Next Article in Journal
Seismic Analysis of Baffle-Reinforced Elevated Storage Tank Using Finite Element Method
Next Article in Special Issue
Development of an Ultra-High-Performance Fibre-Reinforced Concrete (UHPFRC) Manufacturable at Ambient Temperature
Previous Article in Journal
Governance Strategies for Informal Settlements in China: The Case of Guangzhou
Previous Article in Special Issue
Compressive Strength Prediction of High-Strength Concrete Using Long Short-Term Memory and Machine Learning Algorithms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Shear Strength Prediction of Slender Steel Fiber Reinforced Concrete Beams Using a Gradient Boosting Regression Tree Method

by
Amjed Shatnawi
1,
Hana Mahmood Alkassar
2,
Nadia Moneem Al-Abdaly
2,
Emadaldeen A. Al-Hamdany
2,
Luís Filipe Almeida Bernardo
3,* and
Hamza Imran
4
1
Department of Applied Earth Sciences and Environment, Institute of Earth and Environmental Sciences, Al Al-Bayt University, Mafraq 25113, Jordan
2
Department of Civil Engineering, Najaf Technical Institute, Al-Furat Al-Awsat Technical University, Najaf Munazira Str., Najaf 54003, Iraq
3
Centre of Materials and Building Technologies (C-MADE), Department of Civil Engineering and Architecture, University of Beira Interior, 6201-001 Covilhã, Portugal
4
Department of Construction and Project, Al-Karkh University of Science, Baghdad 10081, Iraq
*
Author to whom correspondence should be addressed.
Buildings 2022, 12(5), 550; https://doi.org/10.3390/buildings12050550
Submission received: 2 April 2022 / Revised: 19 April 2022 / Accepted: 21 April 2022 / Published: 25 April 2022

Abstract

:
For the design or assessment of concrete structures that incorporate steel fiber in their elements, the accurate prediction of the shear strength of steel fiber reinforced concrete (SFRC) beams is critical. Unfortunately, traditional empirical methods are based on a small and limited dataset, and their abilities to accurately estimate the shear strength of SFRC beams are arguable. This drawback can be reduced by developing an accurate machine learning based model. The problem with using a high accuracy machine learning (ML) model is its interpretation since it works as a black-box model that is highly sophisticated for humans to comprehend directly. For this reason, Shapley additive explanations (SHAP), one of the methods used to open a black-box machine learning model, is combined with highly accurate machine learning techniques to build an explainable ML model to predict the shear strength of SFRC slender beams. For this, a database of 330 beams with varying design attributes and geometries was developed. The new gradient boosting regression tree (GBRT) machine learning model was compared statistically to experimental data and current shear design models to evaluate its performance. The proposed GBRT model gives predictions that are very similar to the experimentally observed shear strength and has a better and unbiased predictive performance in comparison to other existing developed models. The SHAP approach shows that the beam width and effective depth are the most important factors, followed by the concrete strength and the longitudinal reinforcement ratio. In addition, the outputs are also affected by the steel fiber factor and the shear-span to effective depth ratio. The fiber tensile strength and the aggregate size have the lowest effect, with only about 1% on average to change the predicted value of the shear strength. By building an accurate ML model and by opening its black-box, future researchers can focus on some attributes rather than others.

1. Introduction

Shear failure of reinforced concrete beams is a significant concern due to its brittle and sudden nature [1]. Traditional steel stirrups used as shear reinforcement have been shown to enhance shear capacity efficiently as well as prevent concrete failure. However, incorporating stirrups in narrow, asymmetrical, or congested areas might be challenging. Placing concrete can become a concern when the spacing between stirrups is small, leading to voids in the concrete [2]. Furthermore, traditional stirrups need much labor effort, resulting in more significant building expenses.
In recent years, steel fibers (SF) have acquired significant impetus when utilized in suitable volume fractions due to their potential to replace minimum shear reinforcement [3,4]. Therefore, it has been suggested that implementing SF in the construction industry could provide several advantages. First, SF can enhance the shear resistance by reducing cracks’ width due to the transmission of tensile loads across diagonal cracks, as indicated by Dinh [3]. Concrete shrinkage behavior and post-cracking toughness can also be improved using SF [5,6]. Additionally, the SF inclusion enhances the resistance of the dowel action, which is due to an increase in the tensile strength of the concrete along with the reinforcements in the splitting plane [7].
Various experimental and computational investigations on steel fiber reinforced concrete (SFRC) have been conducted in the literature to examine the shear strength capability [8]. Furthermore, an analysis of the ultimate behavior of SFRC beams was carried out using finite elements and experimental modeling approach utilizing ANSYS software [9]. Additionally, the shear strength of concrete beams with fibers was predicted using a basic physical design model developed by Spinella et al. [10], which considered crack width and shear crack slips. Moreover, utilizing the slenderness ratio, an equation based on fundamental mechanic principles has been proposed to predict the shear strength of SFRC beams [11]. Additionally, SFRC has been studied extensively by conducting several shear experiments on prismatic beams [12].
Several empirical formulas for estimating the shear strength of SFRC beams have been developed in prior research during the last four decades. References [13,14] provide an overview of the most current shear design models and design guidelines for SFRC beams without stirrups. However, the range of validity of such empirical models is limited, which constitute a significant disadvantage. These models are built on the basis of a small number of data specimens, and their accuracy, when applied to additional data cases that fall beyond their range of validity, is arguable. According to a comparison made by references [13,15], it is still challenging to accurately predict the shear capacity of SFRC beams using the various shear resistance models for SFRC. In addition, shear strength predictions from different methodologies differ from one another and still differ from experimentally determined shear strengths. Given these uncertainties, a proper assessment of the reliability of SFRC beams in the event of shear failure is essential. For this, the predictive model must show the highest possible accuracy and the lowest possible variability.
An increase in the use of artificial intelligence (AI) has taken place in recent years due to innovations in computing. Machine learning (ML) models are built using extensive databases. As a result, they can significantly increase generalization capacity and accuracy for measuring the strength of concrete built with various mixing proportions. Therefore, researchers have been inspired by ML and AI to develop new models that can effectively predict the shear capacity of SFRC beams, while overcoming the abovementioned disadvantages. However, despite the emergence of those models, some researchers have focused on some ML methods rather than others for predicting the shear strength of SFRC beams. Support vector machines and artificial neural networks (ANNs), for example, have been widely utilized to predict the mechanical strength of SFRC beams [5,16,17,18,19,20]. Another popular technique utilized for modeling the shear strength of SFRC beams is gene expression programming (GEP) [21,22,23,24,25,26]. However, there is still room for improvement of the prediction of the shear capacity of SFRC beams, even though ANN and GEP algorithms have been widely utilized in various researches concerning the shear capacity prediction of SFRC beams.
In this research, a highly efficient and widely used ML technique, called gradient boosting regression tree (GBRT), is adopted to simulate the process of predicting the shear strength of SFRC beams. GBRT is a powerful ML technique that employs several weak learners and is specifically intended to minimize overfitting issues [27]. Recent research has shown that GBRT shows an excellent prediction performance when compared with other ML algorithms [28,29,30]. Many researches have used GBRT to tackle civil engineering challenges [31,32,33,34].
While some ML-based techniques, such as random forest, neural networks, and support vector machine, can effectively solve regression problems, their operation is difficult to comprehend as these models are often referred to as “black-box” models [35]. As a result, ML-based models should be better described or interpreted to help researchers to better grasp the underlying mechanisms, such as how input parameters impact outputs and increase the persuasiveness of the created models. The Shapley additive explanations (SHAP) framework introduced by Lundberg and Lee [36] can be utilized to understand ML models. In SHAP, features are quantified based on their impact on the predictions. In addition to being able to explain the ML models globally, SHAP can also explain the ML models locally by looking at how the features impact the outputs for a single sample. The research in [37] used GBRT and SHAP for “Understanding the Factors Influencing Pedestrian Walking Speed over Elevated Facilities”. Three ensemble ML models were developed by [38] to predict the creep behavior of concrete, and SHAP was utilized to interpret the predictions of the models. Last but not least, the XGBoost model was developed in reference [39] for load-carrying capacity prediction, and SHAP was used to interpret the ML models.
This study proposes an explainable ML-based technique for predicting the shear strength of SFRC beams. To the best of the authors’ knowledge, the GBRT model is used for the first time to forecast the shear strength of SFRC slender beams. Furthermore, by interpreting the ML model through SHAP, variables impacting the shear strength of the SFRC beams are quantitatively investigated. For this, a database with 330 beam tests, whose shear strength was reported in the literature, was prepared. In addition, the GBRT model’s optimal hyperparameters were identified using a five-fold cross-validation procedure. Additionally, the GBRT model’s performance was compared to other empirical and ML-based equation models presented by other researchers. Finally, the SHAP approach was used to interpret the GBRT model that was built. The effects of several factors on the GBRT model outputs were also explored.

2. Materials and Methods

2.1. Research Methodology

Figure 1 depicts the whole workflow used to develop the suggested approach. The development of a database with SFRC slender beams was the initial stage. After that, some engineering features and filtering methods were applied to the gathered data. The third stage consisted in randomly dividing the data into two sets: one for training and the other for testing. The GBRT model was trained using the training set, and the model was validated using the testing set. The appropriate hyperparameters of the GBRT model were determined by using the five-fold cross-validation procedure during the training stage, this being the fourth stage. The fifth stage includes using the testing set to verify the performance of the model after it has been optimized for the hyperparameters. If the model performance is satisfactory, it can be termed as a final predictive model. During the final process, the SHAP approach is used to understand the model. Quantitatively, the SHAP approach was used to examine how features impact GBRT model predictions on a large dataset (global interpretation) and on a single sample (sample-specific interpretation or local interpretation). As a result, it is possible to analyze the factors influencing the outputs.

2.2. Dataset

The shear strength of SFRC beams without shear stirrups has been studied in several experiments. Ref. [13] recently compiled a wide database with 488 experiments on SFRC beams without stirrups. Non-slender beams with a shear-span to effective depth ratio of a/d < 2.5 and beams with shear-flexural mode failure were filtered out of the initial 488 trials, leaving a subset containing 330 experimental tests. The database with 330 experiments was used to build the model and is summarized in Appendix A. The evaluation database contains rectangular and flanged slender beams. The database specimens failed substantially by shear compression and diagonal stress with an a/d ratio higher than 2.5. The experimental database includes shear beams with varying geometry and reinforcement. Based on several studies [10,11,13,14] and to build an efficient ML model, several critical parameters that affect the shear strength of SFRC beams were chosen. Table 1 shows the statistical properties of the evaluation database’s primary parameters. The primary parameter is the shear strength Vu as the output variable, whereas the beam effective depth d, beam width bw, longitudinal reinforcement ratio ρ, concrete compressive strength fc, aggregate size da, shear span to effective depth ratio a/d, tensile strength of fiber ft, and steel fiber factor Fsf were considered as predictors. The steel fiber factor depends on the percentage volume Vf, diameter df, and fiber length Lf (Equation (1)).
F s f = V f   L f d f
The histograms for the input and output variables from the evaluation database are presented in Figure 2. In general, the database’s range of parameters matches what can be found in real design scenarios, as illustrated in Table 1 and Figure 2. Despite the lack of data for beams with large sizes, the dataset is thought to be representative of most real-world applications and design conditions covered by existing design codes.

2.3. Data-Splitting Procedure

The developed database from the previous section was divided into two parts to implement the ML model: the training dataset and the testing dataset. The GBRT model was developed using the training database, whereas the same predictive model was evaluated using the testing database. As much as possible, a statistically significant association was ensured between inputs of the training and testing datasets while dividing the database into subsets. Most of the developed database (80% of 330 tests) was used for training, while the remaining part was used for model testing (66 tests).
As can noticed from Figure 2, some of the predictors and outcome variables do not obey the normal distribution curve. As a result, these variables need a feature transformation to prevent larger numeric ranges from dominating smaller numeric ranges [40]. In this case, the log transformation was applied to bring right- or left-skewed distributions to approximately normal distributions. Figure 3 shows the distribution of beams’ effective depths (d) before and after log transformation.

2.4. GBRT Model Development

GBRT uses a statistical boosting method to improve the classic decision tree approach. In this method, instead of creating a single “optimal” model, this strategy aggregates several “weak” models to generate a single “strong” consensus model [41]. When using GBRT, the existing residuals are used to build new decision trees sequentially. Fundamentally, this is a form of a functional gradient descent approach for creating sequential models. Adding a new tree at each stage reduces the loss function, thereby improving the prediction [42].
Training data are assumed to exist in the form of a training set { ( x i , y i ) } i = 1 N , in which x i represents the input features and y i represents the shear capacity. For example, the squared error, the absolute error, the Huber error, etc., are all possible loss functions L ( y , F ( x ) ) that can be used to measure how much the predicted F ( x ) differs from the true shear strength y . The GBRT framework assumes that D decision trees will be built, and hence it begins with an initial model F 0 ( x ) . For each iteration d = 1, 2, …, D, compensating the residues is equivalent to optimizing the expansion coefficients ρ d and α d as shown in Equation (2):
( ρ d , α d )   =     argmin ρ , α i = 1 N L [ y i , F d 1 + ρ h ( x i ; α ) ]
where argmin is an operation that finds the argument that gives the minimum value from a target function, and L [ y i , F d 1 ] is a pre-selected feasible loss function measuring the amount of how the predicted value F ( x ) deviates from the true response y. The weighting coefficients and the base learners are fitted to the training data x in a greedy manner as follows:
F d ( x )   =     F d 1   +   ρ d h ( x ; α d )
Equation (2), on the other hand, is difficult to solve directly. Even so, since the gradient-boosting model is additive, ρ h ( x i ; α ) may be seen as an increment along h ( x i ; α ) . It is possible to find the optimum α d using the least squares method, based on the principle of gradient descent:
α d   =   argmin α , β i = 1 N [ r i β h ( x i ; α ) ] 2
where β is a weight factor and r i is the negative gradient evaluated using the previous model.
r i = [ L ( y i , F ( x i ) ) F ( x i ) ] F ( x ) = F d 1 ( x ) , i = 1 , , N
One-dimensional optimization can be used to further improve the gradient-descent step size or weight of the obtained decision tree:
ρ d = argmin ρ i = 1 N L [ r i , F d 1 + ρ h ( x i ; α d ) ] 2
Finally, according to Equation (3), the prior model will be added to the newly evaluated residue model. Algorithm 1 represents the pseudocode for the generic gradient boosting.
Algorithm 1. The gradient boosting algorithm.
Input the iteration number D, loss function   L ( y , F ( x ) ) training set   { ( x i , y i ) } i = 1 N
Initialize: F0 = argmin ρ 0 i = 1 N L ( y i , ρ 0 )
For d = 1 to D do:
r i   =   [ L ( y i , F ( x i ) ) F ( x i ) ] F ( x ) = F d 1 ( x ) , i   =   1 , , N .
α d   =   argmin α , β i = 1 N [ r i β h ( x i ; α ) ] 2
ρ d = argmin ρ i = 1 N L [ r i , F d 1 + ρ h ( x i ; α d ) ] 2 .
F d ( x )   =     F d 1 + ρ d h ( x ; α d )
end for
Output: the final regression function Fd(x)
Gradient boosting allows for a wide variety of smooth loss functions, including AdaBoost, LogitBoost, and L2Boosting [43]. Because of its simplicity and coherence in solving regression problems, the squared loss function is employed in this study:
L ( y , F D ( x ) ) = i = 1 N ( y i F D ( x i ) ) 2 .
Regularization techniques are typically used during the training stage to reduce overfitting and to boost the model’s generalization capacity. In the following equation, Gradient boosting uses a new variable called ν d to regulate the model’s update rate, which is known as shrinkage or learning rate:
F d ( x )   =   F d 1 ( x )   +   ν d ρ d h ( x ; α d ) ,   0 < ν d < 1
The model is updated more slowly when ν d is smaller. According to [44], utilizing small learning rates leads to better model generalization without shrinkage; however, this comes at the cost of greater computing time because more decision trees are required. In addition, numerous additional parameters that are strongly related to the final tree’s structure and model complexity, such as depths (maximum number of splits) and the number of trees D, must be fine-tuned to maximize the performance of the model.

2.5. Cross-Validation

The division of the complete dataset into three subsets—training, validation, and testing—is a standard approach for evaluating the performance of ML models. While the training set is used to complete the learning process, the validation set tracks the performance of the model. As a final step, the model’s extrapolation skills are tested by running it through a set of samples that it has never seen before (testing set) [40]. However, dividing data into three subsets reduces the size of the dataset, which might result in an inadequately trained model. As a result, cross-validation is a typical strategy for avoiding over-reduction of the training set, particularly for small datasets [40]. Cross-validation is performed in various ways, the most common of which is omitting random data to verify the model. K-fold cross-validation was used in this research. Cross-validation with K-fold is a resampling technique that divides data into k subsets, one for validation and the other k-1 for training.

2.6. Hyperparameter Tuning

The tuning of hyperparameters is an essential step in developing reliable ML models. Tuning an ML model reduces overfitting and increases the model adaptability to new data [45]. Choosing the best hyperparameters is also a key component in improving the accuracy of the model [46]. Many ways to automate hyperparameter selection have been developed to prevent manual tuning, including grid search and random search hyperparameter optimization [47]. The domain of the possible values evaluated in the search effort distinguishes these techniques from each other. Random search methods choose distinct hyperparameter values randomly for a given number of iterations, while grid search investigates all potential values in a pre-defined domain for the hyperparameters [47]. The Scikit-learn package in Python [48] was used to explore possible values of hyperparameters using a grid search technique with five-fold cross-validation (GridSearchCV). Figure 4 depicts the five-fold cross-validation used in this work for training and for the hyperparameter selection of the model.

2.7. Performance Metrics

2.7.1. Model Performance Metrics

Various statistical measures, such as R2, mean absolute error (MAE), root mean squared error and (RMSE), were used to evaluate the performance of the built ML-based models. For example, the best model has an R2 value close to 1, while RMSE and MAE values are close to zero. In order to obtain the MAE value, the absolute difference between actual and predicted values must be averaged. The equation for MAE is the following one:
M A E     =     ( i = 1 n | y i o b s y i p r e | n )
When the R2 value is 1, the predicted and true/actual values are perfectly aligned. R2 has the following mathematical representation:
R 2 = i = 1 n ( y i o b s y o b s ) 2 i = 1 n ( y i o b s y i p r e ) 2 i = 1 n ( y i o b s y o b s ) 2 [ 0 , 1 ] .
where yiobs and yipre are the actual output and predicted values, respectively, and y−obs is the average of all observed data.
The difference between the predicted and actual values is the error and the RMSE is calculated as the square root of the average squared errors. The RMSE is computed as follows:
R M S E = 1 n i = 1 n ( y i o b s y i p r e ) 2

2.7.2. Model Uncertainty Metrics

The model uncertainty, or standard deviation (scatter) of the model error, and the mean (bias) are used to evaluate the built models. Due to the lack of knowledge of the problem, conservative assumptions, and mathematical simplifications, the model uncertainty is defined as a model inability to effectively reflect and express a physical phenomenon (in this case, the shear strength). The model uncertainty is described as a random variable with a standard deviation, mean value, and probability distribution in the structural reliability framework. Shear reliability analysis has been proven to be significantly impacted by it. In this work, the predictive model uncertainty related to beam x is equal to the ratio between the experimental and the predicted shear strength, as stated in Equations (12) and (13).
M x = R exp , R p r e d , x ( X )
M ( μ M , σ M )
For a single beam test x, Mx is the model uncertainty. The predictors for the GBRT model (a/d, d, bw, ρ, fc, ft, da, and Fsf) are represented by X. The mean and standard deviation of the model uncertainty are represented by σ M and μ M , respectively.
It is better to choose a model with a mean μ M close to 1 and a standard deviation σ M close to 0 for the model uncertainty. μ M > 1 indicates that the model underestimates the shear capacity of the beam specimen and, consequently, underestimates its failure load. However, if μ M <   1 , it suggests that the model overestimates the shear resistance.

2.8. Shapley Additive Explanations (SHAP) Framework

Lately, Explainable black-box ML models have attracted more study interest because they allow users to trust the created ML models by helping them to comprehend the ML models’ involved mechanism. SHAP is a method for explaining “black-box” ML models. Lundberg and Lee [36] were the first to suggest SHAP, which is based on the notion of Shapley game theory. The SHAP seeks to assess the contribution of each input variable or feature to the observation, and it can determine whether the contribution of each feature is positive or negative. To help with the global and local explanation of ML models, SHAP can calculate the contribution from each feature for every observation. SHAP creates a model of explanation that can be written as:
g ( z ) = ϕ 0 +   j = 1 K ϕ j z j
where z { 0 , 1 } K and K represent the number of input features; ϕ j is the SHAP value for the j–th feature; ϕ 0 is the constant if all inputs are missing.
The SHAP value for the j–th feature can be calculated as:
ϕ j = S F \ { i } | S | ! ( | F | | S | 1 ) ! | F | ! [ f S { i } ( x S { i } f S ( x S ) ) ]
where F is the set of all features and xS is the value of the input.

2.9. Programming Languages and Softwares

In this research, the Python programming language combined with the Scikit-learn library was used to build the system for the estimation of the shear capacity and the interpretation of the GBRT model. Python is a high-level, easy-to-learn, open-source, extensible, and object-oriented programming language (OOP). Python is also an interpreted and versatile language widely used in many fields, such as for building independent programs using graphical interfaces and web applications. In addition, it can be used as a scripting language to control the performance of many programs. It is often recommended for beginners in programming to learn this language because it is among the fastest programming languages to learn [49].
On the other hand, Scikit-learn [48] is an ML library in Python. It contains many algorithms and methods used in the field of ML, such as classification, clustering, and regression, in addition to being used in the stages of data processing and model evaluation. It was built based on the libraries of Scipy, Numpy, Matplotlib, and many others. This study implemented the data preprocessing, filtering techniques, and GBRT modeling using the Python programming language and the Scikit-learn library. At the same time, the plots and figures were created using OriginLab software.

3. Model Results

3.1. K-Fold Cross-Validation

Before running the model, there is the need to fine-tune several of GBRT’s hyperparameters. The hyperparameters of the GBRT model were optimized using a grid search process and a five-fold cross-validation. The most critical hyperparameters for the GBRT model are the n estimators and the learning rate, representing the number of the model’s weak learners and the weights assigned to each estimator, respectively. Additionally, the GBRT model prediction performance can be considerably affected by its max depth parameter, which indicates the complexity of each tree, and its subsample parameter, which represents the fraction of samples to be used for fitting the individual base learners [50]. The tuned values for each of the four hyperparameters are shown in Table 2. The coefficient of determination (R2) was closely examined as a statistical error to obtain hyperparameters with the maximum accuracy while minimizing over-fitting. To execute GBRT modeling and tuning, the Scikit-learn program [48] was used.
A total of 264 data records was used to train the GBRT model, and 66 samples were used to test it. The five-fold cross validation results are shown in Figure 5. Again, there is no noticeable fluctuation in the results of the five folds, and the overall accuracy remains excellent. For example, Fold 1 has a minimum R2 value of 0.9580, and Fold 2 has a maximum R2 value of 0.9852. Table 3 provides the full statistical breakdown of the folds’ results. The coefficient of variation (COV) is only 1.1246% based on the average R2 of 0.9692 and the standard deviation (SD) of 0.0109.

3.2. GBRT Performance on Testing Set

The prediction performance of the proposed method can be tested after the hyperparameters have been identified. The prediction findings are shown in Figure 6, with the X and Y axes representing the experimental and predicted shear strengths, respectively. The training and testing outcomes are represented by the blue dot and red triangle, respectively. In most cases, the difference between the predicted and actual shear strength is within a margin of error of 20% or less. There were three further iterations of the experiment, each using a different mix of training and testing datasets. The predictions for all four experiments are reported in Table 4. The testing RMSE and MAE were always less than 30 and 17, respectively. The mean absolute percentage error, or MAPE, was less than 14%. That is to say, for every sample and instance, the deviation between the predicted and actual shear strength was less than 17 kN (equivalently 14%). These results show that the GBRT approach can be considered an effective tool for estimating the shear capacity of SFRC beams.
Figure 7 presents the histogram of the predicted shear strength (Vpred) from the GBRT model compared to the actual shear strength (Vact) (case 1). Again, most of the shear strengths predicted by GBRT are within a margin of 20% or less of error. The standard deviation ( σ M ) and mean value ( μ M ) of the ratio Vact/Vpred are taken into consideration when evaluating the accuracy of the GBRT model. The standard deviations ( σ M ) for the training and testing data were 0.058 and 0.145, respectively, whereas the mean values ( μ M ) were 1.002 and 0.980, respectively. A normally distributed relationship between the GBRT-predicted values and the experimental data shows that the error is dispersed randomly.

3.3. The Reliability of the GBRT Model Prediction

For a total of 330 SFRC beams, the statistical evaluation of the various models [1,21,26,51,52,53] used to estimate the shear strength, and also for the model developed in the present study, can be found in Table 5. The formulas from the models used for the comparative analysis can be found in Table 6. Ashour et al. [52] proposed two sets of equations based on a regression model for observed data gathered from 18 high-strength SFRC beam specimens. An essential parameter, the fiber factor (F), which accounts for the influence of steel fiber size and shape, was incorporated into the first equation, taken from the ACI Building Code’s shear equation. In addition, to account for the role of reinforcement and concrete in the shear capacity, the authors incorporated the shear span to effective depth (a/d) ratio in their equation. The second equation of Ashour et al. [52] is based on a modified version of Zsutty’s equation [54], which includes the fiber factor. Deep beams ( a d   <   2.5 ) and slender beams ( a d     2.5 ) have unique formulas in the two sets of the second equation. A modified version of the ACI Building Code equation was also established for the shear capacity by Khuntia et al. [1]. In their equation, the effect of fiber is incorporated. Khuntia et al. [1] used the post-cracking tensile properties of fiber reinforced concrete to build the equation. The experimental data of 68 SFRC beam specimens were used to validate the equation. An equation presented by Sharma [51] omits some of the essential parameters, such as the ratio lf/df and F, which substantially impact the shear capacity of SFRC. The referred author used 41 experiments to validate his equation. Rather than incorporating the actual reinforcement ratio, the equation from Greenough and Nehdi [21] simplifies a formula derived from genetic programming by using a percentage for ρ. Additionally, 208 SFRC beam test results from earlier research were analyzed using multi-expression programming to obtain the formula presented by Sarveghadi et al. [26]. The authors have produced two sets of equations: one set contains expressions specific to high-strength concrete, and the other set is a composite equation for both types of concrete (normal- and high-strength). An equation for predicting the shear strength of SFRC beams based on 293 previous experiments was recently published by Sabetifar and Nematzadeh [53]. The previous two studies [26,53] built their models based on genetic programming (GP). GP is a machine learning-based approach for developing nonlinear regression. The Darwinian ideas of natural selection and genetic spreading of features chosen by biologically growing organisms are the foundations of GP. Even though both researches [26,53] were published recently, the datasets utilized to train and test the models were quite constrained, resulting in models with only limited application. The equations for the shear capacity of SFRC beams proposed in the previous referred studies are given in Table 6.
The mean value for the ratio of the experimental values to the model predicted values ( μ M ) and their variability (CV) was used to evaluate the performance of the model. The prediction is more accurate when μ M is near to 1 and when the CV is low. As can be observed in Table 5, five of the μ M values are higher than 1.00, implying that the models from [1,21,51,52] underestimate the shear capacity of SFRC beams, while the models from [26,53] slightly overestimate the shear resistance. The explanation for this observation might be linked to the fact that the previously referred equations were produced based on a limited set of data with low variation between specimens’ properties. Furthermore, the equations proposed in the referred literature omit some critical factors that contribute to the shear strength of SFRC.
With μ M = 0.996, STD = 0.08, and COV = 12%, it can be stated that the proposed model in this research beat all previous models. As a result, the model has a reduced error rate and a higher degree of linearity between the anticipated and actual values. Furthermore, the GBRT model has the lowest coefficient of variation (COV) when compared to the other models, indicating that its projected values have the slightest variance around the mean.
The experimental to prediction ratios for each input variable are presented in Figure 8, in order to check if a bias exists between the prediction of the GBRT model and one or more input variables. From Figure 8, it seems that a significant trend or preference toward these variables does not exist. The accuracy of the GBRT prediction for the shear strength seems robust. This indicates that the proposed model can be employed with high confidence within the ranges of independent variables used to construct the model.

3.4. Interpretation of the GBRT Model

The SHAP approach was used to understand the developed GBRT model and how its inputs impact its outputs. The summary of the SHAP values and the feature importance factor for all of the input features can be seen in Figure 9. Each dot on the graphs represents a dataset instance and its corresponding feature SHAP value. The x-axis indicates each feature’s effectiveness on the dependent variable, while the y-axis shows the model’s ranking of features by significance. A red dot denotes a high feature value, corresponding to a higher SHAP value. The significance of each feature is determined as the mean absolute SHAP values for the whole dataset, as shown in Figure 9a.
Figure 9a shows that both the beam width (bw) and the effective depth (d) are the most important factors, followed by the concrete strength (fc) and the longitudinal reinforcement ratio (ρ). In addition, the outputs are also affected by the steel fiber factor (Fsf) and the shear-span to effective depth ratio (a/d). Finally, the fiber tensile strength (ft) and the aggregate size (da) have the lowest effect. According to the results in Figure 9a, bw and d have the ability to alter the estimated value of shear strength by an average of 16%, while ft and da have the lowest ability with only about 1%. As can be seen in Figure 9b, most of the features mentioned above positively influence the model outcome, which indicates that when one of those features increases, the shear capacity of the SFRC slender beam increases. The only exception among those features is a/d. This observation can be explained due to the influence of the arch action, which depicts the compressive force created along with the beam supports and the loading points. Loads are borne in part by the arch action in the area of small shear spans. The applied shear is resisted by the arch action, which leads to a decreased shear for higher a/d [55]. The conclusions presented in this section can assist constructers and designers in determining the importance of each feature in SFRC slender beams for the output shear strength, and whether it is positive or negative.

4. Conclusions

An investigation on the use of an explainable ML method for the prediction of the shear strength of SFRC slender beams was conducted in this study. Using SHAP to interpret the ML model, the factors impacting the shear strength were examined. A database with 330 SFRC slender beam tests was created and randomly divided into testing and training sets. Using a five-fold cross-validation procedure paired with a grid search strategy, optimal hyperparameters of the GBRT model were found based on the training dataset. The testing dataset was used to validate the performance of the built GBRT model. Meanwhile, six empirical and machine learning-based equation models were chosen and compared to comprehensively analyze the performance of the GBRT model. Additionally, to analyze the GBRT model globally across the whole dataset, the SHAP approach was used. The SHAP values were used to discuss factors that impact the model results. The following are the main key conclusions that can be derived from the research findings:
  • The GBRT model predicts the shear capacity of SFRC slender beams with high accuracy. The model has R2 values of 0.963 and 0.972 for the testing and training sets, respectively. In addition, both the training and testing sets of the GBRT model have low RMSE and MAE values, indicating that the prediction capability of the GBRT model can be trusted with high confidence;
  • A comparison between the predicted and experimental shear strengths was also performed, using previously established equations from the literature. The results show that the predicted values from previous models do not apply to a wide range of data and have a high variance;
  • Most of the proposed equations from the literature show a mean value for the model uncertainty larger than 1, implying that they all underestimate the shear capacity of the SFRC slender beams from the database;
  • With low error measurements and μ M near unity, the results showed that the GBRT method surpassed the other models mentioned in this study.

Author Contributions

A.S.—conceptualization, modeling, and write up. H.M.A.—visualization and review. N.M.A.-A.—writing, validation and supervision. E.A.A.-H.—review, editing and visualization. L.F.A.B.—review and writing. H.I.—review, writing original draft and funding. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not Applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

S.Nobw (mm)d (mm)ρa/dda (mm)fc (MPa)ft (MPa)FstVu (kN)
11502510.02673.4912.528.111000.49113
21502510.02673.4912.525.311000.4979
31502510.02673.4912.527.911000.65109
41502510.02673.4912.526.211000.65123
51502510.02673.4912.528.111000.98111
61502510.02673.4912.527.311000.98131
71502510.02673.4912.527.510500.4065
81502510.02673.4912.524.910500.4077
91502510.02673.4912.527.810500.6091
101502510.02673.4912.527.310500.60102
111502510.02673.4912.526.310500.80116
121502510.02673.4912.527.110500.80105
131502510.02673.4912.553.411000.49113
141502510.02673.4912.554.111000.49126
151502510.02673.4912.553.211000.65144
161502510.02673.4912.555.311000.65166
171502510.02673.4912.564.611000.98195
181502510.02673.4912.559.911000.98160
191502510.02673.4912.547.810500.40128
201502510.02673.4912.549.510500.40152
211502510.02673.4912.555.310500.60146
221502510.02673.4912.556.410500.60178
231502510.02673.4912.553.410500.80128
241502510.02673.4912.551.010500.80157
251502510.02673.4912.527.810250.3879
261502510.02673.4912.527.210250.3878
271502510.02673.4912.527.610500.6499
281502510.02673.4912.527.910500.6481
291502510.02673.4912.534.710250.3899
301502510.02673.4912.536.210250.38100
311502510.02673.4912.537.010500.64110
321502510.02673.4912.538.310500.64104
331502610.01953.4520.032.911000.60108
341502610.01953.4520.023.811000.8093
351502610.01953.4520.024.111001.00114
361401750.01282.5012.082.011000.4063
371401750.01282.5012.083.211000.8079
381401750.01282.5012.083.811001.20135
391502000.01342.5022.033.711000.5565
401502000.01342.5022.024.511000.5544
411502000.01342.5022.021.411001.0950
421502000.01342.5012.09.811001.6439
431502000.01343.5022.020.211000.5533
441502000.01343.5022.021.411001.0943
451502000.01343.5012.027.911001.6459
461502000.01344.5022.024.511000.5543
471523810.02713.4010.049.211000.80172
481523810.02713.4010.031.011000.90148
491523810.02713.4010.044.911000.90189
501523810.02713.4010.044.911000.90190
511523810.02713.4010.049.211000.80218
521523810.02713.4010.031.011000.90195
531523810.02713.5010.038.111000.60147
541523810.02713.5010.038.111000.60200
551523810.01973.5010.038.111000.60175
561523810.01973.5010.038.111000.60179
572002600.01812.5010.040.011000.17108
582002600.01812.5010.038.711000.51144
592002600.01152.5010.040.011000.1782
602002600.01152.5010.038.711000.51107
612004600.02803.4010.037.711000.34244
622004600.02803.4010.038.811000.34252
632004600.02803.4010.037.711000.34259
642004600.02803.4010.037.711000.34263
652002600.03563.5010.046.911000.17110
662002600.03563.5010.043.711000.34120
672002600.03563.5010.048.311000.51155
682002600.02833.5010.037.711000.34111
692002600.02833.5010.038.811000.34132
702005400.02733.5010.037.711000.17153
712005600.02733.5010.038.811000.34230
722002600.01814.0010.041.211000.1782
732002600.01814.0010.040.311000.51117
741502170.01852.9510.035.011000.6084
753006220.01982.8110.034.023000.21274
763006220.01982.8110.036.023000.45344
77851300.02052.529.651.920000.1930
78851300.02053.029.651.920000.1931
79851300.02052.529.633.320000.1923
80851300.02053.029.633.320000.1921
81851300.02053.029.651.720000.5036
82851300.02053.029.630.620000.5022
83851300.02053.029.631.020000.7533
84851300.02052.529.651.720000.5041
85851300.02053.529.641.720000.5029
86851300.02052.529.648.720001.0049
87851300.02053.529.648.820001.0033
88851280.03703.069.641.720000.5032
89851260.05723.119.641.720000.5038
90851280.03703.069.630.620000.5024
91851260.05723.119.630.620000.5025
92851280.03703.069.648.820001.0048
93851260.05723.119.648.820001.0054
94851260.05723.119.653.620001.1352
95851260.05723.119.643.220001.5053
96851280.03703.069.653.620001.1349
971502190.01912.8010.040.911150.6096
981502190.01912.8010.040.911151.20103
991252120.01523.0019.030.810790.3168
1001001300.03093.0810.038.713030.3058
1011001300.03093.0810.042.413030.6074
1021523810.01963.4410.044.811000.41171
1031523810.01963.4410.044.811000.41160
1041523810.01963.4410.038.111000.55169
1051523810.01963.4410.038.111000.55172
1061523810.02633.4410.031.011000.83148
1071523810.02633.4410.031.011000.83196
1081523810.02633.4410.044.911000.83191
1091523810.02633.4410.044.911000.83189
1101523810.02633.4410.049.211000.80172
1111523810.02633.4410.049.211000.80218
1121523810.01963.4410.043.323000.60193
1131523810.01963.4410.043.323000.60189
1142056100.01963.5010.050.811000.41363
1152056100.01963.5010.050.811000.41335
1162056100.01963.5010.028.711000.60349
1172056100.01963.5010.028.711000.60341
1182056100.01523.5010.042.311000.41345
1192056100.01523.5010.029.611000.60265
1202056100.01523.5010.029.611000.60222
1212056100.01963.5010.044.411000.83432
1222056100.01963.5010.042.811001.20418
1231503400.03082.5012.558.911500.65260
1241503400.03082.5012.551.711501.30291
1251507350.01063.8112.542.012000.94352
1261507350.01063.8112.538.012000.75352
1271252250.03492.8910.090.012000.75157
1281502020.01172.9710.021.311000.2848
1291502020.01172.9710.019.611000.5557
1303004370.01503.0910.021.311000.28154
1313004370.01503.0910.019.611000.55198
1322004350.01042.5120.024.811000.19129
1332004350.01042.5120.033.511000.19115
1342004350.01042.5120.033.513330.33137
1352004350.01042.5120.038.611000.19136
1362004550.00992.5115.024.411000.13154
1372009100.01042.5020.024.411000.13247
1382009100.01042.5020.055.011000.13328
1391252100.01534.0019.044.611000.3135
1401252250.03492.8910.090.012000.75138
1411252250.03492.8910.090.012000.75138
1421522210.01202.5010.034.011300.3058
1431522210.02392.5010.034.011300.6083
1441522210.02392.5010.034.011300.3064
1451522210.02393.5010.034.011300.3049
1461501970.01362.8020.029.112600.3053
1471501970.01363.6020.029.112600.3045
1481501970.01362.8020.029.912600.4560
1491501970.02042.8020.029.912600.4565
1501501970.01362.8020.020.612600.4545
1511501970.02042.8020.020.612600.4560
1521501970.02042.8020.033.412600.4586
1531522540.02483.5010.029.010960.50120
1546102540.02473.5010.029.010960.50478
1551523940.02863.6110.039.010960.50161
1561523940.02863.6110.039.010960.50194
1572035410.02543.4510.050.010960.50267
1582035410.02543.4510.050.010960.50380
1592548130.02703.5010.050.010960.50683
1602548130.02703.5010.050.010960.50704
16130511180.02553.5010.050.010960.501045
16230511180.02553.5010.050.010960.501008
1632001800.04473.3316.090.626000.20299
1642001800.04473.3316.083.218500.36295
1652001800.04473.3316.080.522000.43252
1662001800.04473.3316.080.522000.64262
1672001950.03093.0816.039.418500.36189
1682002350.04282.7716.091.411000.50310
1692002350.04282.7716.093.326000.20363
1702002350.04282.7716.089.618500.36407
1712004100.03062.9318.076.826000.20289
1722004100.03062.9318.076.826000.20336
1732004100.03062.9318.072.018500.36367
1742004100.03062.9318.072.018500.36327
1752004100.03062.9318.069.322000.43264
1762004100.03062.9318.069.322000.43312
1772004100.03062.9318.060.222000.64339
1782004100.03062.9318.075.722000.64292
1793005700.02872.9818.076.826000.20445
1803005700.02872.9818.072.018500.36596
1813005700.02872.9818.060.222000.64509
1822003140.03503.500.4131.520000.75251
1832003140.03503.500.4154.520000.75318
1842003140.03503.500.4145.620000.75357
1852003140.03503.500.4132.820000.38266
1862003140.03503.500.4143.320000.38199
1872003140.03503.500.4152.920000.38308
1881252150.00374.0010.092.62600.7524
1891252150.00376.0010.093.72600.7515
1901252150.02834.0010.095.42600.3861
1911252150.02836.0010.095.82600.3852
1921252150.02834.0010.097.52600.7585
1931252150.02836.0010.0100.52600.7553
1941252150.02834.0010.097.12601.1394
1951252150.02836.0010.0101.32601.1353
1961252150.04584.0010.093.82600.75104
1971252150.04586.0010.095.02600.7579
1981403400.01672.5019.036.011000.60154
1991503500.05612.862.0121.120000.52340
2001503500.05612.862.0120.320001.04531
2012603400.01724.0010.021.013360.45114
2022603400.01724.0010.056.013360.45204
203641020.02203.002.453.010000.2117
2041272040.02213.002.453.010000.2151
205641020.02203.002.450.210000.4321
2061272040.02213.002.450.210000.4366
207641020.02203.002.462.610000.2118
2081272040.02213.002.462.610000.2161
209641020.02202.502.462.610000.2121
210641020.02202.752.462.610000.2118
211641020.01103.002.462.610000.2113
212641020.03303.002.462.610000.2118
213641020.03303.002.454.110000.4325
2141272040.02213.009.022.711720.6079
215641020.02203.009.022.711720.6020
216641020.01103.009.022.711720.6016
2171272040.02213.009.026.011721.0079
218641020.02203.009.026.011721.0023
219552650.04313.4314.041.915700.7559
220552650.04314.9114.036.915700.7543
221552650.02763.4314.033.915700.7546
2222002650.01783.0210.047.911000.2591
2232002650.01783.0210.038.011000.38106
2242002650.01783.0210.042.211000.50149
2252002650.01783.0210.045.411000.13115
2262002650.01783.0210.044.411000.19144
2272002650.01783.0210.040.311000.25147
2282002650.01783.0210.053.711000.11107
2292002650.01783.0210.046.011000.16123
2302002650.01783.0210.042.211000.21151
2312003100.01132.559.539.811000.30131
2322002850.03332.779.539.811000.30220
2332002600.03553.4614.046.411000.16110
2342002600.03553.4614.043.211000.33120
2352002600.03553.4614.047.611000.49155
2362002600.01812.5014.039.111000.16108
2372002600.01812.5014.038.611000.49144
2382002600.01814.0414.040.711000.1683
2392002600.01814.0414.042.411000.49117
2402002600.01812.5014.026.511000.11100
2412002600.01812.5014.027.211000.34120
2422002600.01812.5014.046.811000.33158
2431752100.04014.5010.036.410500.3080
2441752100.04014.5010.038.410500.60114
2451752100.04014.5010.040.810500.90115
2461752100.04014.5010.038.510500.6069
2471011270.03094.402.033.211000.1132
2481011270.03094.202.033.211000.1131
2491011270.03094.202.033.211000.1128
2501011270.03094.202.033.211000.1125
2511011270.03094.302.033.211000.1130
2521011270.03094.302.033.211000.1128
2531011270.03094.002.040.211000.2233
2541011270.03094.002.040.211000.2231
2551011270.03094.002.040.211000.2233
2561011270.03094.402.033.211000.1128
2571011270.03094.402.033.211000.1127
2581011270.03094.002.033.211000.1030
2591011270.03094.002.033.211000.1030
2601011270.03094.002.033.211000.1033
2611011270.03094.602.033.211000.1026
2621011270.03094.402.033.211000.1027
2631011270.03094.402.033.211000.1026
2641011270.03095.002.033.211000.1024
2651011270.03094.802.033.211000.1022
2661011270.03094.002.040.211000.2031
2671011270.03094.202.040.211000.2034
2681011270.03094.202.040.211000.2030
2691011270.03094.202.040.211000.2032
2701011270.03093.202.039.711000.4137
2711011270.03093.402.039.711000.4134
2721011270.03093.402.039.711000.4133
2731011270.03093.402.039.711000.4142
2741011270.03093.402.039.711000.4139
2751011270.03094.802.033.211000.1024
2761011270.03094.802.033.211000.1023
2771011270.03094.802.033.211000.1026
2781001270.01993.602.020.749130.1321
2791001270.01993.602.020.723500.4229
2801001270.01994.802.020.723500.4224
2811001750.03593.0013.080.018560.2556
2821001750.03593.0013.080.018560.5072
2831001750.03594.5013.080.018560.2549
2841001750.03594.5013.080.018560.5060
2852003000.03082.5010.0110.020000.56284
2862003000.03083.5010.0111.520000.56209
2872003000.03084.5010.0110.820000.56212
2881522830.01992.509.533.111001.00136
289152.42830.01992.509.533.211001.00145
2901522830.01992.509.533.011002.00134
2911522830.01992.509.534.411002.00138
2921001660.03433.0210.039.412000.3031
2931001660.03433.0210.039.212000.6052
2941001660.03433.0210.040.012000.9054
2951001660.03433.0210.035.512001.2048
2961001590.04783.1410.058.012000.6074
2971001590.04783.1410.080.112000.3073
2981001590.04783.1410.088.012000.6081
2991502190.01912.8010.080.011000.55114
3001252120.01523.7710.059.411000.2743
3011252120.01523.7710.049.611000.4045
3021252100.02283.8110.049.711000.4144
3031252100.02283.8110.051.511000.5558
3041252100.02283.8112.054.511000.5559
3051001400.01122.5012.536.111000.3141
306100850.01663.5210.054.811000.9520
307100850.01663.5210.049.311001.4322
308100850.01663.5210.049.311001.4319
309100850.01663.5210.053.711002.8620
310100850.01663.5210.053.511000.7123
311100850.01663.5210.053.511000.7118
3122002730.03482.7522.0110.910000.48201
3132002730.03482.7522.0109.210000.50209
314801650.01712.994.041.28000.5033
315801650.01712.994.039.98000.7541
3163004200.03223.2120.062.314000.49411
3174506480.03273.2620.062.314000.49793
3186008870.03433.2620.062.314000.491430
319702700.03322.5610.050.011000.3381
3201102700.02122.5610.050.011000.3396
3211502700.01552.5610.050.011000.33109
3223102580.02503.0010.023.011000.55210
3233102400.04033.0010.041.011000.55280
3243005310.01883.0010.023.011000.55248
3253005230.02553.0010.023.011000.55238
3263005230.02553.0010.041.011000.55440
3273009230.01443.0010.041.011000.55479
3283009200.02033.0010.041.011000.55484
3293009230.01443.0010.080.011000.55633
3303009200.02033.0010.080.011000.55631

References

  1. Khuntia, M.; Stojadinovic, B.; Goel, S.C. Shear strength of normal and high-strength fiber reinforced concrete beams without stirrups. Struct. J. 1999, 96, 282–289. [Google Scholar]
  2. Shahnewaz, M.; Alam, M.S. Genetic algorithm for predicting shear strength of steel fiber reinforced concrete beam with parameter identification and sensitivity analysis. J. Build. Eng. 2020, 29, 101205. [Google Scholar] [CrossRef]
  3. Dinh, H.H. Shear Behavior of Steel Fiber Reinforced Concrete Beams without Stirrup Reinforcement. Ph.D. Thesis, University of Michigan, Ann Arbor, MI, USA, 2009. [Google Scholar]
  4. Raju, R.A.; Akiyama, M.; Lim, S.; Kakegawa, T.; Hosono, Y. A novel casting procedure for SFRC piles without shear reinforcement using the centrifugal forming technique to manipulate the fiber orientation and distribution. Constr. Build. Mater. 2021, 303, 124232. [Google Scholar] [CrossRef]
  5. Yaseen, Z.M.; Tran, M.T.; Kim, S.; Bakhshpoori, T.; Deo, R.C. Shear strength prediction of steel fiber reinforced concrete beam using hybrid intelligence models: A new approach. Eng. Struct. 2018, 177, 244–255. [Google Scholar] [CrossRef]
  6. Kim, K.S.; Lee, D.H.; Hwang, J.-H.; Kuchma, D.A. Shear behavior model for steel fiber-reinforced concrete members without transverse reinforcements. Compos. Part B Eng. 2012, 43, 2324–2334. [Google Scholar] [CrossRef]
  7. Narayanan, R.; Darwish, I. Use of steel fibers as shear reinforcement. Struct. J. 1987, 84, 216–227. [Google Scholar]
  8. Gandomi, A.; Alavi, A.; Yun, G. Nonlinear modeling of shear strength of SFRC beams using linear genetic programming. Struct. Eng. Mech. 2011, 38, 1–25. [Google Scholar] [CrossRef]
  9. Özcan, D.M.; Bayraktar, A.; Şahin, A.; Haktanir, T.; Türker, T. Experimental and finite element analysis on the steel fiber-reinforced concrete (SFRC) beams ultimate behavior. Constr. Build. Mater. 2009, 23, 1064–1077. [Google Scholar] [CrossRef]
  10. Spinella, N.; Colajanni, P.; Recupero, A. Simple plastic model for shear critical SFRC beams. J. Struct. Eng. 2010, 136, 390–400. [Google Scholar] [CrossRef] [Green Version]
  11. Arslan, G. Shear strength of steel fiber reinforced concrete (SFRC) slender beams. KSCE J. Civ. Eng. 2014, 18, 587–594. [Google Scholar] [CrossRef]
  12. Hanai, J.B.; Holanda, K.M.A. Similarities between punching and shear strength of steel fiber reinforced concrete (SFRC) slabs and beams. IBRACON Struct. Mater. J. 2008, 1, 1–16. [Google Scholar]
  13. Lantsoght, E.O. Database of shear experiments on steel fiber reinforced concrete beams without stirrups. Materials 2019, 12, 917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Lantsoght, E.O. How do steel fibers improve the shear capacity of reinforced concrete beams without stirrups? Compos. Part B Eng. 2019, 175, 107079. [Google Scholar]
  15. Marì Bernat, A.; Spinella, N.; Recupero, A.; Cladera, A. Mechanical model for the shear strength of steel fiber reinforced concrete (SFRC) beams without stirrups. Mater. Struct. 2020, 53, 28. [Google Scholar] [CrossRef]
  16. Keshtegar, B.; Bagheri, M.; Yaseen, Z.M. Shear strength of steel fiber-unconfined reinforced concrete beam simulation: Application of novel intelligent model. Compos. Struct. 2019, 212, 230–242. [Google Scholar] [CrossRef]
  17. Abambres, M.; Lantsoght, E.O. ANN-based shear capacity of steel fiber-reinforced concrete beams without stirrups. Fibers 2019, 7, 88. [Google Scholar] [CrossRef] [Green Version]
  18. El-Chabib, H.; Nehdi, M.; Said, A. Predicting shear capacity of NSC and HSC slender beams without stirrups using artificial intelligence. Comput. Concr. Int. J. 2005, 2, 79–96. [Google Scholar] [CrossRef]
  19. Nehdi, M.; El Chabib, H.; Said, A. Evaluation of shear capacity of FRP reinforced concrete beams using artificial neural networks. Smart Struct. Syst. 2006, 2, 81–100. [Google Scholar] [CrossRef]
  20. Chaabene, W.B.; Nehdi, M.L. Novel soft computing hybrid model for predicting shear strength and failure mode of SFRC beams with superior accuracy. Compos. Part C Open Access 2020, 3, 100070. [Google Scholar] [CrossRef]
  21. Greenough, T.; Nehdi, M. Shear behavior of fiber-reinforced self-consolidating concrete slender beams. ACI Mater. J. 2008, 105, 468. [Google Scholar]
  22. Kara, I.F. Empirical modeling of shear strength of steel fiber reinforced concrete beams by gene expression programming. Neural Comput. Appl. 2013, 23, 823–834. [Google Scholar] [CrossRef]
  23. Mukherjee, S.; Manju, S. An improved parametric formulation for the variationally correct distortion immune three-noded bar element. Struct. Eng. Mech. 2011, 38, 261. [Google Scholar] [CrossRef]
  24. Shahnewaz, M.; Tannert, M. Shear strength prediction of steel fiber reinforced concrete beams from genetic programming and its sensitivity analysis. In Proceedings of the FRC: The Modern Landscape BEFIB 2016 9th Rilem International Symposium on Fiber Reinforced Concrete, Vancouver, BC, Canada, 19–21 September 2016. [Google Scholar]
  25. Slater, E.; Moni, M.; Alam, M.S. Predicting the shear strength of steel fiber reinforced concrete beams. Constr. Build. Mater. 2012, 26, 423–436. [Google Scholar] [CrossRef]
  26. Sarveghadi, M.; Gandomi, A.H.; Bolandi, H.; Alavi, A.H. Development of prediction models for shear strength of SFRCB using a machine learning approach. Neural Comput. Appl. 2019, 31, 2085–2094. [Google Scholar] [CrossRef]
  27. Prettenhofer, P.; Louppe, G. Gradient Boosted Regression Trees in Scikit-Learn. 2014. Available online: https://orbi.uliege.be/handle/2268/163521 (accessed on 23 February 2014).
  28. Zhang, J.; Li, D.; Wang, Y. Toward intelligent construction: Prediction of mechanical properties of manufactured-sand concrete using tree-based models. J. Clean. Prod. 2020, 258, 120665. [Google Scholar] [CrossRef]
  29. Chen, S.-Z.; Zhang, S.-Y.; Han, W.-S.; Wu, G. Ensemble learning based approach for FRP-concrete bond strength prediction. Constr. Build. Mater. 2021, 302, 124230. [Google Scholar] [CrossRef]
  30. Fu, B.; Feng, D.-C. A machine learning-based time-dependent shear strength model for corroded reinforced concrete beams. J. Build. Eng. 2021, 36, 102118. [Google Scholar] [CrossRef]
  31. Xiao, Q.; Li, C.; Lei, S.; Han, X.; Chen, Q.; Qiu, Z.; Sun, B. Using Hybrid Artificial Intelligence Approaches to Predict the Fracture Energy of Concrete Beams. Adv. Civ. Eng. 2021, 2021, 6663767. [Google Scholar] [CrossRef]
  32. Zhang, Z.; Yang, W.; Wushour, S. Traffic accident prediction based on LSTM-GBRT model. J. Control. Sci. Eng. 2020, 2020, 4206919. [Google Scholar] [CrossRef]
  33. Hu, X.; Li, B.; Mo, Y.; Alselwi, O. Progress in Artificial Intelligence-based Prediction of Concrete Performance. J. Adv. Concr. Technol. 2021, 19, 924–936. [Google Scholar] [CrossRef]
  34. Qi, C.; Fourie, A.; Zhao, X. Back-analysis method for stope displacements using gradient-boosted regression tree and firefly algorithm. J. Comput. Civ. Eng. 2018, 32, 04018031. [Google Scholar] [CrossRef]
  35. Feng, D.-C.; Wang, W.-J.; Mangalathu, S.; Hu, G.; Wu, T. Implementing ensemble learning methods to predict the shear strength of RC deep beams with/without web reinforcements. Eng. Struct. 2021, 235, 111979. [Google Scholar] [CrossRef]
  36. Lundberg, S.M.; Lee, S.-I. A unified approach to interpreting model predictions. In Proceedings of the NIPS’17: 31st International Conference on Neural Information Processing Systems, Long Beach, CA, USA, 4–9 December 2017; Volume 30. [Google Scholar]
  37. Banerjee, A.; Raoniar, R.; Maurya, A.K. Understanding the Factors Influencing Pedestrian Walking Speed over Elevated Facilities using Tree-Based Ensembles and Shapley Additive Explanations. Res. Sq. 2021. Available online: https://www.researchsquare.com/article/rs-373997/v1 (accessed on 2 August 2021).
  38. Liang, M.; Chang, Z.; Wan, Z.; Gan, Y.; Schlangen, E.; Šavija, B. Interpretable Ensemble-Machine-Learning models for predicting creep behavior of concrete. Cem. Concr. Compos. 2022, 125, 104295. [Google Scholar] [CrossRef]
  39. Bakouregui, A.S.; Mohamed, H.M.; Yahia, A.; Benmokrane, B. Explainable extreme gradient boosting tree-based prediction of load-carrying capacity of FRP-RC columns. Eng. Struct. 2021, 245, 112836. [Google Scholar] [CrossRef]
  40. Kuhn, M.; Johnson, K. Applied Predictive Modeling; Springer: Berlin/Heidelberg, Germany, 2013; Volume 26. [Google Scholar]
  41. Lin, H.-T.; Liang, T.-J.; Chen, S.-M. Estimation of battery state of health using probabilistic neural network. IEEE Trans. Ind. Inform. 2012, 9, 679–685. [Google Scholar] [CrossRef]
  42. Elith, J.; Leathwick, J.R.; Hastie, T. A working guide to boosted regression trees. J. Anim. Ecol. 2008, 77, 802–813. [Google Scholar] [CrossRef] [PubMed]
  43. Bühlmann, P.; Yu, B. Boosting with the L 2 loss: Regression and classification. J. Am. Stat. Assoc. 2003, 98, 324–339. [Google Scholar] [CrossRef]
  44. Hastie, T.; Tibshirani, R.; Friedman, J.H.; Friedman, J.H. The Elements of Statistical Learning: Data Mining, Inference, and Prediction; Springer: Berlin/Heidelberg, Germany, 2009; Volume 2. [Google Scholar]
  45. Bardenet, R.; Brendel, M.; Kégl, B.; Sebag, M. Collaborative hyperparameter tuning. In Proceedings of the International Conference on Machine Learning, Atlanta, GA, USA, 16–21 June 2013; pp. 199–207. [Google Scholar]
  46. Bergstra, J.; Yamins, D.; Cox, D. Making a science of model search: Hyperparameter optimization in hundreds of dimensions for vision architectures. In Proceedings of the International Conference on Machine Learning, Atlanta, GA, USA, 16–21 June 2013; pp. 115–123. [Google Scholar]
  47. Bergstra, J.; Bengio, Y. Random search for hyper-parameter optimization. J. Mach. Learn. Res. 2012, 13, 281–305. [Google Scholar]
  48. Varoquaux, G.; Buitinck, L.; Louppe, G.; Grisel, O.; Pedregosa, F.; Mueller, A. Scikit-learn: Machine learning without learning the machinery. GetMob. Mob. Comput. Commun. 2015, 19, 29–33. [Google Scholar] [CrossRef]
  49. Srinath, K.R. Python—The fastest growing programming language. Int. Res. J. Eng. Technol. (IRJET) 2017, 4, 354–357. [Google Scholar]
  50. 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]
  51. Sharma, A. Shear strength of steel fiber reinforced concrete beams. J. Proc. 1986, 83, 624–628. [Google Scholar]
  52. Ashour, S.A.; Hasanain, G.S.; Wafa, F.F. Shear behavior of high-strength fiber reinforced concrete beams. Struct. J. 1992, 89, 176–184. [Google Scholar]
  53. Sabetifar, H.; Nematzadeh, M. An evolutionary approach for formulation of ultimate shear strength of steel fiber-reinforced concrete beams using gene expression programming. Structures 2021, 34, 4965–4976. [Google Scholar] [CrossRef]
  54. Zsutty, T.C. Beam shear strength prediction by analysis of existing data. J. Proc. 1968, 65, 943–951. [Google Scholar]
  55. Jeong, J.-P.; Kim, W. Shear resistant mechanism into base components: Beam action and arch action in shear-critical RC members. Int. J. Concr. Struct. Mater. 2014, 8, 1–14. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Workflow used to develop the explainable ML model.
Figure 1. Workflow used to develop the explainable ML model.
Buildings 12 00550 g001
Figure 2. Relative frequency distributions for the input and output variables.
Figure 2. Relative frequency distributions for the input and output variables.
Buildings 12 00550 g002
Figure 3. Histogram of beam effective depths: (a) before log transformation; (b) after log transformation.
Figure 3. Histogram of beam effective depths: (a) before log transformation; (b) after log transformation.
Buildings 12 00550 g003
Figure 4. Hyperparameter tuning using five-fold cross validation (GridSearchCV).
Figure 4. Hyperparameter tuning using five-fold cross validation (GridSearchCV).
Buildings 12 00550 g004
Figure 5. Results for five-fold cross validation.
Figure 5. Results for five-fold cross validation.
Buildings 12 00550 g005
Figure 6. Prediction results: (a) case 1; (b) case 3.
Figure 6. Prediction results: (a) case 1; (b) case 3.
Buildings 12 00550 g006
Figure 7. Histogram for Vact/Vpred (case 1): (a) training set; (b) testing set.
Figure 7. Histogram for Vact/Vpred (case 1): (a) training set; (b) testing set.
Buildings 12 00550 g007
Figure 8. The relationship between the shear input variables and the ratio of experimental to model shear strength.
Figure 8. The relationship between the shear input variables and the ratio of experimental to model shear strength.
Buildings 12 00550 g008
Figure 9. The relative importance of each feature and SHAP summary plot: (a) relative importance; (b) SHAP summary plot.
Figure 9. The relative importance of each feature and SHAP summary plot: (a) relative importance; (b) SHAP summary plot.
Buildings 12 00550 g009
Table 1. Statistical measures of the variables.
Table 1. Statistical measures of the variables.
Statisticsft (MPa)da (mm) ρfc (MPa) a/dbw (mm) d (mm) FstVu (kN)
Median1100.010.00.0340.73.4150.0251.00.55108.0
Mean1269.810.50.0348.73.4157.9282.00.61153.2
Minimum260.00.40.0049.82.555.085.30.1113.0
Maximum4913.022.00.06154.06.0610.01118.03.821481.0
Range4653.0216.00.05144.23.5555.01032.83.711468.0
Standard deviation470.35.10.01 25.80.668.7178.00.40168.9
Table 2. Hyperparameters for the GBRT model.
Table 2. Hyperparameters for the GBRT model.
Hyperparametern EstimatorsLearning RateMax DepthSubsample
Values15000.0180.2
Table 3. Cross-validation measurement results.
Table 3. Cross-validation measurement results.
StatisticsFolds
12345SDAverageCOV%
R20.95940.98530.96440.97900.96920.01090.96921.1246
Table 4. Testing results of four repeated experiments.
Table 4. Testing results of four repeated experiments.
ExperimentRMSE (kN)MAEMAPER2
Case 129.56116.4440.13690.943
Case 219.2769.4100.0650.977
Case 315.448.3130.0570.990
Case 425.4912.560.0670.978
Table 5. Statistical measures of the proposed equations.
Table 5. Statistical measures of the proposed equations.
Model μ M STDCOVMinMax
Sarveghadi et al. [26] 0.991 0.26 27 % 0.22 1.92
Greenough and Nehdi [21] 1.20 0.37 30 % 0.31 3.11
Khuntia et al. [1] 1.48 0.45 31 % 0.18 4.03
Sharma [51] 1.11 0.33 30 % 0.18 2.28
Sabetifar and Nematzadeh [53] 0.968 0.22 22 % 0.33 1.83
Ashour et al. [52] 1.15 0.40 35 % 0.24 3.14
Ashour et al. [52] 1.35 0.35 26 % 0.47 3.22
Proposed GBRT0.9960.0812%0.641.26
Table 6. Previous equations for the shear capacity of SFRC beams.
Table 6. Previous equations for the shear capacity of SFRC beams.
ReferenceAuthorEquation
[26]Sarveghadi et al. V u = [ ρ + ρ v b + 1 a d ( ρ f t ( ρ + 2 ) ( f t a d 3 v b ) a d   + f t ) + v b ] b w d f t = 0.79 f c v b = 0.41 τ F   with   τ = 4.15 MPa
[21]Greenough and Nehdi V u = [ 0.35 ( 1 + 400 d ) ( f c ) 0.18 ( ( 1 + F ) ρ d a ) 0.4 + 0.9 η o τ F ] b w d
[1]Khuntia et al. V u = [ ( 0.167 + 0.25 F ) f c   ] b w d
[51]Sharma V u = ( 2 3 × 0.8 f c ( d a ) 0.25 ) b w d
[52]Ashour et al. V u = [ ( 0.7 f c   + 7 F ) d a + 17.2 ρ d a ] b w d V u = [ ( 2.11 f c   3 + 7 F ) ( ρ d a ) 0.333 ] b w d   for   a d 2.5 V u = [ ( ( 2.11 f c   3 + 7 F ) ( ρ d a ) 0.333 ) 2.5 π d + v b ( 2.5 a d ) ] b w d   for   a d < 2.5
[53]Sabetifar and Nematzadeh V u = [ F + 2 ρ + ρ f c ( F + 3.58 ) 2 ( a / d ) ρ 2 ( f c + 8.52 ) + F ( F 0.73 ) ( ρ ρ a / d ) ] b w d
where: f c and f t are the compressive and tensile strengths of concrete, respectively; F is the fiber factor; η o is the fiber orientation factor, τ is the average fiber–matrix interfacial bond stress.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shatnawi, A.; Alkassar, H.M.; Al-Abdaly, N.M.; Al-Hamdany, E.A.; Bernardo, L.F.A.; Imran, H. Shear Strength Prediction of Slender Steel Fiber Reinforced Concrete Beams Using a Gradient Boosting Regression Tree Method. Buildings 2022, 12, 550. https://doi.org/10.3390/buildings12050550

AMA Style

Shatnawi A, Alkassar HM, Al-Abdaly NM, Al-Hamdany EA, Bernardo LFA, Imran H. Shear Strength Prediction of Slender Steel Fiber Reinforced Concrete Beams Using a Gradient Boosting Regression Tree Method. Buildings. 2022; 12(5):550. https://doi.org/10.3390/buildings12050550

Chicago/Turabian Style

Shatnawi, Amjed, Hana Mahmood Alkassar, Nadia Moneem Al-Abdaly, Emadaldeen A. Al-Hamdany, Luís Filipe Almeida Bernardo, and Hamza Imran. 2022. "Shear Strength Prediction of Slender Steel Fiber Reinforced Concrete Beams Using a Gradient Boosting Regression Tree Method" Buildings 12, no. 5: 550. https://doi.org/10.3390/buildings12050550

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