
  17. Qureshi, T.S.; Panesar, D.K. Impact of graphene oxide and highly reduced graphene oxide on cement based composites. Constr. Build. Mater. 2019, 206, 71–83. [Google Scholar] [CrossRef]
  18. Li, X.; Wang, L.; Liu, Y.; Li, W.; Dong, B.; Duan, W.H. Dispersion of graphene oxide agglomerates in cement paste and its effects on electrical resistivity and flexural strength. Cem. Concr. Compos. 2018, 92, 145–154. [Google Scholar] [CrossRef]
  19. Sartipi, F.; Ghari Zadeh, A.; Gamil, M. Electrical resistance of graphene reinforced cement paste. J. Constr. Mater. 2019. [Google Scholar] [CrossRef]
  20. Guo, R.; Suo, Y.; Xia, H.; Yang, Y.; Ma, Q.; Yan, F. Study of Piezoresistive Behavior of Smart Cement Filled with Graphene Oxide. Nanomaterials 2021, 11, 206. [Google Scholar] [CrossRef]
  21. Hu, Y.-G.; Awol, J.F.; Chen, S.; Jiang, J.N.; Pu, X.; Jia, X.; Xu, X.Q. Experimental study of the electrical resistance of graphene oxide-reinforced cement-based composites with notch or rebar. J. Build. Eng. 2022, 51, 104331. [Google Scholar] [CrossRef]
  22. Yan, Y.; Tian, L.; Zhao, W.; Lazaro, S.A.M.; Li, X.; Tang, S. Dielectric and mechanical properties of cement pastes incorporated with magnetically aligned reduced graphene oxide. Dev. Built Environ. 2024, 18, 100471. [Google Scholar] [CrossRef]
  23. Xiong, G.; Ren, Y.; Wang, C.; Zhang, Z.; Zhou, S.; Kuang, C.; Zhao, Y.; Guo, B.; Hong, S. Effect of power ultrasound assisted mixing on graphene oxide in cement paste: Dispersion, microstructure and mechanical properties. J. Build. Eng. 2023, 69, 106321. [Google Scholar] [CrossRef]
  24. Li, W.; Li, X.; Chen, S.J.; Liu, Y.M.; Duan, W.H.; Shah, S.P. Effects of graphene oxide on early-age hydration and electrical resistivity of Portland cement paste. Constr. Build. Mater. 2017, 136, 506–514. [Google Scholar] [CrossRef]
  25. Zhang, P.; Kong, F.; Hai, L. Strength Prediction of Smart Cementitious Materials Using a Neural Network Optimized by Particle Swarm Algorithm. Buildings 2024, 14, 2033. [Google Scholar] [CrossRef]
  26. Li, H.; Pan, Z.; Yang, Y.; Wang, X.; Tang, H.; Ma, F.; Zheng, L. Predicting Residual Flexural Strength of Corroded Prestressed Concrete Beams: Comparison of Chinese Code, Eurocode and ACI Standard. Buildings 2024, 14, 2047. [Google Scholar] [CrossRef]
  27. Liu, Q.; Hu, D.; Jin, Q.; Zhu, L.; Xu, K.; Zhou, Z.; Su, W. Projection-Pursuit Regression-Based Optimization of Frost Resistance and Mechanical Performance in Alkali-Activated Slag Cement Pavements. Buildings 2024, 14, 2034. [Google Scholar] [CrossRef]
  28. Huang, J.S.; Liew, J.X.; Liew, K.M. Data-driven machine learning approach for exploring and assessing mechanical properties of carbon nanotube-reinforced cement composites. Compos. Struct. 2021, 267, 113917. [Google Scholar] [CrossRef]
  29. Montazerian, A.; Arve Øverli, J.; Goutianos, S. Thermal conductivity of cementitious composites reinforced with graphene-based materials: An integrated approach combining machine learning with computational micromechanics. Constr. Build. Mater. 2023, 395, 132293. [Google Scholar] [CrossRef]
  30. Bang, J.; Park, S.; Jeon, H. Piezoresistive Prediction of CNTs-Embedded Cement Composites via Machine Learning Approaches. Comput. Mater. Contin. 2022, 71, 1503–1519. [Google Scholar] [CrossRef]
  31. Kekez, S.; Kubica, J. Application of Artificial Neural Networks for Prediction of Mechanical Properties of CNT/CNF Reinforced Concrete. Materials 2021, 14, 5637. [Google Scholar] [CrossRef] [PubMed]
  32. Li, Y.; Li, H.; Jin, C.; Shen, J. The study of effect of carbon nanotubes on the compressive strength of cement-based materials based on machine learning. Constr. Build. Mater. 2022, 358, 129435. [Google Scholar] [CrossRef]
  33. Alyousef, R.; Nassar, R.-U.-D.; Fawad, M.; Farooq, F.; Gamil, Y.; Najeh, T. Predicting the properties of concrete incorporating graphene nano platelets by experimental and machine learning approaches. Case Stud. Constr. Mater. 2024, 20, e03018. [Google Scholar] [CrossRef]
  34. Cosoli, G.; Mobili, A.; Tittarelli, F.; Revel, G.M.; Chiariotti, P. Electrical resistivity and electrical impedance measurement in mortar and concrete elements: A systematic review. Appl. Sci. 2020, 10, 9152. [Google Scholar] [CrossRef]
  35. Fan, X.M. Effects of environmental temperature and humidity on the electrical properties of carbon fiber graphite cement mortar. Adv. Mat. Res. 2011, 143, 1022–1026. [Google Scholar] [CrossRef]
  36. Dong, W.; Li, W.; Guo, Y.; Qu, F.; Wang, K.; Sheng, D. Piezoresistive performance of hydrophobic cement-based sensors under moisture and chloride-rich environments. Cem. Concr. Compos. 2022, 126, 104379. [Google Scholar] [CrossRef]
  37. Eskandari-Naddaf, H.; Kazemi, R. ANN prediction of cement mortar compressive strength, influence of cement strength class. Constr. Build. Mater. 2017, 138, 1–11. [Google Scholar] [CrossRef]
  38. Sun, J.; Wang, Y.; Yao, X.; Ren, Z.; Zhang, G.; Zhang, C.; Chen, X.; Ma, W.; Wang, X. Machine-learning-aided prediction of flexural strength and ASR expansion for waste glass cementitious composite. Appl. Sci. 2021, 11, 6686. [Google Scholar] [CrossRef]
  39. Zhang, J.; Niu, W.; Yang, Y.; Hou, D.; Dong, B. Machine learning prediction models for compressive strength of calcined sludge-cement composites. Constr. Build. Mater. 2022, 346, 128442. [Google Scholar] [CrossRef]
  40. Marani, A.; Nehdi, M.L. Machine learning prediction of compressive strength for phase change materials integrated cementitious composites. Constr. Build. Mater. 2020, 265, 120286. [Google Scholar] [CrossRef]
  41. Sun, J.; Ma, Y.; Li, J.; Zhang, J.; Ren, Z.; Wang, X. Machine learning-aided design and prediction of cementitious composites containing graphite and slag powder. J. Build. Eng. 2021, 43, 102544. [Google Scholar] [CrossRef]
  42. Nguyen, N.-H.; Abellán-García, J.; Lee, S.; Garcia-Castano, E.; Vo, T.P. Efficient estimating compressive strength of ultra-high performance concrete using XGBoost model. J. Build. Eng. 2022, 52, 104302. [Google Scholar] [CrossRef]
  43. Dong, W.; Huang, Y.; Lehane, B.; Ma, G. Multi-objective design optimization for graphite-based nanomaterials reinforced cementitious composites: A data-driven method with machine learning and NSGA-II. Constr. Build. Mater. 2022, 331, 127198. [Google Scholar] [CrossRef]
  44. Sug, H. Applying randomness effectively based on random forests for classification task of datasets of insufficient information. J. Appl. Math. 2012, 2012, 13. [Google Scholar] [CrossRef]
  45. Natras, R.; Soja, B.; Schmidt, M. Ensemble machine learning of Random Forest, AdaBoost and XGBoost for vertical total electron content forecasting. Remote Sens. 2022, 14, 3547. [Google Scholar] [CrossRef]
  46. Alabdullah, A.A.; Iqbal, M.; Zahid, M.; Khan, K.; Amin, M.N.; Jalal, F.E. Prediction of rapid chloride penetration resistance of metakaolin based high strength concrete using light GBM and XGBoost models by incorporating SHAP analysis. Constr. Build. Mater. 2022, 345, 128296. [Google Scholar] [CrossRef]
  47. Ahmad, A.S.; Hassan, M.Y.; Abdullah, M.P.; Rahman, H.A.; Hussin, F.; Abdullah, H.; Saidur, R. A review on applications of ANN and SVM for building electrical energy consumption forecasting. Renew. Sustain. Energy Rev. 2014, 33, 102–109. [Google Scholar] [CrossRef]
  48. da Costa, N.L.; de Lima, M.D.; Barbosa, R. Evaluation of feature selection methods based on artificial neural network weights. Expert Syst. Appl. 2021, 168, 114312. [Google Scholar] [CrossRef]
  49. Rasamoelina, A.D.; Adjailia, F.; Sinčák, P. A review of activation function for artificial neural network. In Proceedings of the 2020 IEEE 18th World Symposium on Applied Machine Intelligence and Informatics (SAMI), Herlany, Slovakia, 23–25 January 2020; pp. 281–286. [Google Scholar]
  50. Liu, J.; Liu, Y.; Zhang, Q. A weight initialization method based on neural network with asymmetric activation function. Neurocomputing 2022, 483, 171–182. [Google Scholar] [CrossRef]
  51. Kohavi, R. A study of cross-validation and bootstrap for accuracy estimation and model selection. In Proceedings of the International Joint Conference on Articial Intelligence, Montreal, QC, Canada, 20–25 August 1995; pp. 1137–1145. [Google Scholar]
  52. Tafesse, M.; Shiferaw Alemu, A.; Yang, B.; Park, S.; Kim, H.-K. Crack monitoring strategy for concrete structures in various service conditions via multiple CNT-CF/cement composite sensors: Experiment and simulation approaches. Cem. Concr. Compos. 2023, 143, 105249. [Google Scholar] [CrossRef]
  53. Shi, L.; Lin, S.T.K.; Lu, Y.; Ye, L.; Zhang, Y.X. Artificial neural network based mechanical and electrical property prediction of engineered cementitious composites. Constr. Build. Mater. 2018, 174, 667–674. [Google Scholar] [CrossRef]
  54. Chu, J.; Liu, X.; Zhang, Z.; Zhang, Y.; He, M. A novel method overcomeing overfitting of artificial neural network for accurate prediction: Application on thermophysical property of natural gas. Case Stud. Therm. Eng. 2021, 28, 101406. [Google Scholar] [CrossRef]
  55. Qureshi, T.S.; Panesar, D.K. Nano reinforced cement paste composite with functionalized graphene and pristine graphene nanoplatelets. Compos. Part B Eng. 2020, 197, 108063. [Google Scholar] [CrossRef]
  56. Kaur, R.; Kothiyal, N.C. Comparative effects of sterically stabilized functionalized carbon nanotubes and graphene oxide as reinforcing agent on physico-mechanical properties and electrical resistivity of cement nanocomposites. Constr. Build. Mater. 2019, 202, 121–138. [Google Scholar] [CrossRef]
  57. Lu, D.; Shi, X.; Wong, H.S.; Jiang, Z.; Zhong, J. Graphene coated sand for smart cement composites. Constr. Build. Mater. 2022, 346, 128313. [Google Scholar] [CrossRef]
Figure 1. Flowchart of RF.
Figure 1. Flowchart of RF.
Buildings 14 02455 g001
Figure 2. Flowchart of XGB.
Figure 2. Flowchart of XGB.
Buildings 14 02455 g002
Figure 3. Structural diagram of ANN.
Figure 3. Structural diagram of ANN.
Buildings 14 02455 g003
Figure 4. Schematic diagram of the 10-fold cross-validation.
Figure 4. Schematic diagram of the 10-fold cross-validation.
Buildings 14 02455 g004
Figure 5. Pearson correlation coefficients between input features and the electrical resistivity.
Figure 5. Pearson correlation coefficients between input features and the electrical resistivity.
Buildings 14 02455 g005
Figure 6. Correlation between features and the electrical resistivity. (a) w/c; (b) Ks; (c) Kh.
Figure 6. Correlation between features and the electrical resistivity. (a) w/c; (b) Ks; (c) Kh.
Buildings 14 02455 g006
Figure 7. Relationship between the true and predicted electrical resistivity of the GORCCs. (a) RF; (b) XGB.
Figure 7. Relationship between the true and predicted electrical resistivity of the GORCCs. (a) RF; (b) XGB.
Buildings 14 02455 g007
Figure 8. Results of the test set on RF and XGB.
Figure 8. Results of the test set on RF and XGB.
Buildings 14 02455 g008
Figure 9. Feature importance. (a) RF; (b) XGB.
Figure 9. Feature importance. (a) RF; (b) XGB.
Buildings 14 02455 g009
Figure 10. Predictive and true values of the electrical resistivity of the GORCCs for the ANN models. (a) 15-32-1; (b) 15-64-32-1; (c) 15-64-32-16-1; (d) 15-64-32-16-8-1.
Figure 10. Predictive and true values of the electrical resistivity of the GORCCs for the ANN models. (a) 15-32-1; (b) 15-64-32-1; (c) 15-64-32-16-1; (d) 15-64-32-16-8-1.
Buildings 14 02455 g010
Figure 11. Distribution of errors of the four ANN models. (a) 15-32-1; (b) 15-64-32-1; (c) 15-64-32-16-1; (d) 15-64-32-16-8-1.
Figure 11. Distribution of errors of the four ANN models. (a) 15-32-1; (b) 15-64-32-1; (c) 15-64-32-16-1; (d) 15-64-32-16-8-1.
Buildings 14 02455 g011
Figure 12. Taylor diagrams of the ML models for electrical resistivity.
Figure 12. Taylor diagrams of the ML models for electrical resistivity.
Buildings 14 02455 g012
Table 1. Description of the collected data.
Table 1. Description of the collected data.
FeaturesCountsMeanMinimumMaximumMedianStandard Deviation
w/c1370.430.40.50.450.02
Curing temperature (°C)13721.721823231.64
Curing age (days)13716.18128159.77
Ks1371.290.662.5610.65
Kh1370.880.31.4110.29
Diameter (μm)1371.560.131.60.90
Thickness (nm)1371.95132.070.57
Layers1371.951320.84
Content of GO (wt%)1370.0500.160.040.04
Cement type 1102-01--
Cement type 235-01--
Dispersion method 1113-01--
Dispersion method 224-01--
Measurement method 1102-01--
Measurement method 2
Electrical resistivity (Ω·m)
35
171
-
123.78
0
6.19
1
660.71
-
19.6
-
176.26
Table 2. Results of RF and XGB for the electrical resistivity of the GORCCs.
Table 2. Results of RF and XGB for the electrical resistivity of the GORCCs.
IndexModelsMAERMSEMAPE
1RF3.9510.335.75%
2XGB0.763.181.04%
Table 3. Predictions of the ANN models with four different hidden layers.
Table 3. Predictions of the ANN models with four different hidden layers.
IndexModelsMAERMSEMAPE
1ANN:15-32-114.8827.7343.9%
2ANN:15-64-32-13.8410.255.21%
3ANN:15-64-32-16-11.113.072.36%
4ANN:15-64-32-16-8-14.1512.664.76%
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

Chen, R.; Feng, C.; Yang, J.; Hang, Z.; Fan, Y.; Zhang, J. Data-Driven Prediction of Electrical Resistivity of Graphene Oxide/Cement Composites Considering the Effects of Specimen Size and Measurement Method. Buildings 2024, 14, 2455. https://doi.org/10.3390/buildings14082455

AMA Style

Chen R, Feng C, Yang J, Hang Z, Fan Y, Zhang J. Data-Driven Prediction of Electrical Resistivity of Graphene Oxide/Cement Composites Considering the Effects of Specimen Size and Measurement Method. Buildings. 2024; 14(8):2455. https://doi.org/10.3390/buildings14082455

Chicago/Turabian Style

Chen, Runyang, Chuang Feng, Jinlong Yang, Ziyan Hang, Yucheng Fan, and Jinzhu Zhang. 2024. "Data-Driven Prediction of Electrical Resistivity of Graphene Oxide/Cement Composites Considering the Effects of Specimen Size and Measurement Method" Buildings 14, no. 8: 2455. https://doi.org/10.3390/buildings14082455

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