You are on page 1of 17

Case Studies in Construction Materials 20 (2024) e02800

Contents lists available at ScienceDirect

Case Studies in Construction Materials


journal homepage: www.elsevier.com/locate/cscm

Case study

Prediction of California bearing ratio and modified proctor


parameters using deep neural networks and multiple linear
regression: A case study of granular soils
Rodrigo Polo-Mendoza a, Jose Duque a, b, *, David Mašín a
a
Faculty of Science, Charles University, Prague, Czech Republic
b
Department of Civil & Environmental Engineering, Universidad de la Costa, Barranquilla, Colombia

A R T I C L E I N F O A B S T R A C T

Keywords: The California Bearing Ratio (CBR) and modified proctor parameters belong to the soil
California bearing ratio geotechnical properties used to assess soil behavior. Direct measurement of these properties can
Granular soils be quite time-consuming in large-scale applications or when immediate results are required.
Modified proctor test
Therefore, significant research efforts have been made in the literature to develop indirect
Neural networks
methods for their estimation. However, some gaps in the state-of-the-art can be highlighted in
these topics, such as the deficiency in computational models to calculate the maximum dry unit
weight (γd(max) ), optimum moisture content (wopt ) and CBR, and the lack of methods that consider
their intrinsic influence on each other. Hence, in this investigation, mathematical and compu­
tational models were created to obtain the above-mentioned variables from the soil grain size
distribution. The mathematical model was based on Multiple Linear Regression (MLR) correla­
tions. Meanwhile, the computational model was constructed from a custom-made Deep Neural
Networks (DNNs) architecture. Subsequently, the accuracy of these models was validated with an
experimental case study. The results demonstrated that the proposed methods in this study are
more precise than previous approaches in the literature. Accordingly, the main contribution of
this manuscript to the industry is the formation of models with high exactness to predict the
γd(max) , wopt and CBR of granular soils.

1. Introduction

The proper characterization of the soil geotechnical properties is of great importance for the design and construction of road
infrastructure [1–4]. Even today, the California Bearing Ratio (CBR) and the modified proctor parameters (maximum dry unit weight
-γd(max) - and optimum moisture content -wopt -) are among the most used properties for estimating the quality of subgrades, embank­
ments, and pavement structures. On the one hand, CBR is widely used to estimate subgrades’ bearing capacity under different satu­
ration and compaction degree conditions [5–7]. On the other hand, the modified proctor parameters are typically employed to
evaluate the mechanical performance of soils in terms of density and moisture content [8–10]. In this way, these soil variables are
essential for adequately designing transportation infrastructures. The preceding is particularly meaningful because the road infra­
structure is crucial for the economic growth of the communities [11–13].

* Corresponding author at: Faculty of Science, Charles University, Prague, Czech Republic.
E-mail address: jduque@cuc.edu.co (J. Duque).

https://doi.org/10.1016/j.cscm.2023.e02800
Received 16 March 2023; Received in revised form 24 November 2023; Accepted 17 December 2023
Available online 19 December 2023
2214-5095/© 2024 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

The primary issue with soil geotechnical properties such as CBR and modified proctor parameters is that the reproducibility (in
terms of large extent) of laboratory tests consumes significant amounts of financial resources and time, thus delaying the infrastructure
design and construction processes. Therefore, in the literature, an extensive repertoire of mathematical models (founded mainly on
Multiple Linear Regression -MLR-) has been developed to estimate these properties based on parameters that are easier to measure,
such as the soil grain size distribution, Atterberg limits, plasticity index and the specific gravity of solids [14–16]. Nonetheless, the
traditional mathematical models are being progressively replaced in state-of-the-art by computational models based on Artificial
Intelligence (AI) techniques [17–19]. The preceding is because this type of computational approach can simultaneously consider the
almost-hidden relationships between large amounts of data, thus providing greater precision in value predictions [20,21]. Notably, in
geotechnical engineering, a segment of AI called deep learning is the one that has had the most remarkable expansion in the industry
[21–24].
Within the range of deep learning techniques, Artificial Neural Networks (ANNs) have been used to predict the CBR of granular soils
[25–27]. However, the aforementioned computational models are limited to only covering the CBR estimation, i.e., other properties
with a high potential for correlations, such as the modified proctor parameters, are left out. The preceding shows a gap in the literature,
which is discussed in greater depth in the next section of this manuscript. In this way, the present investigation aims to design a
computational model based on ANNs to predict the CBR and modified proctor parameters of granular soils. Therefore, the main
contribution of this research to the literature is to create a procedure to estimate precisely-quickly some of the essential soil
geotechnical properties.
The remaining sections of this manuscript are organized as follows. First, Section 2 presents a brief review concerning the
implementation of ANNs to predict soil geotechnical properties. Then, in Section 3 , the characteristics of the materials used in this
study are detailed. Succeeding, Section 4 shows the mathematical and computational models developed. For its part, Section 5 ex­
hibits a comparative assessment of the models’ performance. Meanwhile, in Section 6 , a case study is employed to verify the accuracy
of the proposed models. Finally, Section 7 concludes this investigation’s principal findings and recommendations.

2. Background

AI can be understood as the set of techniques based on computational models that try to imitate the logic, learning capacity and
implicit processes of human intelligence to solve problems in diverse fields of science [21,28]. For its part, machine learning is a branch
of AI that utilizes large amounts of data to train computational models based on statistical methods, mathematical functions, and
optimizing parameters [29–31]. Furthermore, within machine learning, there are various sub-branches such as classification/re­
gression, clustering, deep learning, dimensionality reduction, and support-vector machine [29,32,33]. Likewise, among the machine
learning procedures, deep learning stands out for its high capacity to model and adapt to different problems [34–36]. The preceding is
due to the fact that deep learning uses the behavior of the human brain as inspiration to learn from extensive quantities of data [29,37].
Notably, this sub-branch uses architectures of ANNs to discover correlations between variables, thus finding more advantageous paths
to establish connections between parameters that did not have a direct affinity [38,39]. Some of the main architectures are
Bi-Directional Recurrent Neural Network (BDRNN), Biological Neural Network (BNN), Convolutional Neural Network (CNN), Deep
Neural Network (DNN), Feedforward Neural Network (FNN), Probabilistic Neural Network (PNN), Recurrent Neural Network (RNN),
Self-Delimiting Neural Network (SDNN), Spiking Neural Network (SNN), and Time-Delay Neural Network (TDNN) [34,40–42].
The ANNs have been widely used in different applications within geotechnical engineering, for instance, assessing the soil moisture
distribution [43], estimating the effective thermal conductivity in sands [44], evaluating the spatial soil erosion [45], forecasting the
geo-mechanical properties of stabilized clays [46], modeling the load-settlement behavior of single piles [47], predicting the un­
confined compressive strength of geopolymer stabilized clayey soil [48], among many others. However, relatively few studies based on

Table 1
Review of research on the prediction of geotechnical soil properties with AI.
References Type of soils Number of samples Estimated properties Models based on

Granular 90 CBR, wopt , γd(max) MOGAEPRA


[49]
Coal Gangue 384 CBR, UCS ANNs and RF
[50]
[51] Fine 90 qu ANNs
[25] Granular 61 CBR ANNs
[52] Granular 2320 CBR, Mr DT, KNN, MLP, MLR, NEAT, and RF
[53] Granular and fine 158 CBR ANNs
[54] Marine 82 UCS ANNs
[26] Granular 207 CBR ANNs and MLR
[27] Fine 151 CBR ANNs and GEP
[55] Granular and fine 389 CBR ANNs and GEP
[56] Granular and fine 214 CBR DT
[57] Granular 124 CBR ANNs and MLR

Legend: DT - decision tree; GEP - genetic expression programming; KNN - k-nearest neighbors; MLP - multilayer perceptron; MLR - multiple linear
regression; MOGAEPRA - multi-objective genetic algorithm evolutionary polynomial regression analysis; Mr- resilient modulus; NEAT - neuro­
evolution of augmenting topologies; qu - bearing capacity; RF - random forest; UCS - unconfined compressive strength.

2
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

ANNs have focused on the CBR and modified proctor parameters; some notable examples are described in Table 1. The preceding is
because the principal research efforts to estimate these variables have concentrated on developing mathematical correlations. Table 2
shows some of these empirical functions. Notably, most of the expressions presented in the previous tables employ the results of tests
on soil grain size distribution, Atterberg limits, plasticity index and the specific gravity of solids as input data.

3. Materials

This research used a dataset of 90 granular soil samples to create mathematical and computational models. These data were
extracted from an investigation that precedes this study, i.e., by Duque et al. [15]. The experimental information can be accessed by a
public GitHub repository (created especially for this study) through the following link: https://github.com/rpoloe/DNN. These data
include the grain size diameters (D10, D30, D50, and D60), coefficients of uniformity (Cu, according to Eq. 1), coefficients of curvature
(Cc, according to Eq. 2), γ d(max) , wopt , and CBR. Notably, the grain size distribution, CBR and modified proctor parameters were
determined following the standards ASTM D6913, ASTM D1833 and ASTM D1557, respectively. Remarkably, all the granular soil
samples have varying grain size distributions. Moreover, all the CBR tests were carried out at the maximum dry unit weight.
D60
Cu = (1)
D10

D230
Cc = (2)
D10 *D60
Table 3 and Fig. 1 show the statistical analysis of the 90 experimental datasets. On the one hand, Table 3 presents its minimum,
maximum, average, median, standard deviation, variance, skewness, and kurtosis values for each parameter. Meanwhile, Fig. 1 ex­
hibits the scatter chart of the variables, thus allowing to observe the correlation and variability between them. According to this graph,
a substantial relationship exists between γ d(max) , wopt , and CBR with the grain size distribution of the granular soils. However, these
trends are appreciably not straightforward to follow. Consequently, in the following section of this manuscript, mathematical and
computational models are proposed in order to (accurately) emulate these behaviors.

4. Methods

In this investigation, mathematical and computational models based on MLR and ANNs were proposed to estimate the CBR and
modified proctor parameters of granular soils by employing as input variables the grain size distribution (i.e., D10, D30, D50, D60, Cu,
and Cc). The construction of these models is detailed below.

Table 2
Mathematical models to estimate soil geotechnical properties of granular soils.
Estimated Model Units Statistical References
property indicator

CBR CBR = 11.03 + 6.61D60 CBR in percent, and D60 in mm R2 = 0.93


[15]
CBR = 99.08 − 5.162wopt CBR and wopt in percent R2 = 0.74
[58]

⎨ CBR = 5%, ifD60 ≤ 0.01mm CBR in percent, and D60 in mm R2 = 0.84
[14]
CBR = 28.09(D60 )0.358 , if0.01mm < D60 < 30mm

CBR = 95%, if D60 ≥ 30mm
CBR = 6.508D50 + 1.48Cu + 3.97 CBR in percent, D50 in mm, and Cu R2 = 0.85
[16]
dimensionless
γd(max) γd(max) = 26.75 − 7.1D10 + 3.17ln(D10 ) + 0.53ln(D50 ) γd(max) in kN/m3, D10 and D50 in mm 2
R = 0.70
[15]
γd(max) = 13.778C0.166 γd(max) in kN/m3, and Cu dimensionless R2 = 0.90
u [59]
(
P#200
)0.71 γd(max) in kN/m3, P#200 in percent, γs and PL R2 = 0.76
γd(max) = 4.1 2.31γ0.5
s + 0.27PL
0.73
+ [60]
40 dimensionless
γd(max) = 4.49log(Cu ) + 1.51log(CE) + 10.2 γd(max) in kN/m3, CE in kJ/m3, and Cu R = 0.90
[61]
dimensionless
wopt wopt = 9.92D−500.175 C−c 0.058 wopt in percent, D50 in mm, and Cc dimensionless R2 = 0.72
[15]
wopt = 0.1LL + 0.07PL1.44 + 0.09(P#200) + 2e0.27 wopt , LL, PL, P#200, and e in percent R2 = 0.76
[60]
( )
log wopt = 1.67 − 0.193log(Cu ) − 0.153log(CE) wopt in percent, CE in kJ/m3, and Cu R = 0.84
[61]
dimensionless
wopt = 2.0627Cc + 9.6066 wopt in percent, and Cc dimensionless R2 = 0.81
[62]

Legend: CE - compaction energy; e - void ratio; LL - liquid limit; PL - plastic limit;


P#200 - percentage passing No. 200 US sieve; γs -specific unit weight.

3
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Table 3
Statistical parameters of variables used in this study.
Variables Unit Minimum Maximum Average Median Standard Deviation Variance Skewness Kurtosis

D10 mm 0.07 0.90 0.21 0.17 0.15 0.02 3.12 10.75


D30 mm 0.11 3.00 0.49 0.25 0.51 0.26 2.57 7.88
D50 mm 0.18 10.00 1.24 0.75 1.68 2.83 3.31 12.10
D60 mm 0.20 14.00 1.70 1.13 2.37 5.60 3.27 11.83
Cu - 1.69 77.78 8.93 5.58 13.37 178.88 3.63 13.56
Cc - 0.04 4.09 0.85 0.87 0.56 0.31 2.60 12.72
γd(max) kN/m3 16.29 21.90 19.63 19.45 1.25 1.57 -0.15 -0.31
wopt % 6.20 15.40 10.95 11.20 2.21 4.89 -0.22 -0.70
CBR % 6.00 90.00 22.06 17.50 16.69 278.41 2.45 6.63

Fig. 1. Correlation and variability of the parameters considered in this study.

4.1. MLR

In order to create and evaluate the mathematical models, the R programming language was employed via RStudio. This software is
an integrated development environment that offers interactive capabilities to facilitate statistical analysis of a large number of data
[63–65]. By implementing MLR techniques, the data from the 90 granular soil samples were analyzed. Notably, in order to improve the
accuracy of the correlations, it was necessary to create four new variables, namely A (Eq. 3), B (Eq. 4), C (Eq. 5), and D (Eq. 6). The
above variables were obtained by transforming the original input variables through a trial-and-error fine-tuning procedure. Thus, the
definitive mathematical models are presented in Eq. 7, Eq 8, and Eq. 9 for predicting the CBR, γd(max) , and wopt , respectively.

A = D210 (3)

B = D50 *ln(D30 + Cu + Cc + A) (4)

C = log10 (D50 + D60 ) (5)

D = ln(A) (6)

CBR = − 47.8 + 115.3D10 + 17.3D50 + 0.8Cu + 2.4Cc − 67.3A − 3.4B − 9.1D (7)

4
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

γd(max) = 24.1 − 0.7D60 − 6.1A + 0.2B + 2.2C + 1.2D (8)

wopt = 34.3 − 58.2D10 + 38.9A − 3.9C + 3.8D (9)

Tables 4–6 show the estimated coefficients, standard deviations, confidence intervals (at 95% confidence level), and two statistical
tests (Student’s t and ANOVA tests) for each of the correlations developed in this study. Moreover, these tables offer general infor­
mation concerning the models: the residual standard error, multiple R-squared, adjusted R-squared, F-statistic, and p-value. Notably,
these two statistical tests are of outstanding importance because they validate the model’s significance and its parameters (intercept
and variables) [66–70]. On the one hand, the student’s t-test (also called t-test) is a method to evaluate hypotheses regarding the mean
of samples using the t-value, which is defined as the ratio of the difference between the mean of sample sets and their respective
variations [71–73]. Meanwhile, the ANOVA test determines the F-value to portray the explanatory capability of a group of inde­
pendent variables on the variability of the dependent variable [66,74,75]. Remarkably, all the statistical indicators assessed exhibit
that the three proposed models (and their parameters) have high statistical significance in predicting the CBR, γ d(max) , and wopt .

4.2. ANNs

The main issue to address in the design procedure of ANNs is to define the most accurate and faultless architecture that guarantees a
correct reading and dissecting of the input data to process it later and thus predict the output data [76–78]. Overall, for the vast
majority of ANNs typologies, the design process of the architecture can be summarized in five steps: (i) commence with basic network
architecture, (ii) set the network parameters, (iii) train the ANNs, (iv) test the ANNs, and (v) reconfigure the architecture to improve
exactness [79–81]. Notably, unduly complex network architecture may generate defects such as overfitting and excessive time con­
sumption during execution [82–85]. Hence, high-complexity architecture designs do not ensure sufficient explanatory and predictive
capacities [86,87]. Accordingly, the statistical learning theory claims that the prevalence of learning with low complexity models
should be preferred [88–91]. Therefore, this research aimed to encounter the simplest competent model to adequately predict the
geotechnical soil properties (under consideration).
In order to address the analyzed problem, it was chosen to implement a type of ANN denominated DNN. The DNNs are defined as
ANNs designed under an architecture with more than one hidden layer and densely connected neurons [77,92–94]. The DNNs have
been selected because this architecture guarantees an elevated capacity to find relationships between variables and generate automatic
learning based on data representations, even in problems concerning high dimensions and non-smooth functions [95–98]. Thus,
various network configurations were explored using the canonical architecture of the DNNs as a starting point. Fig. 2 shows the most
efficient design of layers-neurons-connections encountered. This computational model was defined by scrutinizing the simplest
network that would remarkably fit the data. It is essential to highlight that no normalization or statistical transformation was per­
formed on the input data. Hence, it was possible to uncover the flawless balance between precision and complexity. In this regard, it
was avoided developing high-complex models that could cause overfitting or excessive consumption of computational resources [82,
99,100]. This research employed two error measures to recognize the most appropriate DNN model, i.e., the Mean Squared Error
(MSE) and the Mean Squared Logarithmic Error (MSLE). Besides, the MSE was selected as the loss function. Respectively, Eq. 10 and Eq
11 present the mathematical formula that defines MSE and MSLE [78,101].

1∑ n
( )2
MSE = ODi − EDi (10)
n i=1

1∑ n
( ( ) ( ) )2
MSLE = ln 1 + ODi − ln 1 + EDi (11)
n i=1

Where,

Table 4
Parameters and statistical tests associated with the MLR model regarding CBR.
Parameters Coefficients 95% confidence interval Statistical tests

Estimate Std. error Lower limit Upper limit t value Pr (>|t|) F value Pr (>F)
Intercept -47.7786 15.5124 -78.6377 -16.9196 -3.080 0.00282

D10 115.2960 39.3333 37.0494 193.542 2.931 0.00437 87.9248 1.282e-14


D50 17.3138 1.7948 13.7432 20.8843 9.646 3.74e-15 1820.3223 2.2e-16
Cu 0.8349 0.1058 0.6245 1.0453 7.893 1.14e-11 19.0600 3.669e-05
Cc 2.3802 0.8905 0.6087 4.1518 2.673 0.00907 0.0013 0.971362
A -67.2617 26.6142 -120.2057 -14.3176 -2.527 0.01342 1.4246 0.236089
B -3.4276 0.5109 -4.4439 -2.4112 -6.709 2.34e-09 48.4531 7.641e-10
D -9.0779 2.7018 -14.4526 -3.7031 -3.360 0.00119 11.2892 0.001185
Model information
Residual standard error: 3.459 on 82 degrees of freedom, Multiple R-squared: 0.9604, Adjusted R-squared (R2 ): 0.957
F-statistic: 284.1 on 7 and 82 DF, p-value: < 2.2e-16

5
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Table 5
Parameters and statistical tests associated with the MLR model regarding γd(max) .
Parameters Coefficients 95% confidence interval Statistical tests

Estimate Std. error Lower limit Upper limit t value Pr (>|t|) F value Pr (>F)

Intercept 24.14678 0.57133 23.01062 25.28294 42.264 < 2e-16


D60 -0.71155 0.22506 -1.15910 -0.26400 -3.162 0.00218 61.585 1.223e-11
A -6.12832 0.91056 -7.93907 -4.31757 -6.730 1.95e-09 10.371 0.001821
B 0.18072 0.06151 0.05839 0.30305 2.938 0.00426 10.788 0.001491
C 2.21878 0.43211 1.35947 3.07809 5.135 1.80e-06 106.147 < 2.2e-16
D 1.17337 0.14736 0.88034 1.46641 7.963 7.20e-12 63.406 7.200e-12
Model information
Residual standard error: 0.6438 on 84 degrees of freedom, Multiple R-squared: 0.7502, Adjusted R-squared (R2 ): 0.7354
F-statistic: 50.46 on 5 and 84 DF, p-value: < 2.2e-16

Table 6
Parameters and statistical tests associated with the MLR model regarding wopt .
Parameters Coefficients 95% confidence interval Statistical tests

Estimate Std. error Lower limit Upper limit t value Pr (>|t|) F value Pr (>F)

Intercept 34.3375 5.1161 24.1654 44.5096 6.712 2.03e-09


D10 -58.1942 12.4961 -83.0398 -33.3486 -4.657 1.17e-05 49.679 4.460e-10
A 38.8711 8.4413 22.0876 55.6546 4.605 1.43e-05 47.834 8.109e-10
C -3.9381 0.3370 -4.6082 -3.2680 -11.684 < 2e-16 124.522 < 2.2e-16
D 3.7643 0.8976 1.9796 5.5490 4.194 6.70e-05 17.587 6.703e-05
Model information
Residual standard error: 1.158 on 85 degrees of freedom, Multiple R-squared: 0.7382, Adjusted R-squared (R2 ): 0.7258
F-statistic: 59.91 on 4 and 85 DF, p-value: < 2.2e-16

Fig. 2. Proposed DNN to predict the CBR and modified proctor parameters of granular soils.

n - number of data points, i - data index, OD - observed data, ED - estimated data.


The proposed DNN architecture is composed of five layers. The first layer is the input layer, which has a neuron for each input
variable, i.e., 6 neurons. The second layer is a hidden layer composed of 3 neurons with Rectified Linear Unit (ReLU) as an activation
function. Similarly, the third layer is another hidden layer formed of 8 neurons with Sigmoid as an activation function. Subsequently
follows a hidden layer with only 2 neurons. Finally, the last layer is the output layer, which has a neuron for each output variable, i.e., 3
neurons. Notably, all neurons are densely connected due to the typology of ANN selected; thus, all possible neural connections were
established. It is essential to highlight that the datasets were distributed as follows for the learning process: 80% for training and 20%
for validation. Fig. 3 displays the statistical evaluations of the computational model suggested in Fig. 2. The statistical assessment
indicates that this DNN is exceptionally fitting because, in less than 2000 epochs (training and validation cycles), it reduces the errors
to a negligible magnitude. Furthermore, in this graph, it is possible to appreciate the stability of the DNN.
The precision reached by the proposed computational model is associated with using two activation functions, i.e., ReLU and

6
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Sigmoid. The activation functions are mathematical operations done to the data after leaving one neuron, i.e., before reaching the next
one in the subsequent layer [102–104]. Selecting an appropriate activation function for a respective layer can substantially improve
the performance of an ANN [105–108]. Such is the importance of these functions that they can decisively influence the stability of the
model and its predictions [109–111]. On the one hand, the ReLU function enhances the convergence of DNNs and minimizes the
phenomenon of vanishing gradients [112–114]. Meanwhile, the Sigmoid is a continuously differentiable function that allows con­
trolling the gradient calculation and output data at high exactness due to its smoothness [115–118].
Algorithm 1 exhibits the simplified pseudo-code of the proposed computational model. This model uses RMSprop as the optimizer
algorithm. RMSprop is a widely used optimizer to train ANNs outside stochastic gradient descent methods [119–121]. As is shown in
Algorithm 2, RMSprop strives to keep a moving average of the square of gradients, divide the gradient by the root of this average, and
apply plain momentum [122–124]. In this way, RMSprop encounters the maximum likelihood [120,123,125], i.e., θ = µ (according to
the nomenclature of Algorithm 2). Notably, the DNN was created in Python (a general-purpose programming language) utilizing
Google Colab as the development environment. The authors uploaded this DNN model to the following GitHub repository: https://
github.com/rpoloe/DNN. Likewise, the authors guarantee free and prompt access to the research community by furnishing an open
license for use, modification, and distribution.
Algorithm 1. : Simplified pseudo-code to create the proposed DNN.

Algorithm 2. : Simplified pseudo-code to employ the RMSprop optimizer in ANNs. Adapted from: [120,126].

Legend: t - iteration index; T - maximum number of iterations; θ - parameter to optimize; µ - current parameter vector; gt -
subgradient of ft at θt; ▽θ - parameter changes; ft - arbitrary convex function;
s - scaling vector; βt and α - step-sizes; δ - small constant.

5. Models’ performance

In order to assess the predictive capacity of the proposed models, the errors associated with the estimations (obtained from the DNN
model and MLR models) were computed against the data measured in the laboratory. Table 7 shows these calculations, as well as the
R2 values. According to this table, the DNN architecture is significantly more precise than mathematical models, even reducing errors
to almost zero. The preceding can be visually verified in Fig. 4. This graph depicts the trends of the 1:1 lines of the predicted values
versus the measured values for the γd(max) , wopt , and CBR.
The statistical analyzes in Tables 4–7 are enough to understand the performance of the MLR models. Conversely, for the DNN
model, it is still necessary to carry out a deeper inspection. For this purpose, several SHapely Additive exPlanations (SHAP) assessments
were conducted. Thus, Fig. 5 shows five SHAP evaluations: a SHAP global interpretation using a summary plot (see Fig. 5a) and four
SHAP local interpretations, namely a waterfall plot (see Fig. 5b), a decision plot for the training datasets (see Fig. 5c), a decision plot for
the validation datasets (see Fig. 5d), and a force plot (see Fig. 5e). These graphs are explained below.
Fig. 5a presents the summary plot of the SHAP values for the proposed DNN model. The higher the SHAP values related to a specific
input variable, the greater the influence of that input variable on the output variables [127,128]. Accordingly, for the forecast of the
γd(max) , wopt , and CBR, the parameters with more predictive significance (ordered from largest to smallest) are the D60, D50, Cu, D30, Cc,
and finally, the D10. It is essential to highlight that the only parameter with a minimal contribution to the total convergence of the

7
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 3. Statistical evaluation of the proposed DNN.

model is D10 (i.e., its SHAP value is close to 0). Similarly, Fig. 5b shows the waterfall plot of the DNN model. In this graph, the input
variables marked with red color exhibit a positive contribution to the output variables, whilst the blue color denotes a negative
contribution [129]. Likewise, as the amplitude of the f(x) values increases for an input variable, the model’s sensitivity for that
parameter also augments [129]. Therefore, three main conclusions can be drawn from the waterfall plot: (i) D30 and Cu are the input
parameters that generate the most substantial sensitivity in the quantitative responses of the DNN model; (ii) Cu, D50, Cc, and D10 cause
a positive contribution to the estimated valued; and (iii) contrariwise, D30 and D60 provoke a negative contribution.
Respectively, Fig. 5c and Fig. 5d show the average decision plot for the training and validation datasets. This graph displays the
trend followed by the model’s output values by "traversing" the input parameters [130]. In other words, the decision plot charts the
average amplitude of each change generated in the input variables throughout the computational architecture to reach the output
variables [130]. Since Fig. 5c and Fig. 5d pattern the same shape and behavior, it can be concluded that the validation data has a
similar response to the exhibited by the training data. The preceding is particularly meaningful because it reveals that the overfitting
and underfitting phenomena have not occurred. Finally, Fig. 5e portrays the force plot of the proposed DNN model. The force plot
follows the same notation as the waterfall plot, i.e., red color for input variables with positive contribution and blue color for negative
contribution. However, in this graph, the f(x) values indicated the fitting of the output values concerning the base values [128,131].
Hence, a higher f(x) value for a specific input variable alludes to more "identicality" (in classification problems) or "closeness" (in
regression problems) to the expected outputs [128,131]. In light of the above, D30, Cu and D50 are the parameters with the highest
likeness regarding the γ d(max) , wopt , and CBR.

6. Case study

In order to verify the efficiency of the proposed models, a final testing procedure is performed. For this purpose, an experimental
case study was conducted. Consequently, a granular soil sample extracted in the Northern Region of Colombia (specifically from the
city of Barranquilla) was analyzed. Fig. 6a shows the results of the grain size distribution test. From this granulometry, the grain size
diameters are 0.171, 0.462, 1.081, and 1.445 mm for the D10, D30, D50, and D60, respectively. Subsequently, the Cu and Cc take values
of 8.450 and 0.864. Besides, Fig. 6b presents the laboratory measurements of γ d(max) , wopt , and CBR.
The mathematical and computational models previously proposed (in Eqs. 7–9 and Fig. 2, respectively) were implemented to
estimate the soil geotechnical parameters of the granular soil sample from Fig. 6. Furthermore, with the aim of comparing these models
with other expressions in the literature, these calculations were also carried out employing some of the equations presented in Table 2.
Specifically, the correlations of CBR proposed by Duque et al. [15] and Rehman et al. [16], γd(max) proposed by [15] and Gurtug et al.
[59], and wopt proposed by [15] and Hassan et al. [62] were addressed. Fig. 7 shows the predicted values from the different methods. In
addition to presenting the estimations, this graph reports the percentage error concerning the average experimental values (see Fig. 6).

Table 7
Models’ performance in terms of goodness-of-fit parameters.
Metrics γd(max) wopt CBR

DNN MLR MLR/DNN DNN MLR MLR/DNN DNN MLR MLR/DNN

MSE 0.03326 0.41619 12.51 0.00381 1.29188 339.08 0.01215 11.02047 907.03
MSLE 0.00008 0.00100 12.50 0.00003 0.00915 305.00 0.00002 0.03742 1871.00
R2 0.9804 0.7354 - 0.9992 0.7258 - 0.9999 0.9570 -

8
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 4. Precision of the models to estimate the soil geotechnical properties.

It is essential to highlight that the ANNs present stochastic components in their internal functioning. Therefore, the proposed DNN
model yields slightly different outcomes on each execution. Hence, for this testing procedure, the average of 100 independent exe­
cutions was taken as the DNN model’s predicted value. Thus, from Fig. 7, it is possible to obtain the following findings:

• The most accurate predicted method for the modified proctor parameters was the proposed DNN model. Subsequently, the pro­
posed MLR models ranked second in exactness. Meanwhile, the correlations from the literature (i.e., [15,59,62]) exhibited the most
elevated errors.
• For the CBR, the lower percentage error was associated with the proposed DNN model, closely followed by the correlation from
Rehman et al. [16]. Thus, the third place (concerning the highest precision) was occupied by the proposed MLR model. Finally, the
worst performance was exhibited in the mathematical model by Duque et al. [15].
• Notably, the computational model based on DNNs is noticeably more accurate than its mathematical counterparts. However, it
must be noted that, as the validation was performed using one sample only, the results must only be considered indicative regarding
predictive capability. Therefore, testing employing more samples is needed for more general conclusions.

6.1. Partial dependence analysis

In order to examine the importance of each input variable within the general performance of the proposed DNN model, a partial
dependence analysis was conducted. In this way, the Sensitivity Ratio (SR) was employed as performance criteria. Eq. 12 shows the
mathematical expression that defines the SR, which was adapted from [132,133]. According to this equation, the SR is calculated for
each parameter (i.e., input variable). Notably, more elevated SR values imply a more influence on the model’s outcomes [133,134].
|Initial Prediction− New Prediction|
SR = |Initial
Initial Prediction
Parameter− New Parameter|
(12)
Initial Parameter

For the computing of the SRs, the "initial parameter" values were taken from the case study (i.e., 0.171 mm, 0.462 mm, 1.081 mm,
1.445 mm, 8.450 and 0.864 for D10, D30, D50, D60, Cu, and Cc, respectively). Likewise, the "initial prediction" values are the ones
reported in Fig. 7(i.e., 20.22 kN/m3, 9.96% and 22.51% for γd(max) , wopt and CBR, respectively). Meanwhile, the "new parameter" values

9
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 5. SHAP assessments for the proposed DNN model.

correspond to a 1D arrays that goes from 0 to 1 (with steps of 0.01) for the grain size diameters and from 0 to 15 for the grain size
coefficients (equally, with steps of 0.01). Clearly, the "new prediction" values are obtained from the outcomes of the proposed DNN
model. The results from this analysis are presented in Fig. 8 as a partial dependence plot. From this graph, the subsequent deductions
can be drawn:

• Variations in D50 and D60 have a low impact on the γ d(max) , wopt and CBR estimations. However, the influence of these variables is
more noticeable for greater diameter values, i.e., from 0.8 mm.
• D30 is the grain size diameter with the most remarkable impact on the model’s outcomes. Notably, the peak of influence is located in
the diameters between 0.32-0.6 mm.
• D10 is the second most crucial grain size diameter for the three output variables. Also, this variable presents a peak of importance in
the diameters between 0.12-0.22 mm.
• Cu has more influence on the DNN model’s outcomes than Cc. On the one hand, the influence peak of Cu is located between 5.76 and
11.08. On the other hand, this range is between 0.6 and 1.13 for Cc.
• Overall, the SR values for the grain size coefficients (i.e., Cu and Cc) are considerably higher than those for the grain size diameters
(i.e., D10, D30, D50 and D60). Therefore, it is evident that the proposed DNN model is highly sensitive to Cu and Cc (even more than
concerning the grain size diameters). The preceding behavior agrees with the findings of different statistical investigations in the
literature [15,59,61,62].
• It is crucial to highlight that the granulometric parameters (at least for soils) depend on complex physicochemical processes
(including the dynamic arrangement of the soil particles) [135–137]. Nonetheless, the previous findings come from a computa­
tional standpoint of a case study and do not consider these aspects. Therefore, the conclusions obtained from this partial depen­
dence plot should not be generalized.

7. Conclusions

7.1. Summary

This study employs the results from a previous experimental investigation with the aim of developing mathematical and
computational models to predict the CBR and the modified proctor parameters of granular soils. In this regard, the main goal of this

10
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 6. Results of experimental tests for the granular soil sample.

research is to propose models to forecast the bearing capacity of coarse-grained soils from their grain size distribution. On the one
hand, the mathematical models were based on MLR and were evaluated with a statistical approach (i.e., F-test and t-test). On the other
hand, the computational model was a DNN, which was inspected with performance analysis and SHAP assessments. Subsequently, the
mathematical and computational models were compared to each other in terms of predictive capabilities (i.e., MSE, MSLE, and R2).
Finally, experimental verification of the models was carried out.

7.2. Main findings

From the investigative effort performed, the following conclusions can be drawn:

• For granular soils, there is a substantial dependency between the grain size distribution parameters and their γd(max) , wopt , and CBR.
• The computational model that best predicted the CBR and modified proctor parameters from the grain size distribution was a DNN
composed of five (5) layers as follows: 6-neuron input layer (one neuron for each input variable), 3-neuron hidden layer with ReLU
as activation function, 8-neuron hidden layer with Sigmoid as activation function, 2-neuron hidden layer, and finally a 3-neuron
output layer (one neuron for each output variable).
• The mathematical models proposed to estimate the γ d(max) , wopt , and CBR have R2 values of 0.735, 0.726 and 0.957, respectively.
Therefore, these correlations have a remarkable explanatory capacity for the phenomenon under study. Meanwhile, the proposed
DNN presents R2 values between 0.9804-0.9999. Consequently, it is clear that the computational model is significantly more
precise than the mathematical counterparts.
• The accuracy of the proposed models was evaluated in terms of error metrics, i.e., MSE and MSLE. The correlation for the
γd(max) generates around 12.5 times more errors than the computational model. The preceding ratio is more elevated for the wopt
and CBR, which are between 305-339 and 907-1871, respectively. Hence, it was possible to develop a high-precision DNN ar­
chitecture to estimate all the parameters considered in this study.
• The statistical tests and the SHAP assessments demonstrated that all the input variables (i.e., D10, D30, D50, D60, Cu, and Cc) are
statistically significant for the MLR and DNN models, respectively.

11
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 7. Comparison of the different models to predict the soil geotechnical parameters.

• The SHAP values demonstrate that D60 and D50 are the variables with the most predictive importance on the internal functioning of
the proposed DNN model. Meanwhile, the partial dependence plot reveals that these input variables have a minimum effect on the
DNN model’s outcomes variability. Therefore, it is possible to deduce that D60 and D50 are the parameters responsible for sustaining
the forecasting stability of the proposed computational model.
• Regarding the experimental verification of the predictive capabilities of the models, it was determined that: (i) for the γd(max) , the
more significant errors were associated with the models by Gurtug et al. [59] and Duque et al. [15], whilst the proposed MLR and
DNN models generated the lowest errors; (ii) likewise, for the wopt , the models with deficient exactness were the ones developed by
Hassan et al. [62] and Duque et al. [15]; and (iii) concerning the CBR, the maximum discrepancy in the predicted value was caused
by the model developed by Duque et al. [15], followed by the proposed MLR model, then the one created by Rehman et al. [16], and
finally, the more accurate model was the DNN. Overall, the models proposed in this investigation outperformed their counterparts
from previous research in exactness. Notably, the computational model (i.e., DNN) is more accurate than the mathematical model
(i.e., correlations formed from MLR).

7.3. Future research lines

Based on this study’s scope and limitations, several recommendations were proposed to develop future research lines. These are
listed below: (i) expand the datasets used so that the models developed can cover fine-grained soils; (ii) carry out an extensive
experimental study to measure the effectiveness of the proposed models, thus achieving correction parameters for certain types of soils;

12
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

Fig. 8. Partial dependence plot based on SR values.

13
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

(iii) evaluate the advantages of other types of ANN architectures, such as the BDRNN, BNN, FNN, PNN, RNN, SDNN, SNN, and TDNN;
and (iv) consider other machine learning techniques (or even other branches of AI) to predict soil geotechnical properties.

CRediT authorship contribution statement

Duque Jose: Conceptualization, Investigation, Supervision, Validation, Writing – original draft, Writing – review & editing. Mašín
David: Conceptualization, Funding acquisition, Project administration, Resources, Validation, Writing – review & editing. Polo-
Mendoza Rodrigo: Data curation, Formal analysis, Investigation, Methodology, Software, Visualization, Writing – original draft.

Declaration of Competing Interest

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

Data availability

The authors publicly share the computational model produced in this research through the following GitHub repository: https://
github.com/rpoloe/DNN.

Acknowledgments

The authors appreciate the financial support given by the Czech Science Foundation grant No. 21–35764J. The first and the last
authors acknowledge the institutional support by the Center for Geosphere Dynamics (UNCE/SCI/006).

References

[1] Z. Zhou, R. Bai, M. Shen, Q. Wang, The effect of overconsolidation on monotonic and cyclic behaviours of frozen subgrade soil, Transp. Geotech. vol. 32
(100710) (2022) 1–17, https://doi.org/10.1016/j.trgeo.2021.100710.
[2] W. Fuentes, J. Duque, C. Lascarro, M. Gil, Study of the bearing capacity of closely spaced square foundations on granular soils, Geotech. Geol. Eng. vol. 37 (3)
(2019) 1401–1410, https://doi.org/10.1007/s10706-018-0694-5.
[3] R. Polo-Mendoza, J. Duque, D. Mašín, E. Turbay, C. Acosta, Implementation of deep neural networks and statistical methods to predict the resilient modulus of
soils, Int. J. Pavement Eng. vol. 24 (1) (2023), 2257852, https://doi.org/10.1080/10298436.2023.2257852.
[4] D. Mašín, Chapter 7: advanced modelling approaches (W. Wu, Ed.). Modelling of Soil Behaviour with Hypoplasticity: Another Approach to Soil Constitutive
Modelling, Springer International Publishing, 2018, pp. 119–189, https://doi.org/10.1007/978-3-030-03976-9_7 (W. Wu, Ed.).
[5] S. Mukherjee, P. Ghosh, Soil behavior and characterization: effect of improvement in CBR characteristics of soil subgrade on design of bituminous pavements,
Indian Geotech. J. vol. 51 (3) (2021) 567–582, https://doi.org/10.1007/s40098-021-00533-8.
[6] Y.K. Raju, C.V. Kumar, Experimental investigation on design of thickness for flexible pavement subgrade soils using CBR approach, E3S Web Conf. vol. 184
(01087) (2020) 1–3, https://doi.org/10.1051/e3sconf/202018401087.
[7] H. Haghighi, A. Arulrajah, A. Mohammadinia, S. Horpibulsuk, A new approach for determining resilient moduli of marginal pavement base materials using the
staged repeated load CBR test method, Road. Mater. Pavement Des. vol. 19 (8) (2018) 1848–1867, https://doi.org/10.1080/14680629.2017.1352532.
[8] M. Izquierdo, X. Querol, E. Vazquez, Procedural uncertainties of proctor compaction tests applied on MSWI bottom ash, J. Hazard. Mater. vol. 186 (2–3)
(2011) 1639–1644, https://doi.org/10.1016/j.jhazmat.2010.12.045.
[9] G. Cerni, S. Camilli, Comparative analysis of gyratory and proctor compaction processes of unbound granular materials, Road. Mater. Pavement Des. vol. 12
(2) (2011) 397–421, https://doi.org/10.1080/14680629.2011.9695251.
[10] S. Cápayová, Z. Štefunková, S. Unčík, A. Zuzulová, Requirements for pavement base layers with unbound granular material, Slovak J. Civ. Eng. vol. 27 (3)
(2019) 21–28, https://doi.org/10.2478/sjce-2019-0018.
[11] A.O. Acheampong, J. Dzator, M. Dzator, R. Salim, Unveiling the effect of transport infrastructure and technological innovation on economic growth, energy
consumption and CO2 emissions, Technol. Forecast. Soc. Chang. vol. 182 (121843) (2022) 1–24, https://doi.org/10.1016/j.techfore.2022.121843.
[12] Y. Zhang, L. Cheng, The role of transport infrastructure in economic growth: empirical evidence in the UK, Transp. Policy vol. 133 (2023) 223–233, https://
doi.org/10.1016/j.tranpol.2023.01.017.
[13] M. Munday, L. Reynolds, A. Roberts, Re-appraising ‘in-process’ benefits of strategic infrastructure improvements: capturing the unexpected socio-economic
impacts for lagging regions, Transp. Policy vol. 134 (2023) 119–127, https://doi.org/10.1016/j.tranpol.2023.02.012.
[14] NCHRP, Guide for mechanistic-empirical design of new and rehabilitated pavement structures - appendix CC-1: correlation of CBR values with soil index
properties, Natl. Coop. Highw. Res. Prog. (2001) 4–8.
[15] J. Duque, W. Fuentes, S. Rey, E. Molina, Effect of grain size distribution on California bearing ratio (CBR) and modified proctor parameters for granular
materials, Arab. J. Sci. Eng. vol. 45 (10) (2020) 8231–8239, https://doi.org/10.1007/s13369-020-04673-6.
[16] A. ul Rehman, K. Farooq, H. Mujtaba, Prediction of California Bearing Ratio (CBR) and compaction characteristics of granular soils, Acta Geotech. Slov. vol. 14
(1) (2017) 63–72.
[17] S.C. Jong, D.E.L. Ong, E. Oh, State-of-the-art review of geotechnical-driven artificial intelligence techniques in underground soil-structure interaction, Tunn.
Undergr. Sp. Technol. Inc. Trench Technol. Res. vol. 113 (103946) (2021) 1–23, https://doi.org/10.1016/j.tust.2021.103946.
[18] S. Sharma, S. Ahmed, M. Naseem, W.S. Alnumay, S. Singh, G.H. Cho, A survey on applications of artificial intelligence for pre-parametric project cost and soil
shear-strength estimation in construction and geotechnical engineering, Sensors vol. 21 (463) (2021) 1–44, https://doi.org/10.3390/s21020463.
[19] A. Baghbani, T. Choudhury, S. Costa, J. Reiner, Application of artificial intelligence in geotechnical engineering: a state-of-the-art review, Earth-Sci. Rev. vol.
228 (103991) (2022) 1–26, https://doi.org/10.1016/j.earscirev.2022.103991.
[20] Z. Yin, Y. Jin, Z. Liu, Practice of artificial intelligence in geotechnical engineering, J. Zhejiang Univ. Sci. A Appl. Phys. Eng. vol. 21 (6) (2020) 407–411,
https://doi.org/10.1631/jzus.A20AIGE1.
[21] M. Jaksa, Z. Liu, Applications of artificial intelligence and machine learning in geotechnical engineering, Geosciences vol. 11 (399) (2021) 1–2, https://doi.
org/10.3390/geosciences11100399.
[22] H. Moayedi, M. Mosallanezhad, A.S.A. Rashid, W.A.W. Jusoh, M.A. Muazu, A systematic review and meta-analysis of artificial neural network application in
geotechnical engineering: theory and applications, Neural Comput. Appl. vol. 32 (2) (2020) 495–518, https://doi.org/10.1007/s00521-019-04109-9.

14
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

[23] W. Zhang, H. Li, Y. Li, H. Liu, Y. Chen, X. Ding, Application of deep learning algorithms in geotechnical engineering: a short critical review, Artif. Intell. Rev.
vol. 54 (8) (2021) 5633–5673, https://doi.org/10.1007/s10462-021-09967-1.
[24] Z.Z. Wang, S.H. Goh, A maximum entropy method using fractional moments and deep learning for geotechnical reliability analysis, Acta Geotech. vol. 17 (4)
(2022) 1147–1166, https://doi.org/10.1007/s11440-021-01326-2.
[25] Y. Erzin, D. Turkoz, Use of neural networks for the prediction of the CBR value of some Aegean sands, Neural Comput. Appl. vol. 27 (5) (2016) 1415–1426,
https://doi.org/10.1007/s00521-015-1943-7.
[26] S. Taha, A. Gabr, S. El-Badawy, Regression and neural network models for California bearing ratio prediction of typical granular materials in Egypt, Arab. J.
Sci. Eng. vol. 44 (10) (2019) 8691–8705, https://doi.org/10.1007/s13369-019-03803-z.
[27] T. Taskiran, Prediction of California bearing ratio (CBR) of fine grained soils by AI methods, Adv. Eng. Softw. vol. 41 (6) (2010) 886–892, https://doi.org/
10.1016/j.advengsoft.2010.01.003.
[28] D. Minh, H.X. Wang, Y.F. Li, T.N. Nguyen, Explainable artificial intelligence: a comprehensive review, Artif. Intell. Rev. vol. 55 (5) (2022) 3503–3568, https://
doi.org/10.1007/s10462-021-10088-y.
[29] E.R. Okoroafor, C.M. Smith, K.I. Ochie, C.J. Nwosu, H. Gudmundsdottir, M. (Jabs) Aljubran, Machine learning in subsurface geothermal energy: two decades in
review, Geothermics vol. 102 (102401) (2022) 1–16, https://doi.org/10.1016/j.geothermics.2022.102401.
[30] L. Li, Y. Cai, Q. Zhou, A survey on machine learning-based routing for VLSI physical design, Integr. VLSI J. vol. 86 (2022) 51–56, https://doi.org/10.1016/j.
vlsi.2022.05.003.
[31] S. Ledesma, M.-A. Ibarra-Manzano, D.-L. Almanza-Ojeda, J.G. Avina-cervantes, E. Cabal-yepez, On removing conflicts for machine learning, Expert Syst. Appl.
vol. 206 (117835) (2022) 1–8, https://doi.org/10.1016/j.eswa.2022.117835.
[32] N. Pudjihartono, T. Fadason, A.W. Kempa-Liehr, J.M. O´Sullivan, A review of feature selection methods for machine learning-based disease risk prediction,
Front. Bioinforma. vol. 2 (927312) (2022) 1–17, https://doi.org/10.3389/fbinf.2022.927312.
[33] K.A. AlAfandy, H. Omara, M. Lazaar, M. Al Achhab, Chapter 5: machine learning. Computational Intelligence and Applications for Pandemics and Healthcare,
IGI Global, 2022, pp. 83–113, https://doi.org/10.4018/978-1-7998-9831-3.ch005.
[34] A.A. Kashyap, S. Raviraj, A. Devarakonda, S.R. Nayak K, K.V. Santhosh, S.J. Bhat, Traffic flow prediction models – a review of deep learning techniques,
Cogent Eng. vol. 9 (1) (2022), 2010510, https://doi.org/10.1080/23311916.2021.2010510.
[35] S. Zhang, L. Yao, A. Sun, Y. Tay, Deep learning based recommender system: a survey and new perspectives, ACM Comput. Surv. vol. 52 (1) (2019) 1–38,
https://doi.org/10.1145/3285029.
[36] O.B. Sezer, M.U. Gudelek, A.M. Ozbayoglu, Financial time series forecasting with deep learning: a systematic literature review: 2005–2019, Appl. Soft Comput.
J. vol. 90 (106181) (2020) 1–32, https://doi.org/10.1016/j.asoc.2020.106181.
[37] S. Pouyanfar, et al., A survey on deep learning: algorithms, techniques, and applications, 92, ACM Comput. Surv. vol. 51 (5) (2018) 1–36, https://doi.org/
10.1145/3234150.
[38] B. Qiu, et al., Automatic segmentation of mandible from conventional methods to deep learning - a review, J. Pers. Med. vol. 11 (629) (2021) 1–26, https://
doi.org/10.3390/jpm11070629.
[39] Y. LeCun, Y. Bengio, G. Hinton, Deep learning, Nature vol. 521 (7553) (2015) 436–444, https://doi.org/10.1038/nature14539.
[40] S. Gaba, I. Budhiraja, V. Kumar, S. Garg, G. Kaddoum, M.M. Hassan, A federated calibration scheme for convolutional neural networks: models, applications
and challenges, Comput. Commun. vol. 192 (2022) 144–162, https://doi.org/10.1016/j.comcom.2022.05.035.
[41] J. Schmidhuber, Deep learning in neural networks: an overview, Neural Netw. vol. 61 (2015) 85–117, https://doi.org/10.1016/j.neunet.2014.09.003.
[42] Y. Yang, T. Wang, J.P. Woolard, W. Xiang, Guaranteed approximation error estimation of neural networks and model modification, Neural Netw. vol. 151
(2022) 61–69, https://doi.org/10.1016/j.neunet.2022.03.023.
[43] M. Dursun, S. Özden, An efficient improved photovoltaic irrigation system with artificial neural network based modeling of soil moisture distribution - a case
study in Turkey, Comput. Electron. Agric. vol. 102 (2014) 120–126, https://doi.org/10.1016/j.compag.2014.01.008.
[44] W. Fei, G.A. Narsilio, M.M. Disfani, Predicting effective thermal conductivity in sands using an artificial neural network with multiscale microstructural
parameters, Int. J. Heat. Mass Transf. vol. 170 (120997) (2021) 1–12, https://doi.org/10.1016/j.ijheatmasstransfer.2021.120997.
[45] V. Gholami, M.J. Booij, E.N. Tehrani, M.A. Hadian, Spatial soil erosion estimation using an artificial neural network (ANN) and field plot data, Catena vol. 163
(2018) 210–218, https://doi.org/10.1016/j.catena.2017.12.027.
[46] J.J. Jeremiah, S.J. Abbey, C.A. Booth, A. Kashyap, Results of application of artificial neural networks in predicting geo-mechanical properties of stabilised clays
- a review, Geotechnics vol. 1 (2021) 147–171, https://doi.org/10.3390/geotechnics1010008.
[47] F.P. Nejad, M.B. Jaksa, Load-settlement behavior modeling of single piles using artificial neural networks and CPT data, Comput. Geotech. vol. 89 (2017) 9–21,
https://doi.org/10.1016/j.compgeo.2017.04.003.
[48] R.A. Mozumder, A.I. Laskar, Prediction of unconfined compressive strength of geopolymer stabilized clayey soil using artificial neural network, Comput.
Geotech. vol. 69 (2015) 291–300, https://doi.org/10.1016/j.compgeo.2015.05.021.
[49] S. Alzabeebee, S.A. Mohamad, R.K.S. Al-Hamd, Surrogate models to predict maximum dry unit weight, optimum moisture content and California bearing ratio
form grain size distribution curve, Road. Mater. Pavement Des. (2021) 1–18, https://doi.org/10.1080/14680629.2021.1995471.
[50] M.N. Amin, et al., Prediction of strength and CBR characteristics of chemically stabilized coal gangue: ANN and random forest tree approach, Materials vol. 15
(4330) (2022) 1–19, https://doi.org/10.3390/ma15124330.
[51] M. Das, A.K. Dey, Prediction of bearing capacity of stone columns placed in soft clay using ANN model, Geotech. Geol. Eng. vol. 36 (3) (2018) 1845–1861,
https://doi.org/10.1007/s10706-017-0436-0.
[52] S. Hao, T. Pabst, Prediction of CBR and resilient modulus of crushed waste rocks using machine learning models, Acta Geotech. vol. 17 (4) (2022) 1383–1402,
https://doi.org/10.1007/s11440-022-01472-1.
[53] T.F. Kurnaz, Y. Kaya, Prediction of the California bearing ratio (CBR) of compacted soils by using GMDH-type neural network, Eur. Phys. J. vol. 134 (326)
(2019) 1–15, https://doi.org/10.1140/epjp/i2019-12692-0.
[54] M.M. Nujid, M.E. Michael, D.A. Tholibon, Failure assessment of strength and bearing capacity on marine stabilized subgrade soil, J. Fail. Anal. Prev. vol. 21 (6)
(2021) 1925–1942, https://doi.org/10.1007/s11668-021-01232-5.
[55] A.R. Tenpe, A. Patel, Application of genetic expression programming and artificial neural network for prediction of CBR, Road. Mater. Pavement Des. vol. 21
(5) (2020) 1183–1200, https://doi.org/10.1080/14680629.2018.1544924.
[56] D.K. Trong, et al., On random subspace optimization-based hybrid computing models predicting the California bearing ratio of soils, Materials vol. 14 (6516)
(2021) 1–20, https://doi.org/10.3390/ma14216516.
[57] B. Yildirim, O. Gunaydin, Estimation of California bearing ratio by using soft computing systems, Expert Syst. Appl. vol. 38 (5) (2011) 6381–6391, https://doi.
org/10.1016/j.eswa.2010.12.054.
[58] V.Y. Katte, S.M. Mfoyet, B. Manefouet, A.S. Ludovic Wouatong, L.A. Bezeng, Correlation of California bearing ratio (CBR) value with soil properties of road
subgrade soil, Geotech. Geol. Eng. vol. 37 (1) (2019) 217–234, https://doi.org/10.1007/s10706-018-0604-x.
[59] Y. Gurtug, A. Sridharan, Compaction behaviour and prediction of its characteristics of fine grained soils with particular reference to compaction energy, Soils
Found. vol. 44 (5) (2004) 27–36, https://doi.org/10.3208/sandf.44.5_27.
[60] A.V. Hohn, R.F. Leme, F.C. da Silva Filho, T.E. Moura, G.R. Llanque A, Empirical models to predict compaction parameters for soils in the State of Ceará,
Northeastern Brazil, e86328, Ing. e Investig. vol. 42 (1) (2022) 1–11, https://doi.org/10.15446/ing.investig.v42n1.86328.
[61] H. Mujtaba, K. Farooq, N. Sivakugan, B.M. Das, Correlation between gradational parameters and compaction characteristics of sandy soils, Int. J. Geotech. Eng.
vol. 7 (4) (2013) 395–401, https://doi.org/10.1179/1938636213Z.00000000045.
[62] W. Hassan, M. Ahmad, A. Farooq, A. Ajwad, H.Q. Ali, Y. Ilyas, Correlation of maximum laboratory dry density and optimum moisture content of soil with soil
parameters, NFC-IEFR J. Eng. Sci. Res. vol. 5 (1) (2017) 1–6, https://doi.org/10.24081/nijesr.2017.1.0003.

15
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

[63] I. Narváez-Bandera, D. Suárez-Gómez, C.E. Isaza, M. Cabrera-Ríos, Multiple criteria optimization (MCO): a gene selection deterministic tool in RStudio, PLoS
One vol. 17 (1) (2022), e0262890, https://doi.org/10.1371/journal.pone.0262890.
[64] J. Racine, RStudio: a platform-independent IDE for R and sweave, J. Appl. Econom. vol. 27 (1) (2012) 167–172, https://doi.org/10.1002/jae.1278.
[65] C.J. Shedlock, K.A. Stumpo, Data parsing in mass spectrometry imaging using R studio and cardinal: a tutorial, J. Mass Spectrom. Adv. Clin. Lab vol. 23 (2022)
58–70, https://doi.org/10.1016/j.jmsacl.2021.12.007.
[66] R. Polo-Mendoza, R. Peñabaena-Niebles, F. Giustozzi, G. Martinez-Arguelles, Eco-friendly design of warm mix asphalt (WMA) with recycled concrete aggregate
(RCA): a case study from a developing country, Constr. Build. Mater. vol. 326 (126890) (2022) 1–16, https://doi.org/10.1016/j.conbuildmat.2022.126890.
[67] T.L. Weissgerber, O. Garcia-Valencia, V.D. Garovic, N.M. Milic, S.J. Winham, Why we need to report more than ’data were analyzed by t-tests or ANOVA’, Elife
vol. 7 (e36163) (2018) 1–16, https://doi.org/10.7554/eLife.36163.
[68] K. Anders, Resolution of students t-tests, ANOVA and analysis of variance components from intermediary data, Biochem. Med. vol. 27 (2) (2017) 253–258,
https://doi.org/10.11613/BM.2017.026.
[69] P. Mishra, U. Singh, C. Pandey, P. Mishra, G. Panday, Application of student’s t-test, analysis of variance, and covariance, Ann. Card. Anaesth. vol. 22 (4)
(2019) 407–411, https://doi.org/10.4103/aca.ACA_94_19.
[70] J. Duque, W. Fuentes, J. Barros, Effect of grain size distribution on the maximum and minimum void ratios of granular soils, Acta Geotech. Slov. vol. 17 (2)
(2020) 26–33, https://doi.org/10.18690/actageotechslov.17.2.26-33.2020.
[71] F. Yang-chun, H. Yan-chun, M. Xiu-min, The application of Student’s t-test in internal quality control of clinical laboratory, Front. Lab. Med. vol. 1 (3) (2017)
125–128, https://doi.org/10.1016/j.flm.2017.09.002.
[72] S. Zhao, et al., Attach importance of the bootstrap t-test against Student’s t-test in clinical epidemiology: a demonstrative comparison using COVID-19 as an
example, Epidemiol. Infect. vol. 149 (e107) (2021) 1–6, https://doi.org/10.1017/S0950268821001047.
[73] D. Wang, H. Zhang, R. Liu, W. Lv, D. Wang, t-Test feature selection approach based on term frequency for text categorization, Pattern Recognit. Lett. vol. 45 (1)
(2014) 1–10, https://doi.org/10.1016/j.patrec.2014.02.013.
[74] D. Campbell, S. Lele, An ANOVA test for parameter estimability using data cloning with application to statistical inference for dynamic systems, Comput. Stat.
Data Anal. vol. 70 (2014) 257–267, https://doi.org/10.1016/j.csda.2013.09.013.
[75] G. Zhang, R. Christensen, J. Pesko, Parametric boostrap and objective Bayesian testing for heteroscedastic one-way ANOVA, Stat. Probab. Lett. vol. 174
(109095) (2021) 1–7, https://doi.org/10.1016/j.spl.2021.109095.
[76] M. Bahrami, M. Akbari, S.A. Bagherzadeh, A. Karimipour, M. Afrand, M. Goodarzi, Develop 24 dissimilar ANNs by suitable architectures & training algorithms
via sensitivity analysis to better statistical presentation: measure MSEs between targets & ANN for Fe–CuO/Eg–Water nanofluid, Phys. A vol. 519 (2019)
159–168, https://doi.org/10.1016/j.physa.2018.12.031.
[77] W. Liu, Z. Wang, X. Liu, N. Zeng, Y. Liu, F.E. Alsaadi, A survey of deep neural network architectures and their applications, Neurocomputing vol. 234 (2017)
11–26, https://doi.org/10.1016/j.neucom.2016.12.038.
[78] R. Polo-Mendoza, G. Martinez-Arguelles, R. Peñabaena-Niebles, Environmental optimization of warm mix asphalt (WMA) design with recycled concrete
aggregates (RCA) inclusion through artificial intelligence (AI) techniques, Results Eng. vol. 17 (100984) (2023) 1–15, https://doi.org/10.1016/j.
rineng.2023.100984.
[79] M. Mohanraj, S. Jayaraj, C. Muraleedharan, Applications of artificial neural networks for refrigeration, air-conditioning and heat pump systems - a review,
Renew. Sustain. Energy Rev. vol. 16 (2012) 1340–1358, https://doi.org/10.1016/j.rser.2011.10.015.
[80] H. Rizk, M. Abbas, M. Youssef, Device-independent cellular-based indoor location tracking using deep learning, Pervasive Mob. Comput. vol. 75 (2021),
101420, https://doi.org/10.1016/j.pmcj.2021.101420.
[81] M. Dudzik, A.M. Strȩk, ANN architecture specifications for modelling of open-cell aluminum under compression, Math. Probl. Eng. vol. 2834317 (2020) 1–26,
https://doi.org/10.1155/2020/2834317.
[82] W. Cao, X. Wang, Z. Ming, J. Gao, A review on neural networks with random weights, Neurocomputing vol. 275 (2018) 278–287, https://doi.org/10.1016/j.
neucom.2017.08.040.
[83] O. Abdeljaber, O. Avci, S. Kiranyaz, M. Gabbouj, D.J. Inman, Real-time vibration-based structural damage detection using one-dimensional convolutional
neural networks, J. Sound Vib. vol. 388 (2017) 154–170, https://doi.org/10.1016/j.jsv.2016.10.043.
[84] Z. Yang, F. Gao, S. Fu, M. Tang, D. Liu, Overfitting effect of artificial neural network based nonlinear equalizer: from mathematical origin to transmission
evolution, Sci. China Inf. Sci. vol. 63 (160305) (2020) 1–13, https://doi.org/10.1007/s11432-020-2873-x.
[85] A.P. Piotrowski, J.J. Napiorkowski, A comparison of methods to avoid overfitting in neural networks training in the case of catchment runoff modelling,
J. Hydrol. vol. 476 (2013) 97–111, https://doi.org/10.1016/j.jhydrol.2012.10.019.
[86] L. Li, Z. Qiao, Y. Liu, Y. Chen, A convergent smoothing algorithm for training max-min fuzzy neural networks, Neurocomputing vol. 260 (2017) 404–410,
https://doi.org/10.1016/j.neucom.2017.04.046.
[87] V.Q. Tran, H.-V.Thi Mai, T.-A. Nguyen, H.-B. Ly, Investigation of ANN architecture for predicting the compressive strength of concrete containing GGBFS, PLoS
One vol. 16 (12) (2021), e0260847, https://doi.org/10.1371/journal.pone.0260847.
[88] L. Zhang, P.N. Suganthan, A comprehensive evaluation of random vector functional link networks, Inf. Sci. vol. 367–368 (2016) 1094–1105, https://doi.org/
10.1016/j.ins.2015.09.025.
[89] M.J. Kearns, U.V. Vazirani, An introduction to computational learning theory, ISBN: 9780262111935, MIT press, 1994.
[90] V. Vapnik, R. Izmailov, Rethinking statistical learning theory: learning using statistical invariants, Mach. Learn. vol. 108 (3) (2019) 381–423, https://doi.org/
10.1007/s10994-018-5742-0.
[91] B.P. Chapman, A. Weiss, P.R. Duberstein, Statistical learning theory for high dimensional prediction: application to criterion-keyed scale development,
Psychol. Methods vol. 21 (4) (2016) 603–620, https://doi.org/10.1037/met0000088.
[92] A. Tavanaei, M. Ghodrati, S.R. Kheradpisheh, T. Masquelier, A. Maida, Deep learning in spiking neural networks, Neural Netw. vol. 111 (2019) 47–63, https://
doi.org/10.1016/j.neunet.2018.12.002.
[93] N. Kriegeskorte, T. Golan, Neural network models and deep learning, Curr. Biol. vol. 29 (2019) R231–R236, https://doi.org/10.1016/j.cub.2019.02.034.
[94] C. Menares, P. Perez, S. Parraguez, Z. Fleming, Forecasting PM2.5 levels in Santiago de Chile using deep learning neural networks, Urban Clim. vol. 38 (2021),
100906, https://doi.org/10.1016/j.uclim.2021.100906.
[95] S. Basu, et al., Deep neural networks for texture classification - a theoretical analysis, Neural Netw. vol. 97 (2018) 173–182, https://doi.org/10.1016/j.
neunet.2017.10.001.
[96] K. Xu, A.M. Tartakovsky, J. Burghardt, E. Darve, Learning viscoelasticity models from indirect data using deep neural networks, Comput. Methods Appl. Mech.
Eng. vol. 387 (2021), 114124, https://doi.org/10.1016/j.cma.2021.114124.
[97] J. Han, A. Jentzen, Weinan E, Solving high-dimensional partial differential equations using deep learning, Proc. Natl. Acad. Sci. U. S. A. vol. 115 (34) (2018)
8505–8510, https://doi.org/10.1073/pnas.1718942115.
[98] A. Müller, A. Taras, Prediction of the local buckling strength and load-displacement behaviour of SHS and RHS members using deep neural networks (DNN) –
introduction to the deep neural network direct stiffness method (DNN-DSM), Steel Constr. vol. 15 (2022) 78–90, https://doi.org/10.1002/stco.202100047.
[99] Y. Bengio, A. Courville, P. Vincent, Representation learning: a review and new perspectives, IEEE Trans. Pattern Anal. Mach. Intell. vol. 35 (8) (2013)
1798–1828, https://doi.org/10.1109/TPAMI.2013.50.
[100] A. Lozano-Diez, R. Zazo, D.T. Toledano, J. Gonzalez-Rodriguez, An analysis of the influence of deep neural network (DNN) topology in bottleneck feature
based language recognition, PLoS One vol. 12 (8) (2017), e0182580, https://doi.org/10.1371/journal.pone.0182580.
[101] E. Mansouri, M. Manfredi, J.-W. Hu, Environmentally friendly concrete compressive strength prediction using hybrid machine learning, Sustainability vol. 14
(12990) (2022) 1–17, https://doi.org/10.3390/su142012990.
[102] A. Apicella, F. Donnarumma, F. Isgrò, R. Prevete, A survey on modern trainable activation functions, Neural Netw. vol. 138 (2021) 14–32, https://doi.org/
10.1016/j.neunet.2021.01.026.

16
R. Polo-Mendoza et al. Case Studies in Construction Materials 20 (2024) e02800

[103] M. Murray, V. Abrol, J. Tanner, Activation function design for deep networks: linearity and effective initialisation, Appl. Comput. Harmon. Anal. vol. 59 (2022)
117–154, https://doi.org/10.1016/j.acha.2021.12.010.
[104] Z. Wang, H. Liu, F. Liu, D. Gao, Why KDAC? A general activation function for knowledge discovery, Neurocomputing vol. 501 (2022) 343–358, https://doi.
org/10.1016/j.neucom.2022.06.019.
[105] A. Yilmaz, R. Poli, Successfully and efficiently training deep multi-layer perceptrons with logistic activation function simply requires initializing the weights
with an appropriate negative mean, Neural Netw. vol. 153 (2022) 87–103, https://doi.org/10.1016/j.neunet.2022.05.030.
[106] Z. Liao, Z. Wang, H. Yamahara, H. Tabata, Echo state network activation function based on bistable stochastic resonance, Chaos Solitons Fractals Nonlinear Sci.
Nonequilibrium Complex Phenom. vol. 153 (111503) (2021) 1–14, https://doi.org/10.1016/j.chaos.2021.111503.
[107] B.N. Örnek, S.B. Aydemir, T. Düzenli, B. Özak, Some remarks on activation function design in complex extreme learning using Schwarz lemma,
Neurocomputing vol. 492 (2022) 23–33, https://doi.org/10.1016/j.neucom.2022.04.010.
[108] S.-L. Shen, N. Zhang, A. Zhou, Z.-Y. Yin, Enhancement of neural networks with an alternative activation function tanhLU, Expert Syst. Appl. vol. 199 (117181)
(2022) 1–13, https://doi.org/10.1016/j.eswa.2022.117181.
[109] K. Adem, Impact of activation functions and number of layers on detection of exudates using circular Hough transform and convolutional neural networks,
Expert Syst. Appl. vol. 203 (117583) (2022) 1–7, https://doi.org/10.1016/j.eswa.2022.117583.
[110] Y. Liu, Z. Wang, Q. Ma, H. Shen, Multistability analysis of delayed recurrent neural networks with a class of piecewise nonlinear activation functions, Neural
Netw. vol. 152 (2022) 80–89, https://doi.org/10.1016/j.neunet.2022.04.015.
[111] G. Bingham, R. Miikkulainen, Discovering parametric activation functions, Neural Netw. vol. 148 (2022) 48–65, https://doi.org/10.1016/j.
neunet.2022.01.001.
[112] D. Dũng, V.K. Nguyen, Deep ReLU neural networks in high-dimensional approximation, Neural Netw. vol. 142 (2021) 619–635, https://doi.org/10.1016/j.
neunet.2021.07.027.
[113] D. Boob, S.S. Dey, G. Lan, Complexity of training ReLU neural network, Discret. Optim. vol. 44 (100620) (2022) 1–16, https://doi.org/10.1016/j.
disopt.2020.100620.
[114] Z. Cai, J. Chen, M. Liu, Least-squares ReLU neural network (LSNN) method for scalar nonlinear hyperbolic conservation law, Appl. Numer. Math. vol. 174
(2022) 163–176, https://doi.org/10.1016/j.apnum.2022.01.002.
[115] S. Langer, Approximating smooth functions by deep neural networks with sigmoid activation function, J. Multivar. Anal. vol. 182 (104696) (2021) 1–21,
https://doi.org/10.1016/j.jmva.2020.104696.
[116] Z. Guo, S. Ou, J. Wang, Multistability of switched neural networks with sigmoidal activation functions under state-dependent switching, Neural Netw. vol. 122
(2020) 239–252, https://doi.org/10.1016/j.neunet.2019.10.012.
[117] S. Xu, X. Li, C. Xie, H. Chen, C. Chen, Z. Song, A high-precision implementation of the sigmoid activation function for computing-in-memory architecture,
Micromachines vol. 12 (1183) (2021) 1–12, https://doi.org/10.3390/mi12101183.
[118] M. Tanaka, Weighted sigmoid gate unit for an activation function of deep neural network, Pattern Recognit. Lett. vol. 135 (2020) 354–359, https://doi.org/
10.1016/j.patrec.2020.05.017.
[119] D. Xu, S. Zhang, H. Zhang, D.P. Mandic, Convergence of the RMSProp deep learning method with penalty for nonconvex optimization, Neural Netw. vol. 139
(2021) 17–23, https://doi.org/10.1016/j.neunet.2021.02.011.
[120] M.C. Mukkamala M. Hein ,“Variants of RMSProp and Adagrad with logarithmic regret bounds,” 34th Int. Conf. Mach. Learn. Sydney, Aust., vol. 70, pp.
2545–2553, 2017.
[121] N. Saqib, K.F. Haque, V.P. Yanambaka, A. Abdelgawad, Convolutional-neural-network-based handwritten character recognition: an approach with massive
multisource data, Algorithms vol. 15 (129) (2022) 1–25, https://doi.org/10.3390/a15040129.
[122] KERAS, Keras API reference / optimizers / RMSprop, Keras Website (2021).
[123] M.E. Khan Z. Liu V. Tangkaratt Y. Gal, “Vprop: Variational Inference using RMSprop,” in 31st Conference on Neural Information Processing Systems (NIPS),
Long Beach, CA, USA., 2017, pp. 1–8.
[124] D. Li, J. Zhao, J. Ma, Experimental studies on rock thin-section image classification by deep learning-based approaches, Mathematics vol. 10 (2317) (2022)
1–28, https://doi.org/10.3390/math10132317.
[125] V. Ojha, G. Nicosia, Backpropagation neural tree, Neural Netw. vol. 149 (2022) 66–83, https://doi.org/10.1016/j.neunet.2022.02.003.
[126] R. Polo-Mendoza, G. Martinez-Arguelles, R. Peñabaena-Niebles, E. Covilla-Valera, Neural networks implementation for the environmental optimisation of the
recycled concrete aggregate inclusion in warm mix asphalt, Road. Mater. Pavement Des. (2023) 1–26, https://doi.org/10.1080/14680629.2023.2230298.
[127] A.A. Alabdullah, M. Iqbal, M. Zahid, K. Khan, M.N. Amin, F.E. Jalal, 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. vol. 345 (128296) (2022) 1–13, https://doi.org/10.1016/
j.conbuildmat.2022.128296.
[128] Y. Kim, Y. Kim, Explainable heat-related mortality with random forest and SHapley additive exPlanations (SHAP) models, Sustain. Cities Soc. vol. 79 (103677)
(2022) 1–15, https://doi.org/10.1016/j.scs.2022.103677.
[129] K. Lin, Y. Gao, Model interpretability of financial fraud detection by group SHAP, Expert Syst. Appl. vol. 210 (118354) (2022) 1–9, https://doi.org/10.1016/j.
eswa.2022.118354.
[130] X. Li, et al., Development of an interpretable machine learning model associated with heavy metals’ exposure to identify coronary heart disease among US
adults via SHAP: findings of the US NHANES from 2003 to 2018, Chemosphere vol. 311 (137039) (2023) 1–9, https://doi.org/10.1016/j.
chemosphere.2022.137039.
[131] N. Nordin, Z. Zainol, M.H. Mohd Noor, L.F. Chan, An explainable predictive model for suicide attempt risk using an ensemble learning and Shapley additive
explanations (SHAP) approach, Asian J. Psychiatr. vol. 79 (103316) (2023) 1–7, https://doi.org/10.1016/j.ajp.2022.103316.
[132] L.F. Walubita, G. Martinez-Arguelles, R. Polo-Mendoza, S. Ick-Lee, L. Fuentes, Comparative environmental assessment of rigid, flexible, and perpetual
pavement: a case study of Texas,”, Sustainability vol. 14 (9983) (2022) 1–22, https://doi.org/10.3390/su14169983.
[133] J. Clavreul, D. Guyonnet, T.H. Christensen, Quantifying uncertainty in LCA-modelling of waste management systems, Waste Manag. vol. 32 (12) (2012)
2482–2495, https://doi.org/10.1016/j.wasman.2012.07.008.
[134] D. Vega-Araujo, J. Santos, G. Martinez-Arguelles, Environmental performance evaluation of warm mix asphalt with recycled concrete aggregate for road
pavements, Int. J. Pavement Eng. (2022) 1–14, https://doi.org/10.1080/10298436.2022.2064999.
[135] A. Önalp, E. Bol, A. Özocak, S. Sert, N. Ural, E. Arel, Influence of index properties on the cyclic failure of fine-grained soils, Eng. Geol. vol. 317 (107056) (2023)
1–17, https://doi.org/10.1016/j.enggeo.2023.107056.
[136] F. Stanić, et al., A new multifractal-based grain size distribution model, Geoderma vol. 404 (115294) (2021) 1–12, https://doi.org/10.1016/j.
geoderma.2021.115294.
[137] P.I.J. Thompson, A.J. Dugmore, A.J. Newton, N.A. Cutler, R.T. Streeter, The influence of burial rate on variability in tephra thickness and grain size
distribution in Iceland, Catena vol. 225 (107025) (2023) 1–13, https://doi.org/10.1016/j.catena.2023.107025.

17

You might also like