Uploaded by Vikram Kumar

SS-XGBoost feature engineering

advertisement
Materials Today Communications 35 (2023) 105793
Contents lists available at ScienceDirect
Materials Today Communications
journal homepage: www.elsevier.com/locate/mtcomm
Robust extreme gradient boosting regression model for compressive
strength prediction of blast furnace slag and fly ash concrete
M. Iqbal Khan *, Yassir M. Abbas
Department of Civil Engineering, King Saud University, Riyadh 800–11421, Saudi Arabia
A R T I C L E I N F O
A B S T R A C T
Keywords:
Concrete
Compressive strength
Supplementary cementitious materials
Machine learning
Regression
Grading boost
In this study, a novel machine learning (ML) technique, eXtreme Gradient Boosting (XG Boost), was employed to
train an extremely precise ML model. The developed XG Boost model was highly interpretable, filling the gap and
opening black boxes in the literature. The study provides further a simple and free user interface to support the
design of normal- and high-strength Blast Furnace Slag (BFS), and fly ash (FA) concrete. The compressive
strength of 1030 concrete mixes containing cement (C), BFS, and FA were collected and analyzed. The baseline
model tend to overfit, with R2 values of 0.996 and 0.919 for the training and testing datasets, respectively. The
hyperparameters of the model have been optimized using vector multi-objective optimization to maximize the
prediction capability of the model. The optimized XG Boost model exhibited a superior prediction performance
with R2 of 0.992 and 0.949 for the training and testing datasets. Based on Gini indexes and SHAP values, C, FA,
water, and aggregate were the most significant model parameters. According to this study, the best BFS, FA, sand,
and superplasticizer contents for concrete strength optimization were 100–200, 100–200, 600–800, and
7–13 kg/m3, respectively. The SP has a negligible effect on concrete’s compressive strength at low water contents
(less than 180 kg/m3), but a stochastic effect at high contents. The various chemical properties of high-range
water reducers may have resulted in the randomly generated response in the current study.
1. Introduction
1.1. Concrete– an overview
Due to its availability of raw materials, cost-effectiveness, and ad­
vantageous mechanical properties, concrete is the most consumable
structural material. Every year, around 32 billion tons of concrete are
produced worldwide (more than 4 tons per capita and the demand for
concrete is expected to grow). A growing urban population (the global
population is projected to rise from 30% to 54% in 2050) will drive more
concrete demand for production in the coming decades [1,2]. However,
about 8% of all anthropogenic greenhouse gas emissions can be attrib­
uted to concrete’s current ecological footprint [3]. A variety of supple­
mentary cementitious materials [SCM such as fly ash (FA), blast furnace
slag (BFS), and silica fume (SF)] have been shown to enhance the con­
crete’s strength and durability and scale down its environmental impact
[4–6].
Due to its robust relationship with structural reliability, concrete’s
compressive strength is probably its most significant characteristic [7,
8]. A concrete’s compressive strength is crucial to the design of
structures, and it is practically correlated with various other properties.
The compressive strength of concrete is not solely dependent on the
water–binder ratio but also on the cementitious types and contents [9,
10]. Overall, this mechanical property is extremely nonlinear and
responsive to all constituents of concrete and age. It was shown the
strength of concrete is enhanced by the increase of sodium hydroxide
molarity and sodium hydroxide to sodium silicate, whereas it decreases
by increasing the curing temperature [11]. A key factor affecting a
concrete’s strength is the percentage of FA and BFS and the ratio of
water to binders [12].
1.2. Machine learning – an overview
ML is a form of artificial intelligence that enables systems to learn
just like humans without explicit programming. In general, ML models
are based on side-range datasets (training data) [13,14]. It is typical to
categorize ML methodologies according to the feedback to the learning
algorithm into supervised-, unsupervised-, and reinforcement-learning.
A list of the common supervised and unsupervised ML systems is given
in Table 1. It is noteworthy that the XG Boost and t-Distributed Sto­
chastic Neighbor Embedding (t-SNE) [15] are the most modern and
* Corresponding author.
E-mail address: miqbal@ksu.edu.sa (M.I. Khan).
https://doi.org/10.1016/j.mtcomm.2023.105793
Received 24 January 2023; Received in revised form 7 March 2023; Accepted 10 March 2023
Available online 11 March 2023
2352-4928/© 2023 Elsevier Ltd. All rights reserved.
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
businesses [23].
Moreover, the t-SNE approach is an unsupervised ML approach that
was originally proposed by Hinton and Roweis [24] and considered a
technique of visualization in prototype form [25]. It has been gaining
traction in real-life problems over the past decade. ML of quantum states
has also proven beneficial using such a technique. A study by [26]
concluded that using of t-SNE in this type of application is one of the
most promising techniques. Using a state-of-the-art dimensionality
reduction algorithm, usable data distributions can be represented in a
non-linear manner, resulting in a low-dimensional representation (map)
[27,28]. In order to conduct SNE operations, two steps must be
completed. The first step in SNE consists of permuting the distance be­
tween two data points into a probability based on their similarity in
multidimensional space. Secondly, the second significant component of
SNE is that it combines the conditional probability of a point in high
Nomenclature
n
xi , yi
x
x, y
ai
âi
ai
L(y, f(x))
M
α
size of sample
A sample point indexed with i
The sample mean of x variable and analogously for y
Sample means for variables x and, respectively
The targeted (observed) strength of the UHPC
The predicted strength of the UHPC by the RDF model
The mean of targeted (observed) strength of the UHPC
Loss function with differentiability
Learners with a weak response
Learning rate
Table 1
The popular algorithms of supervised and unsupervised ML.
ML technique
The goal
Example algorithms
Supervised
learning
Classification and
regression
Unsupervised
learning
Clustering and data
visualization
Artificial neural networks (ANNs), support vector machine (SVM), Decision trees (DT), random forest, Lasso regression, Multiple
regression (MR), Multiple additives (MA), Light gradient boosting machine (LGBM), Gradient boosting, and Extreme gradient
boosting (XG Boost).
K-means, Mean-shift clustering, DBSCAN clustering, Gaussian mixture, Spectral Clustering, Agglomerative Clustering, and an
interactive high-dimensional data visualization technique called "t-SNE"[15].
promising ML approaches that have shown superior accuracy, and merit
special mention.
The XG Boost is a decision tree-based collective ML system that
routinely uses gradient boosting to develop a regression of the classifi­
cation models [16]. The algorithm was initially developed by Chen et al.
[17] as an efficient application of the gradient boosting methodology
introduced by Friedman et al. [18]. The XG Boost algorithm is charac­
terized by various advantages over gradient boosting, such as smart
splitting of trees, short leaf nodes, randomization, Newton-Raphson
boosting, and out-of-core modeling [19]. As a result of its integration
into the Python programming language, and use in various Kaggle [20]
competitions, it has become increasingly popular. The XG Boost algo­
rithm has been the basis for wide range of pioneering applications in
recent years. Recent years have seen a wide range of pioneering appli­
cations based on the XG Boost algorithm. This has included but is not
limited to, diagnosing human health problems [21], regarding the
COVID-19 pandemic [22], and forecasting financial bankruptcy in
dimensions with the conditional probability of other map points in low
dimensions in order to arrive at a conditional probability of a point in
high dimensions [29]. Because of the large number of observations that
t-SNE must accommodate, it has the disadvantage of being unable to be
scaled [28,30].
Literature has well-documented successful ML applications in
structural engineering analysis and design [31–33]. The popularity of
ML methods has led scientists to focus more on predicting concrete
properties by using these approaches. Table 2 summarizes the reported
studies in predicting concrete’s compressive strength by ML-based
regression models. Although there have been a variety of ML methods
used, ANN-based approaches are the most common. There is no practical
application of these methods, as they are considered black boxes and
cannot be widely applied. Furthermore, the performance of the modern
gradient boosting ensemble methods (such as AdaBoost, LGBM, and XG
Boost) was the highest. Moreover, reliable models are lacking that are
accurately capable of predicting the compressive strength of concrete
Table 2
Summary of some previous studies.
Ref.
Concrete type
Data size
Number of input variables
[34]
[35]
[9]
High-performance concrete (HPC)
Normal concrete (NC)
HPC
727
864
1030
8
9
8
[36]
HPC
1030
8
[37]
HPC
152
8
[38]
Ultrahigh performance concrete
931
17
2
ML method
R2 (Test data)
ANN
ANN
ANN
MR
SVM
MA
DT
ANN
Bagged ANN
Gradient-boosted ANN
Wavelet bagged ANN
Wavelet Gradient-boosted ANN
LGBM
CAT boost regressor
Gradient boosting regressor
AdaBoost regressor
XG Boost
XG Boost
0.914
0.578
0.909
0.611
0.886
0.911
0.890
0.909
0.928
0.927
0.940
0.953
0.950
0.950
0.960
0.900
0.940
0.892
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
efficiency of a ML model [47]. As such, establishing a model of concrete
compressive strength requires the existence of vast amounts of data from
diverse real-world environments. In the current study, the raw experi­
mental data were collected from the University of California, Irvine
(UCI), repository [48,49]. A total of 1030 concrete mixes containing
OPC (type I), BFS, and FA under normal moisture curing were gathered
from 11 different experimental sources [50–60]. The data attributes
consist of eight inputs and one quantitative output (the compressive
strength of concrete). The units and coding system of these variables are
displayed in Table 3. Among the characteristics of the collected dataset
are a maximum aggregate size of 10 mm and naphthalene-based
superplasticizers. Additionally, concrete strength is determined by
standard methods using 150 × 300 mm cylinders.
Table 4 presents the statistical analysis of the variables in the ML
model, while Fig. 1 displays their frequency distributions. A majority of
the model’s variables have reasonable frequency distributions for use in
ML regression. Further, the developed model will probably apply to
ordinary and high-strength concrete (with strengths ranging from 2.3 to
83.6 MPa) containing cement, BFS, and FA of 102–540, 0–359, and
0–200 kg/m3, respectively.
In this research, an evaluation of the linear correlation between the
model variables was analyzed through the calculation of Pearson’s
correlation coefficient (rxy , Eq. 1) during data preprocessing. This coef­
ficient measures the standard deviation of a line of covariance, with
values ranging from − 1–+1 [61]. The result of this analysis is presented
in Fig. 2. The maximum positive effect of these variables on the model’s
label (CS) was for C, SP, and Age, whereas increases in FA, CA, and water
are likely to reduce the model’s compressive strength. The results are
fairly predictable since FA has a favorable effect on the durability
properties of concrete, but a negative effect on its strength.
Table 3
The database coding system.
Variable
Portland Cement
Blast Furnace Slag
Fly Ash
Water
Superplasticizer
Coarse Aggregate
Sand
Age
Compressive strength
Coded variable
C
BFS
FA
W
SP
CA
S
A
CS
Unit
kg/m
Variable type
3
days
MPa
Predictor (feature)
Target (label)
with different classes (for example, NC and HPC, or HPC and UHPC).
The standard non-optimized ML model generally converges to a local
optimum or overtrains, has slow calculation speeds, and does not
incorporate optimizations [39–41]. In the absence of regular optimiza­
tion of these models, the forecasting precision often remains low due to
the subjective criteria that are used for determining the parameters.
Thus, it is commonly used to optimize parameters via optimization al­
gorithms [e.g., particle swarm optimization algorithm (PSO) or genetic
algorithm (GA)] to improve prediction accuracy [42]. It is important to
recognize, however, that there are a number of inherent limitations
associated with these optimization algorithms. Some of the reasons for
these weaknesses are insufficient calculation speed and inability to
reach a global optimal [43]. The Mind Evolutionary Algorithm (MEA)
has been proposed by Chengyi et al. [44] as a way to overcome the
shortcomings of existing algorithms. It is noteworthy that hyper­
parameters refer to a set of parameters that can significantly impact
forecasting accuracy [45]. Prior to the modeling process, it is important
to optimize the hyperparameters of a ML model to ensure that it will
work successfully.
n
∑
(xi − x)(yi − y)
i=1
̅√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
̅
rxy = √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
n
n
∑
∑
(xi − x)
(yi − y)
1.3. Research objectives, significance, and rationale
i=1
A substantial amount of time and effort is required to optimize the
strength of concrete. Experimental studies would be cost-efficient and
time-saving if a rational and robust prediction model could be devel­
oped. Typically, linear and nonlinear empirical-based models are less
accurate, since they rely on a limited sample of data, ambient and curing
conditions, and testing norms [46]. The shortcomings of empirical
models could be addressed by implementing ML tools. This study aimed
to efficiently model and deploy the compressive strength of NC and HPC
containing BFS and FA. For this purpose, the Python code with the XG
boost regression algorithm was developed and fine-tuned. The model
was eventually made more practical by the development of a graphical
user interface (GUI).
(1)
i=1
Where, n is the size of sample, (xi , yi ) is A sample point indexed with i,
x, y is the sample mean of x variable and analogously for y.
2.2. Feature engineering
Another phase of data preprocessing was performed to identify
outliers in the database. Accordingly, boxplots for the data features were
analyzed, as shown in Fig. 3. There were several outliers for the age
feature because relatively few discrete data were available for long-term
strength measurements (180, 270, and 365 days). It is well reported that
the XG Boost method can handle outliers without notable sensitivity
[38], thus no outliers were removed from the database in the present
study. In addition, the removal of the outlier reduces the generalization
of the model. Feature engineering ultimately culminates in randomly
separating the entire database into training and test dataset sets. In the
current study, the training data sets were split by 75% and 25%,
respectively. Thereby, a total of 772 datasets were used to train the
model, and a further 258 datasets were used to test its validity.
2. Data population
2.1. Data collection, description, and statistical analysis
The generality of training data plays a significant role in the
Table 4
Descriptive statistics of the variables.
Variable
Mean
Std. Dev.
Median
Minimum
Maximum
Q1
Q3
C
BFS
FA
Water
SP
CA
Sand
Age
CS
281.2
73.9
54.2
181.6
6.2
972.9
773.6
45.7
35.8
104.5
86.3
64.0
21.4
6.0
77.8
80.2
63.2
16.7
272.9
22.0
0.0
185.0
6.4
968.0
779.5
28.0
34.4
102.0
0.0
0.0
121.8
0.0
801.0
594.0
1.0
2.3
540.0
359.4
200.1
247.0
32.2
1145.0
992.6
365.0
82.6
192.0
0.0
0.0
164.9
0.0
932.0
730.3
7.0
23.7
350.0
143.0
118.3
192.0
10.2
1029.4
824.3
56.0
46.2
3
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 1. Graphical summary of the variables of the model.
Fig. 2. (a) Pearson’s encoded matrix, and (b) linear correlations between features and label.
Fig. 3. Boxplots of the datasets.
4
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
̂f (x) = ̂f M (x) =
M
∑
(7)
̂f m (x)
m=0
In Eqs. (2–4), L(y, f(x)) denotes the loss function that has differen­
tiability behavior. Further, in Eq. (5), α represents the learning rate.
For XG Boost single trees, the model continually evaluates the loss
function in order to choose the leaf node that has the highest gain.
Splitting features allow the algorithm to add regression trees (i.e.,
introduce a new predictor, ̂f (x), to eliminate the residuals from pre­
m
vious calculations). Finally, the model prediction could be evaluated by
adding up the scores for each predictor. The current investigation was
coded in Python [62], using the flow chart shown in Fig. 4.
3.2. Model performance indicators
Normally, the coefficient of determination (R2 ) is used to test the
results of a regression model. Due to its vulnerability to ML averaging
procedures, it cannot be used solely to evaluate model output [63].
Therefore, the root mean squared error (RMSE), mean absolute percent
error (MAPE), and normalized mean bias error (NMBE) were evaluated
as well. A list of the performance indicators used in this study is pre­
sented in Eqs. (8–11).
Fig. 4. XG Boost algorithm flowchart.
Table 5
Hyperparameters of the main developed models.
Hyperparameter
Baseline model
Optimized model
Maximum tree depth for base learners
Subsample ratio of the training instance
Number of gradient boosted trees
Boosting learning rate
(
)
L1 yi , θ regularization term on weights
Subsample ratio of columns for each level
3
1
100
0.1
0
4
0.8
400
0.175
0.1
1
MAPE =
n
1∑
|ai − âi |
…
n i=1 | âi |
⎛
0.3
1
⎜n
⎜
NMBE = ⎜
⎝
3. XG boost regression
3.1. Model formulation and development
(9)
⎞
n
∑
(ai − âi )2 ⎟
⎟
i=1
⎟
⎠
ai
(10)
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
√∑
√n
√ (ai − âi )2
√
RMSE = i=1
n
As a general rule, XG Boost algorithms are prepared using the
following steps:
Step 1. set the model’s initial value, ̂f , to a constant as:
(11)
(0)
̂f (0) (x) = argθ min
n
∑
N
∑
(2)
L(yi , θ)
2
R = 1−
i=1
Step 2. for m = 1to M (Learners with a weak response):
- Calculate the slopes (̂
g m ) and canvases ( ̂
h m ):
[
]
∂L〈yi , f (xi )〉
̂
g m (xi ) =
∂f (xi )
f (x)=̂
f (m− 1) (x)
[
∂L〈yi , f (xi )〉
̂
h m (xi ) =
∂f (xi )2
(3)
4. Results and discussion
4.1. The baseline model
(4)
1) (x)
This research began by developing a baseline model based on stan­
dardized hyperparameters (Table 5). Table 6 and Fig. 5 summarize the
prediction capacity of this baseline model. The performance indicators
of the test data were marginally inferior to those of the training data,
which indicated that the baseline model exhibited a tendency toward
overfitting. These hyperparameters have thus been tuned to maximize
the model’s prediction abilities.
[
]2
N
∑
̂
g (xi )
1̂
̂ m = argϕ∈Φ min
h m (xi ) − m
ϕ
− ϕ(xi )
̂
2
h m (xi )
i=1
Table 6
Performance indicators of the main developed models.
(5)
Performance indicator
- Modify the model as follows:
̂f m (x) = ̂f (m− 1) (x) + ̂f m (x)…
(12)
âi 2
Where, ai is the targeted (observed) strength of concrete, âi is the pre­
dicted strength of the by the ML model, and ai is the mean of targeted
(observed) strength of concrete.
- Solve the following optimization problem using the training set,
{
}
̂
xi , − g m (xi ) , to fit the base learner (tree):
̂h m (xi )
̂f m (x) = α ϕ
̂ m (x)…
n
∑
i=1
]
f (x)=̂
f (m−
(ai − âi )2
i=1
MAPE
NMBE
RMSE
R2
(6)
Step 3. The final model is
5
Baseline model
Optimized model
Training set
Testing set
Training set
Testing set
0.381
0.999
0.999
0.996
2.945
21.857
4.675
0.919
0.932
2.321
1.524
0.992
2.428
13.780
3.712
0.949
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 5. Model vs. target for the baseline model: (a) training, and (b) testing datasets.
4.2. The optimized model
4.2.2. Important features
Two methods were used in this study to evaluate the feature
importance of the developed ML model. The first approach was based on
the Gini index, which was calculated by evaluating the total gain of all
when the feature was employed [65]. Fig. 8(a) shows the obtained
feature importance by this method, where C, FA, water, and both
aggregate types were the important features of the database. According
to Gini-based feature analysis, significant features with unique values
are more likely to be detected [66].
Likewise, the model’s sensitivity to various features has been
investigated using the SHAP (an acronym from SHapley Additive Ex­
planations) value-based approach. A feature significance analysis re­
veals that low aggregate and cement content has a significant negative
impact on the model’s prediction, while high weights have a strong
positive impact as shown in Fig. 8(b). In contrast, low water content had
a substantial positive impact on the model response, whereas high water
content had a negative effect. It is interesting to note that some SHAP
values are similar to Pearson’s coefficient, especially for features with
negative Pearson’s values (i.e., FA, W, CA, and S).
4.2.1. Properties and performance
During this study, the most significant hyperparameters (Table 5)
were optimized by trial and error to achieve the highest R2 values. This
was achieved by using vector-based (Pareto [64]) optimization. The red
point in Fig. 6 shows the optimum Pareto frontier that was obtained. The
optimized parameters were discovered at this point. The present inves­
tigation demonstrated superior predictive performance compared to
most ML models reported in the literature (Table 2) with scores of 0.992
and 0.949 (Table 6), respectively. As shown in Fig. 7, predicted-target
data were within an accuracy range of ± 85% for the test data, which
proves that the optimized model is a tremendous prediction tool. The
following sections present the feature importance and establish particle
dependence plots (PDPs) established with the aid of the calibrated
model.
4.2.3. Partial dependence analysis
In this study, partial dependence analysis was carried out for each of
the independent variables (i.e., C, BFS, FA, W, SP, CA, S, and A)
employed in the ML model. Fig. 9 shows the partial dependence plots
(PDPs) of concrete’s CS in response to different predictors. Various in­
dependent variables exhibit varying ranges of compressive strength in
this figure. In terms of strength difference, the A, C, and W were the most
influential parameters, while the CA, SP, and FA were the least ones.
This finding is consistent with that obtained from SHAP values [Fig. 8
(b)], and that reported in [67].
Fig. 9 also illustrates that concrete’s strength increases as its cement
content increases; however, an increase in water content will result in a
significant decline in strength. Further, the strength of concrete in­
creases significantly up to about 30 days, but afterward, it remains
Fig. 6. Pareto frontier results for hyperparameter optimization.
Fig. 7. Model vs. target for the optimized model: (a) training, and (b) testing datasets.
6
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 8. Feature importance by: (a) Gini index-, and (b) SHAP value-based methods.
Fig. 9. CS PDPs of: (a) C, (b) BFS, (c) FA, (d)W, (e) SP, (f) CA, (g) S, (h) A.
Fig. 10. PDPs of C and BFS: (a) isoresponse contours, and (b) response surface.
relatively stable. These findings are widely known, which strengthens
the reliability of the developed model. The concrete constituent mate­
rials are also given optimum values in Fig. 9. A general rule of thumb is
that the ideal FA and SP contents are 100–200 and 7–13 kg/m3,
respectively.
The concrete’s compressive strength PDPs shown in Fig. 10 through
Fig. 17 illustrate some of the most important mutual relationships (in the
existence of C and W). These plots would guide the precise selection of
concrete’s constituent materials. The 2D and 3D isoresponses of C and
BFS are provided in Fig. 10. A slight decrease in concrete’s compressive
7
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 11. PDPs of C and FA: (a) isoresponse contours, and (b) response surface.
Fig. 12. PDPs of C and W: (a) isoresponse contours, and (b) response surface.
strength was associated with an increase in BFS content. An observation
similar to that made by Türkmen et al. [68] concluded that BFS incor­
poration led to strength reductions in concrete, especially at early con­
crete ages. As a general guideline, the best content for C and BFS can be
found between 400 and 450 and 100–200 kg/m3, respectively.
An illustration of the concrete’s strength sensitivity to the C-FA
combination is shown in Fig. 11. In this figure, FA had a relatively small
effect on the concrete compressive strength for cement contents below
350 kg/m3, as no noticeable strength changes were observed with FA
addition. The C and W PDPs are presented in Fig. 12. As expected, the
maximum strength was attained at the lowest water content (less than
160 kg/m3) and highest cement content (more than 380 kg/m3).
Fig. 13 depicts the compressive strength of concrete as dependent on
C and S. The figure indicates that a higher ratio of sand to cement will
result in concrete with higher strength. The improvement in fracture
resistance of the concrete is likely to be caused by the increased inter­
locking at higher sand contents [69]. It seems that 600–800 kg/m3 of
sand is the optimal amount for concrete, as shown in Fig. 13.
Fig. 14 shows the development of the concrete’s compressive
strength with age for different cement dosages. It is evident from the
figure that strength increases significantly at early ages (up to 30 days),
but little strength gains occur at older ages. In general, it is known that
concrete materials will significantly increase in strength over time, with
a characteristic strength assessment taking place at 28 days.
Fig. 15 shows the combined effect of water and BFS on concrete’s
compressive strength. This figure illustrates, as discussed earlier, that
increasing BFS content marginally decreases strength at a constant
amount of water. The increase in water, however, significantly reduces
Fig. 13. PDPs of C and S: (a) isoresponse contours, and (b) response surface.
8
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 14. PDPs of C and A: (a) isoresponse contours, and (b) response surface.
Fig. 15. PDPs of W and BFS: (a) isoresponse contours, and (b) response surface.
Fig. 16. PDPs of W and FA: (a) isoresponse contours, and (b) response surface.
the strength. Normally, a water-binder ratio below 0.2 (about 150 kg/
m3 water content) is required for hydration [70]. As more water is
added, hardened cement will break away from the aggregate surface (as
a result of water lubrication at the molecular level).
An illustration of the concrete’s response to water and FA in­
teractions is shown in Fig. 16. No significant change was noted for FA
replacement compared to cement in terms of compressive strength. It is
inconsistent with the findings of various investigators [71–73], that FA
can insignificantly decrease concrete’s compressive strength at early
ages (1–7 days); however, it will not alter its long-term strength.
Fig. 17 shows how water and superplasticizer affect the concrete’s
compressive strength. As shown in the figure, SP has a negligible effect
on compressive strength at low water contents (less than 180 kg/m3) but
has a stochastic effect at high water contents. According to the current
study data, the random response is likely caused by the different
chemical properties of the high-range water reducers.
4.3. Deployment of the model
This study offers a free and easy-to-use graphical user interface (GUI)
to facilitate user interaction with the developed XG Boost model. A
sliding control system has been implemented in Python and Gradio [74],
allowing input values to be limited between minimum and maximum
(Table 4). The GUI developed assists in optimizing and predicting the
strength of concrete containing BFS and FA.
According to Fig. 18, this GUI consists of four main components,
9
M.I. Khan and Y.M. Abbas
Materials Today Communications 35 (2023) 105793
Fig. 17. PDPs of W and SP: (a) isoresponse contours, and (b) response surface.
Fig. 18. GUI for XG Boost model-based prediction of the compressive strength of concrete.
namely input features with slider controls, output results, explanations,
and some examples. The model outputs the concrete’s strength (in MPa),
as well as its class ("normal concrete" if it has a strength lower than
60 MPa, otherwise "high strength concrete"). A GUI explanation is based
on SHAP values [Fig. 8(b)], where the user can see how it is possible to
affect concrete’s compressive strength by varying the amount of the
input variables. The GUI also displays three examples of input variables
that can be chosen and submitted to view the model’s output.
Through the use of the GUI developed in this study, the normal and high
strength of concrete can be optimized in a shorter time period, at a lower
cost, with fewer efforts required. The study further provides 1D, 2D, and
3D PDPs. As a result of this study, the following implications were
drawn:
• Using the developed model, it is likely to be possible to predict the
strength of concrete that contains up to 360 kg/m3 BFS and 200 kg/
m3 FA at up to one year of age.
• As a starting point, a baseline model based on standardized hyper­
parameters was developed. The baseline model exhibited a tendency
toward overfitting, with R2 values of 0.996 and 0.919 for the training
and testing datasets, respectively.
• The hyperparameters of the model have been optimized using vector
multi-objective optimization to maximize the prediction capability
of the model. According to this study, the number of gradient boosted
5. Implications, recommendations, and outlook
This study developed an XG Boost model that accurately predicted
concrete compressive strength. A total of 1030 concrete mixes con­
taining OPC (type I), BFS, and FA were collected from 11 different
laboratories for concrete under normal moisture curing conditions.
Additionally, the study provides a simple and free GUI to support the
design of normal- and high-strength concrete containing BFS and FA.
10
Materials Today Communications 35 (2023) 105793
M.I. Khan and Y.M. Abbas
trees, boosting learning rate, and subsample ratio were the most
influential hyperparameters in the model performance.
• The optimized XG Boost model in the current investigation exhibited
a superior prediction performance with R2 of 0.992 and 0.949 for the
training and testing datasets. The optimized model results in pre­
dicted–target data for the test dataset of accuracy in the range of
[8]
[9]
[10]
± 85%.
• Based on Gini indexes, C, FA, water, and both aggregate types were
the most significant model parameters. An analysis of SHAP values
resulted in consistent findings.
• According to this study, the best BFS, FA, S, and SP contents for
concrete strength optimization were 100–200, 100–200, 600–800,
and 7–13 kg/m3, respectively.
• The SP has a negligible effect on concrete’s compressive strength at
low water contents (less than 180 kg/m3), but a stochastic effect at
high contents. The various chemical properties of high-range water
reducers may have resulted in the randomly generated response in
the current study.
[11]
[12]
[13]
[14]
[15]
[16]
In future studies, the scope would be expanded to include concrete
with a wide variety of SCMs (e.g., SF, metakaolin, rice husk ash,.etc.)
and fiber reinforcement systems. By using unseen data from the model,
the reliability of the model will be further examined.
[17]
[18]
[19]
[20]
[21]
CRediT authorship contribution statement
It is confirmed that neither the manuscript nor any parts of its con­
tent are currently under consideration or published in another journal.
All authors have approved the manuscript and agree with its submission
to Materials Today Communications.
[22]
[23]
Declaration of Competing Interest
[24]
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.
[25]
[26]
Data availability
[27]
[28]
Data will be made available on request.
Acknowledgments
[29]
The authors extend their appreciation to Researcher Supporting
Project number (RSPD2023R692), King Saud University, Riyadh,
Kingdom of Saudi Arabia.
[30]
References
[31]
[1] B. Cohen, Urbanization, city growth, and the new united nations development
agenda, Cornerstone 3 (2) (2015) 4–7.
[2] M. Rangelov, H. Dylla, B. Dobling, J. Gudimettla, N. Sivaneswaran, M. Praul,
Readily implementable strategies for reducing embodied environmental impacts of
concrete pavements in the United States, Transp. Res. Rec. (2022),
03611981221086934.
[3] L. Rodgers, Climate change: The massive CO2 emitter you may not know about,
2022. (Accessed August 27 2022).
[4] D. Ndahirwa, H. Zmamou, H. Lenormand, N. Leblanc, The role of supplementary
cementitious materials in hydration, durability and shrinkage of cement-based
materials, their environmental and economic benefits: a review, Clean. Mater.
100123 (2022).
[5] M.C. Juenger, R. Snellings, S.A. Bernal, Supplementary cementitious materials:
new sources, characterization, and performance insights, Cem. Concr. Res. 122
(2019) 257–273.
[6] Z. Guo, T. Jiang, J. Zhang, X. Kong, C. Chen, D.E. Lehman, Mechanical and
durability properties of sustainable self-compacting concrete with recycled
concrete aggregate and fly ash, slag and silica fume, Constr. Build. Mater. 231
(2020), 117115.
[7] Y.M. Abbas, L.A. Hussain, M.I. Khan, Constitutive compressive stress–strain
behavior of hybrid steel-PVA high-performance fiber-reinforced concrete, J. Mater.
[32]
[33]
[34]
[35]
[36]
[37]
[38]
11
Civ. Eng. 34 (1) (2022), 04021401, https://doi.org/10.1061/(ASCE)MT.19435533.0004041.
I.B. Topcu, M. Sarıdemir, Prediction of compressive strength of concrete containing
fly ash using artificial neural networks and fuzzy logic, Comput. Mater. Sci. 41 (3)
(2008) 305–311.
J.-S. Chou, C.-K. Chiu, M. Farfoura, I. Al-Taharwa, Optimizing the prediction
accuracy of concrete compressive strength based on a comparison of data-mining
techniques, J. Comput. Civ. Eng. 25 (3) (2011) 242–253.
M.S. Reddy, P. Dinakar, B.H. Rao, Mix design development of fly ash and ground
granulated blast furnace slag based geopolymer concrete, J. Build. Eng. 20 (2018)
712–722.
A.A. Aliabdo, M. Abd Elmoaty, M.A. Emam, Factors affecting the mechanical
properties of alkali activated ground granulated blast furnace slag concrete, Constr.
Build. Mater. 197 (2019) 339–355.
L. Baoju, X. Youjun, Z. Shiqiong, L. Jian, Some factors affecting early compressive
strength of steam-curing concrete with ultrafine fly ash, Cem. Concr. Res. 31 (10)
(2001) 1455–1458.
Wikipedia contributors, ML, 2022. (Accessed 27 August 2022.
J.J. Lee, D. Kim, S.K. Chang, C.F.M. Nocete, An improved application technique of
the adaptive probabilistic neural network for predicting concrete strength,
Comput. Mater. Sci. 44 (3) (2009) 988–998.
L. Van der Maaten, G. Hinton, Visualizing data using t-SNE, J. ML Res. 9 (11)
(2008).
V. Rathakrishnan, A.N. Ahmed, Predicting compressive strength of highperformance concrete with high volume ground granulated blast-furnace slag
replacement using boosting ML algorithms, Sci. Rep. 12 (1) (2022) 1–16.
T. Chen, C. Guestrin, Xgboost: A scalable tree boosting system, Proceedings of the
22nd acm sigkdd international conference on knowledge discovery and data
mining, 2016, pp. 785–794.
J. Friedman, T. Hastie, R. Tibshirani, Additive logistic regression: a statistical view
of boosting (with discussion and a rejoinder by the authors), Ann. Stat. 28 (2)
(2000) 337–407.
Wikipedia contributors, XGBoost, 2022. (Accessed August 29 2022).
Wikipedia contributors, Outlier, 2022. (Accessed February 18 2023).
H. Xu, H. Wang, C. Yuan, Q. Zhai, X. Tian, L. Wu, Y. Mi, Identifying diseases that
cause psychological trauma and social avoidance by GCN-Xgboost, BMC
Bioinforma. 21 (16) (2020) 1–16.
M. Kivrak, E. Guldogan, C. Colak, Prediction of death status on the course of
treatment in SARS-COV-2 patients with deep learning and ML methods, Comput.
Methods Prog. Biomed. 201 (2021), 105951.
P. Carmona, A. Dwekat, Z. Mardawi, No more black boxes! explaining the
predictions of a ML XGBoost classifier algorithm in business failure, Res. Int. Bus.
Financ. 61 (2022), 101649.
G.E. Hinton, S. Roweis, Stochastic neighbor embedding, Adv. Neural Inf. Process.
Syst. 15 (2002).
Y. Kawase, K. Mitarai, K. Fujii, Parametric t-stochastic neighbor embedding with
quantum neural network, Phys. Rev. Res. 4 (4) (2022), 043199, https://doi.org/
10.1103/PhysRevResearch.4.043199.
K. Ch’ng, N. Vazquez, E. Khatami, Unsupervised ML account of magnetic
transitions in the Hubbard model, Phys. Rev. E 97 (1) (2018), 013306.
L. Van Der Maaten, Accelerating t-SNE using tree-based algorithms, J. ML Res. 15
(1) (2014) 3221–3245.
A.C. Belkina, C.O. Ciccolella, R. Anno, R. Halpert, J. Spidlen, J.E. Snyder-Cappione,
Automated optimized parameters for T-distributed stochastic neighbor embedding
improve visualization and analysis of large datasets, Nat. Commun. 10 (1) (2019)
5415, https://doi.org/10.6084/m9.figshare.9927986.v1.
H. Liu, J. Yang, M. Ye, S.C. James, Z. Tang, J. Dong, T. Xing, Using t-distributed
Stochastic Neighbor Embedding (t-SNE) for cluster analysis and spatial zone
delineation of groundwater geochemistry data, J. Hydrol. 597 (2021), 126146,
https://doi.org/10.1016/j.jhydrol.2021.126146.
V. van Unen, T. Höllt, N. Pezzotti, N. Li, M.J. Reinders, E. Eisemann, F. Koning,
A. Vilanova, B.P. Lelieveldt, Visual analysis of mass cytometry data by hierarchical
stochastic neighbour embedding reveals rare cell types, Nat. Commun. 8 (1) (2017)
1740, https://doi.org/10.1038/s41467-017-01689-9.
E.M. Golafshani, A. Behnood, M. Arashpour, Predicting the compressive strength of
normal and high-performance concretes using ANN and ANFIS hybridized with
Grey Wolf Optimizer, Constr. Build. Mater. 232 (2020), 117266, https://doi.org/
10.1016/j.conbuildmat.2019.117266.
H. Sun, H.V. Burton, H. Huang, ML applications for building structural design and
performance assessment: state-of-the-art review, J. Build. Eng. 33 (2021), 101816.
J. Naranjo-Pérez, M. Infantes, J.F. Jiménez-Alonso, A. Sáez, A collaborative MLoptimization algorithm to improve the finite element model updating of civil
engineering structures, Eng. Struct. 225 (2020), 111327.
I.-C. Yeh, Modeling of strength of high-performance concrete using artificial neural
networks, Cem. Concr. Res. 28 (12) (1998) 1797–1808.
R. Gupta, M.A. Kewalramani, A. Goel, Prediction of concrete strength using neuralexpert system, J. Mater. Civ. Eng. 18 (3) (2006) 462–466.
H.I. Erdal, O. Karakurt, E. Namli, High performance concrete compressive strength
forecasting using ensemble models based on discrete wavelet transform, Eng. Appl.
Artif. Intell. 26 (4) (2013) 1246–1254.
E. Ghafari, M. Bandarabadi, H. Costa, E. Júlio, Prediction of fresh and hardened
state properties of UHPC: comparative study of statistical mixture design and an
artificial neural network model, J. Mater. Civ. Eng. 27 (11) (2015), 04015017.
N.-H. Nguyen, J. Abellán-García, S. Lee, E. Garcia-Castano, T.P. Vo, Efficient
estimating compressive strength of ultra-high performance concrete using XGBoost
M.I. Khan and Y.M. Abbas
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
Materials Today Communications 35 (2023) 105793
[56] W.S. Langley, G.G. Carette, V. Malhotra, Structural concrete incorporating high
volumes of ASTM class fly ash, Mater. J. 86 (5) (1989) 507–514.
[57] C. Lee, A study on dry shrinkage and creep property of HPC, National Taiwan Univ.
of Science and Technology. Taipei, Taiwan (1994).
[58] M. Lessard, O. Challal, P.-C. Aticin, Testing high-strength concrete compressive
strength, Mater. J. 90 (4) (1993) 303–307.
[59] F. Lin, Mixture proportion and quality of HPC, National Taiwan Univ. of Science
and Technology. Taipei, Taiwan (1994).
[60] H.L. Mo, A study on high performance concrete, National Taiwan Univ. of Science
and Technology, Taipei, Taiwan, 1995.
[61] Wikipedia contributors, Pearson correlation coefficient, 2022. (Accessed 28 August
2022.
[62] G.V. Rossum, J.F. Drake, Python reference manual, Centrum voor Wiskunde en
Informatica Amsterdam 1995.
[63] S.K. Babanajad, A.H. Gandomi, A.H. Alavi, New prediction models for concrete
ultimate strength under true-triaxial stress states: An evolutionary approach, Adv.
Eng. Softw. 110 (2017) 55–68.
[64] V. Pareto, Cours d′ économie politique, Librairie Droz1964.
[65] T. Hastie, R. Tibshirani, J.H. Friedman, J.H. Friedman, The elements of statistical
learning: data mining, inference, and prediction, Springer,, 2009.
[66] J. Gong, S. Chu, R.K. Mehta, A.J. McGaughey, XGBoost model for electrocaloric
temperature change prediction in ceramics, npj Comput. Mater. 8 (1) (2022) 1–10.
[67] H.-V.T. Mai, T.-A. Nguyen, H.-B. Ly, V.Q. Tran, Investigation of ann model
containing one hidden layer for predicting compressive strength of concrete with
blast-furnace slag and fly ash, Adv. Mater. Sci. Eng. 2021 (2021).
[68] İm Türkmen, R. Gül, C. Çel k, R. Dem rboğa, Determination by the Taguchi method
of optimum conditions for mechanical properties of high strength concrete with
admixtures of silica fume and blast furnace slag, Civ. Eng. Environ. Syst. 20 (2)
(2003) 105–118.
[69] G. Giaccio, R. Zerbino, Failure mechanism of concrete: combined effects of coarse
aggregates and strength level, Adv. Cem. Based Mater. 7 (2) (1998) 41–48.
[70] M. Khan, Y. Abbas, G. Fares, Review of high and ultrahigh performance
cementitious composites incorporating various combinations of fibers and
ultrafines, J. King Saud. Univ. -Eng. Sci. 29 (4) (2017) 339–347, https://doi.org/
10.1016/j.jksues.2017.03.006.
[71] E. Kearsley, P. Wainwright, The effect of high fly ash content on the compressive
strength of foamed concrete, Cem. Concr. Res. 31 (1) (2001) 105–112.
[72] X.-Y. Wang, K.-B. Park, Analysis of compressive strength development of concrete
containing high volume fly ash, Constr. Build. Mater. 98 (2015) 810–819.
[73] R. Siddique, M.I. Khan, Supplementary cementing materials, Springer Science &
Business Media, 2011.
[74] A. Abid, A. Abdalla, A. Abid, D. Khan, A. Alfozan, J. Zou, Gradio: Hassle-free
sharing and testing of ML models in the wild, arXiv preprint arXiv:1906.02569
(2019).
model, J. Build. Eng. 52 (2022), 104302, https://doi.org/10.1016/j.
jobe.2022.104302.
Z. Chao, Y. Dang, Y. Pan, F. Wang, M. Wang, J. Zhang, C. Yang, Prediction of the
shale gas permeability: A data mining approach, Geomech. Energy Environ.
(2023), 100435, https://doi.org/10.1016/j.gete.2023.100435.
W. Zhang, D. Shi, Z. Shen, W. Shao, L. Gan, Y. Yuan, P. Tang, S. Zhao, Y. Chen,
Reduction of the calcium leaching effect on the physical and mechanical properties
of concrete by adding chopped basalt fibers, Constr. Build. Mater. 365 (2023),
130080, https://doi.org/10.1016/j.conbuildmat.2022.130080.
Z. Chao, B. Gong, W. Yue, X. Xu, D. Shi, C. Yang, T. Hu, Experimental study on
stress-dependent gas permeability and porosity of artificially cracked cement
mortar, Constr. Build. Mater. 359 (2022), 129290, https://doi.org/10.1016/j.
conbuildmat.2022.129290.
H. Al Khalifah, P. Glover, P. Lorinczi, Permeability prediction and diagenesis in
tight carbonates using ML techniques, Mar. Pet. Geol. 112 (2020), 104096, https://
doi.org/10.1016/j.marpetgeo.2019.104096.
H. Wang, J. Shen, An improved model combining evolutionary algorithm and
neural networks for PV maximum power point tracking, IEEE Access 7 (2018)
2823–2827, https://doi.org/10.1109/ACCESS.2018.2881888.
S. Chengyi, S. Yan, X. Keming, Mind-evolution-based ML and applications,
Proceedings of the 3rd World Congress on Intelligent Control and Automation (Cat.
No. 00EX393), IEEE, 2000, pp. 112–117.
E. Alpaydin, Introduction to ML, MIT press,, 2020.
W.B. Chaabene, M. Flah, M.L. Nehdi, ML prediction of mechanical properties of
concrete: Critical review, Constr. Build. Mater. 260 (2020), 119889.
R. Siddique, P. Aggarwal, Y. Aggarwal, Prediction of compressive strength of selfcompacting concrete containing bottom ash using artificial neural networks, Adv.
Eng. Softw. 42 (10) (2011) 780–786.
Concrete Compressive Strength Data Set, in: P.I.-C. Yeh (Ed.) August 3, 2007.
I.C. Yeh, Modeling of strength of high-performance concrete using artificial neural
networks, Cement. 302 and Concrete Research 28(12) (1998).https://doi.
org/10.1016/S0008–8846(98)00165–3.
T.P. Chang, F.C. Chuang, H.C. Lin, A mix proportioning methodology for highperformance concrete, J. Chin. Inst. Eng. 19 (6) (1996) 645–655.
C. Chang, Research on the mix proportion of high flowing eugenic concrete, Chung
Hua Univ., Hsin Chu, Taiwan (1997).
F. Chung, Study on characteristic of coarse aggregate in high-performance
concrete, National Taiwan Univ. of Science and Technology. Taipei, Taiwan
(1995).
G. Giaccio, C. Rocco, D. Violini, J. Zappitelli, R. Zerbino, High-strength concretes
incorporating different coarse aggregates, Mater. J. 89 (3) (1992) 242–246.
O.E. Gjorv, P.J. Monteiro, P.K. Mehta, Effect of condensed silica fume on the steelconcrete bond, Mater. J. 87 (6) (1990) 573–580.
T. Hwang, Compressive strength of blast furnace slag concrete, National Chiao
Tung Univ. Hsin Chu, Taiwan (1966).
12
Download