You are on page 1of 14

Journal of Building Engineering 78 (2023) 107761

Contents lists available at ScienceDirect

Journal of Building Engineering


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

Experimental study of mixtures soil-industrial waste using simplex


design for application in paving
Klaus Henrique de Paula Rodrigues *, Taciano Oliveira da Silva,
Heraldo Nunes Pitanga, Leonardo Gonçalves Pedroti,
Mateus Henrique Ribeiro Rodrigues
Universidade Federal de Viçosa, Minas Gerais, Brazil

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

Keywords: The large amount of solid waste generated by industry is a concern for society. A possible solution
Optimization of mixtures is to incorporate these wastes into soils to improve their properties and contribute to sustainable
Design of experiments for mixtures development. The objective of this research was to conduct an experimental study on the
Electric arc furnace slag chemical stabilization of two samples of tropical soils using two industrial wastes, electric arc
Fly ash furnace slag (EAFS) and fly ash (FA), using a design of experiments (DoE) of mixtures in a simplex
Soil stabilization
centroid network. Seven mixtures were determined from the DoE for each soil sample. Using
specimens of these mixes prepared at optimum moisture content, after a curing period of 7 days in
a humid chamber, the properties were evaluated: unconfined compressive strength (UCS), Cali­
fornia Bearing Ratio (CBR), and resilient modulus (RM). An improvement in the evaluated me­
chanical properties was observed with the addition of waste, especially with the addition of EAFS.
The statistical software Minitab 18 allowed to obtain response surfaces for each mechanical
property studied as a function of the pseudo-components of the mixtures. To determine the
proportion of components that optimized the soil-waste mixtures, the desirability function was
used, which resulted in the proportions of 80% soil, 20% EAFS, and 0% FA for both soil samples.
Based on SEM and XRD tests, it was found that calcium silicate hydrate (C–S–H), calcium
aluminate silicate hydrate (C-A-S-H), and calcium aluminate hydrate (C-A-H) were formed, which
were responsible for improving the mechanical strength. The use of industrial waste to stabilize
tropical soils with low bearing capacity proved to be a technically and environmentally viable
alternative.

1. Introduction
In Brazil, direct steel wastes, excluding wastes from canteens, administration, civil engineering and others, amounted to about 20.1
million tons in 2017, of which 27% was steel slag [1]. Steel Slag is a partially crystalline byproduct of steelmaking in which metallic
and non-metallic components are separated from the raw ore [2]. Various types of slags generated from different steelmaking processes
have already been used in researches as soil stabilizers, including blast furnace slag [3–5], basic oxygen furnace slag [6,7], electric arc
furnace slag (EAFS) [8], and ladle furnace slag [9–11]. Global production of fly ash (FA) is estimated at about 363 million tons per year

* Corresponding author. Universidade Federal de Viçosa, Viçosa, Minas Gerais, Brazil, 36576-028.
E-mail addresses: klaus@ufv.br (K.H.P. Rodrigues), taciano.silva@ufv.br (T.O. da Silva), heraldo.pitanga@ufv.br (H.N. Pitanga), leonardo.pedroti@ufv.br
(L.G. Pedroti), mateus.ribeiro@ufv.br (M.H.R. Rodrigues).

https://doi.org/10.1016/j.jobe.2023.107761
Received 18 April 2023; Received in revised form 21 August 2023; Accepted 9 September 2023
Available online 14 September 2023
2352-7102/© 2023 Elsevier Ltd. All rights reserved.
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

[12], but only a quarter of the world’s production is used. Unused FA is usually disposed of in landfills, posing a risk of air and
groundwater contamination [13]. This scenario, in which steel waste is generated in such large quantities, shows the need to make its
use profitable [14], construction being one of the economic sectors in which this waste can be used on a large scale [15].
Considering the need to reduce the consumption of natural resources used as raw materials in construction, some studies have been
carried out to investigate the technical aspects related to the use of EAFS and/or FA as alternative materials for the partial replacement
of cement [16], as an aggregate in mortars [17], in the composition of bricks [18,19], concrete [12,20], pavers [21], embankments
[22], and asphalt mixtures [23–25], and in soil stabilization as a replacement for cement and lime [9,26–41]. Research on the use of
EAFS, FA, and other wastes for soil stabilization has generally been conducted using intuitive and empirical measurement methods
[8–10,26,27,29–32,42].
The lack of criteria for the use of certain types of wastes in a particular application is one of the main factors limiting their use on a
large scale [43]. To date, there is no known method in the literature for the dosage of soil-waste mixtures based on criteria that can
optimize the contributions of each component and, consequently, the engineering properties of the final product, in order to meet the
requirements of projects that require the use of compacted soils. In this context, and with the aim of filling this gap, the present
technical-scientific proposal, based on the statistical model of dosage through a design of experiments in a simplex-centroid network,
which takes into account that the properties of interest depend on the proportions of their components, fits the optimization (maxi­
mization or minimization) of one or more desired properties [44], giving greater statistical confidence to the results and, moreover,
reducing the number of experiments [45,46].
According to Yıldırım, Karacasu and Okur [47], an experimental design consists of selecting points representing the mixtures to be
analyzed within the experimental space. The selected mixtures are then tested for the properties of interest to obtain polynomial
functions describing the responses of the mixtures with respect to these properties. The desirability function can then be applied to
determine the best dosage of components to optimize the desired properties.
The protocol for optimizing properties of interest by design of experiments for mixes has been studied extensively in the con­
struction sector, for example in the manufacture of ceramic blocks and artifacts [46], in soil-cement blocks [48], in paints [49], in
mortars [50,51], and in concrete [52–54]. In the field of soil engineering, the design of experiments method was applied on a smaller
scale to mixtures [55–58].
Considering the need to enable the large-scale use of industrial wastes as chemical stabilizers for compacted soils, the main
objective of this research project was to propose a protocol for the dosage of soil-industrial waste mixtures based on the statistical
model of mixture design of experiments in a simplex-centroid network, to determine the optimal dosage of mixtures of tropical soils,
EAFS, and FA considering the optimization of mechanical properties of the Unconfined Compressive Strength (UCS), Resilient Modulus
(RM), and California Bearing Ratio (CBR). As secondary objectives related to the main one, we intend: (i) to analyze the influence of
EAFS and FA on the mechanical properties studied; (ii) to interpret the results relevant to the above properties based on the physical,
chemical, mineralogical and microstructural properties of the constituents.

2. Materials and methods


2.1. Materials
Soil samples S1 and S2 used in this study are from deposits located in the municipality of Viçosa, Minas Gerais, Brazil, and represent
a clayey and a sandy soil sample, respectively. A company located in the Alto Paraopeba region, Minas Gerais, Brazil, involved in metal
pipe manufacturing, provided the Electric Arc Furnace Slag (EAFS) sample. The Fly Ash (FA) sample was provided by a company
located in the municipality of Capivari de Baixo, Santa Catarina, Brazil, which is engaged in the processing and marketing of
pozzolanic ash, a material derived from the combustion of mineral coal. The FA used was classified as class C, according to NBR 12653
[59], and as class F, according to ASTM C618-22 [60].

2.2. Geotechnical and physical characterization


Geotechnical characterization of the soil samples was performed using the following tests: i) particle size distribution, according to

Table 2.1
Results of the characterization and classification of the soil samples.

Parameter S1 S2

% clay (φ < 0.002 mm) 61 5


% silt (0.002 mm < φ < 0.06 mm) 14 24
% sand (0.06 mm < φ < 2 mm) 25 68
% gravel (2 mm < φ < 60 mm) 0 3
LL (%) 78 36
PL (%) 43 18
PI (%) 35 18
ρs (g/cm3) 2.869 2.657
γd,max (kN/m3) 13.95 15.57
wopt (%) 31.03 15.00
TRB A-7-5 (20) A-2-6 (1)
USC MH SC
MCT LA’ NA’

2
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

ASTM D6913-17 and ASTM D7928-21 [61,62]; ii) liquid limit (LL) and plastic limit (PL), according to ASTM D4318-17 [63]; iv)
specific gravity of solids, according to ASTM D854-14 [64]; and v) compaction at standard Proctor energy, according to ASTM D698-12
[65] to determine the maximum dry unit weight (γd,max) and optimum moisture content (wopt) of the compacted soil samples. The
results of these tests are shown in Table 2.1.
To increase the specific surface area of the EAFS sample, it was ground in a Los Angeles machine for two to 3 h, using the portion
that passed through a 0.6 mm sieve (30 meshes), a similar size to that used in previous studies [66–68], to achieve a value close to
1800 cm2/g specific surface area used for steel slag in the study conducted by Sheen, Le and Lam [69].
For the EAFS and the FA samples, which were to be physically characterized, the following tests were performed following the
technical standards for Portland cement: i) fineness by the 75 μm sieve [70], ii) fineness by the air permeability method (specific
surface area - Blaine method) [71], iii) specific gravity [72], and iv) particle size analysis determined by laser diffraction [73].
The EAFS sample had a sieve fineness index of 24.88%, which is a high value for cementitious materials in Brazil [35]. The specific
surface area by the Blaine method was 0.20 m2/g, a value close to that required for Portland cement, which must be greater than 0.24
m2/g [29]. The result for specific gravity was 3.7 g/cm3. The sample FA had a fineness index of 23.71%, a specific surface area of
1788.24 cm2/g and a specific gravity of 2.08 g/cm3.
Fig. 2.1 shows the grain size distribution curves of EAFS, FA, and soil samples S1 and S2 classified by the MCT (Miniature,
Compacted, Tropical) method, according to DNER-CLA 259 [74] and by the TRB (Transportation Research Board), according to
AASHTO M 145–91 [75] and USC (Unified Soil Classification) geotechnical soil classification systems, according to ASTM D2487-17
[76] (Table 2.1).

2.3. Chemical and mineralogical characterizations


The soil, EAFS, and FA samples were subjected to X-ray fluorescence (XRF) analysis. The results are shown in Table 2.2 along with
the loss on ignition test, according to NBR NM 18 [77]. X-ray diffraction (XRD) was used to determine the mineralogical composition of
the samples of the analyzed materials. For this purpose, theta-2theta measurements were performed in the range of 5–80◦ with a step of
0.05◦ and 1 s per step. The results of the XRD analysis are shown in Fig. 2.2a, b, 2.2c, and 2.2d for soil samples S1, S2, EAFS, and FA,
respectively.
As for the XRF tests, the predominance of SiO2, Al2O3 and Fe2O3 compounds was found in the S1 soil sample, which were also found
in the S2 soil sample. A large amount of CaO and Fe2O3 was detected in the EAFS sample. The hydraulicity index value (SiO2
+Al2O3+Fe2O3)/(CaO + MgO) was 1.49, classifying the EAFS as strongly hydraulic [78], indicating that the material has binding
potential. A significant amount of SiO2 was found in the FA sample. These results are consistent with those found in the literature for
most EAFS and FA samples [9,30,79,80]. In the XRD studies, for the EAFS sample, iron (iron ore), calcium (portlandite and larnite),
and silicon (larnite) were found to be the main constituents, which is consistent with the results of the XRF studies. In the sample of FA,
the high intensity of the chemical compound mullite (Al2O3⋅SiO2) was found to be consistent with the XRF results.
The tests for determination of free lime, according to NBR NM 13 [81] and insoluble residue NBR NM 15 [82] were also performed
for the EAFS sample and gave values of 0.74% and 7.94%, respectively. The result obtained for free lime, which is less than 1%, allows
the use of the material in road construction due to its low expansivity [83].

2.4. Methods
2.4.1. Environmental characterization of EAFS and FA samples
The environmental characterization of the EAFS and FA samples was performed according to the NBR 10004 technical standard
[84]. To determine the classification of each industrial waste, leaching and solubilization tests were performed in accordance with
technical standards NBR 10005 [85] and NBR 10006 [86], respectively.

2.4.2. The simplex-centroid design for experiments with mixtures


In order to evaluate the influence of EAFS and FA on the chemical stabilization of the studied soil samples, the statistical method of
experimental design in a simplex-centroid network was used. Minitab 18 statistical software was used to perform this procedure. A
flow chart showing the procedure for determining the optimal dosage is shown in Fig. 2.3.
2.4.2.1. Design of experiments. The wastes (EAFS and FA) were used in proportion to the dry mass of the soil-waste mixtures (soil-
EAFS-FA) in a maximum percentage of 20%, with the EAFS content set between 5 and 20% and the FA content between 0 and 15%.
These limits were established based on previous studies using soil mixtures with EAFS and FA [10,30,87]. For soil samples, the upper
limit was 95% and the lower limit was 80%, the latter value being defined on the basis of the sum of all components, which must be

Fig. 2.1. Grain size distribution curves of the EAFS, FA, and soil samples S1 and S2.

3
­
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Table 2.2
Chemical composition of the EAFS, FA, and soil samples S1 and S2.

Material SiO2 Al2O3 Fe2O3 CaO MgO K2 O Na2O TiO2 SO3 Cl Other LOI*
(%) (%) (%) (%) (%) (%) (%) (%) (%) (%) compounds (%)

S1 27.11 24.52 12.47 0.02 1.28 0.07 1.55 1.53 0.04 0.27 31.14 11.49
S2 36.27 24.31 1.92 0.17 1.41 1.30 2.40 0.18 0.03 0.22 31.79 4.19
EAFS 21.35 5.17 24.50 33.15 1.01 0.30 0.00 0.64 0.18 0.32 13.38 0.19
FA 57.50 22.69 4.00 1.66 1.03 2.67 1.33 1.06 0.35 0.20 7.51 1.50

*Loss on Ignition.

Fig. 2.2. Difractograms of the soil samples S1 (a), S2 (b), EAFS (c), and FA (d).

equal to 100% [88].


Using the design of experiments method in simplex-centroid for three components, 7 points were generated within an equilateral
triangle forming a space called simplex. The equilateral triangle generated with the components in the pure state and the triangle
generated with the pseudo components are shown in Fig. 2.4.
The percentages of each mixture determined using Minitab 18 software for the two soil samples studied (S1 and S2), taking into
account the lower and upper limits of each component, are shown in Table 2.3. The percentages of the materials were determined with
respect to the total dry mass of the mixture.
To perform the laboratory tests used to optimize the experimental design, the compaction curves in standard Proctor energy were
obtained for each mixture to determine the maximum dry density (γd,max) and the optimum moisture content (wopt) [65].
To determine the mechanical properties of each mixture studied, three specimens were molded at optimum moisture content and
subjected to a sealed curing process (in PVC film) for seven days in a humid chamber to evaluate cementation properties [89]. The
coefficient of variation was 10% for all properties evaluated.
The mechanical properties used to evaluate the experimental design of this study were the Unconfined Compressive Strength
(UCS), the Resilient Modulus (RM) and California Bearing Ratio (CBR index) of the compacted specimens of the mixtures analyzed,
according to the respective technical standards NBR 12025 [90], DNIT ME 134 [91] and NBR 9895 [92]. These mechanical properties

4
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Fig. 2.3. Flowchart of the simplex-centroid method for optimizing the experimental design of mixtures.

Fig. 2.4. Mixture design of experiments.

Table 2.3
Soil-waste mixtures defined in the experimental design.

Mixture Soil (%) EAFS (%) FA (%)

M1 95 5 0
M2 87.5 12.5 0
M3 80 20 0
M4 80 12.5 7.5
M5 80 5 15
M6 87.5 5 7.5
M7 85 10 5

5
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

are directly related to the criteria for the selection of materials for the composition of structural layers of road pavements, as rec­
ommended in the technical standard NBR 12253 [93] and DNIT [94].
The analysis of the results was performed in the form of pseudo components, and the method used to fit the mathematical models
used was the stepwise method, in which the non-significant terms were automatically removed. The significance level assumed in the
analyzes was 10%. Therefore, the terms in the statistical analysis that had a p-value greater than 0.10 were removed.
The special cubic model, used in the simplex-centroid is presented on equation (2.1).

n ∑∑
n ∑∑
n ∑
y= bi xi + bij xi xj + bijk xi xj xk (2.1)
i=1 i<j i<j<k

(Where: bi, bij and bijk are the constants of the mathematical model and xi, xj and xk is the proportions of components i, j and k,
respectively).
Expanding equation (2.1), we obtain equation (2.2), which represents the special cubic model used in the simplex-centroid method
for a mixture of three components. The constant b1 is the value estimated by the mathematical model when the proportion of
component x1 in the mixture is maximal. To assess which component(s) have the greatest influence on the increase or decrease of a
particular property, one can calculate the arithmetic mean between the constants b1, b2 and b3. The constant whose value is farthest
from the mean indicates that the material it represents has the greatest influence on the change in the analyzed property, either to
increase the value of the property (constant above the mean) or to decrease it (value below the mean). The constant b12 represents the
interaction between the components x1 and x2, if its value is positive, there is synergism, indicating an increase in the value of the
property; if its value is negative, there is antagonism, indicating a decrease in the value of the property. The same explanation can be
given for the constant b123 [105].
y = b1 x1 + b2 x2 + b3 x3 + b12 x1 x2 + b13 x1 x3 + b23 x2 x3 + b123 x1 x2 x3 (2.2)
2.4.2.2. Optimization through the desirability function. The average values of the results of the mechanical tests used to optimize the
experimental design promoted the response surfaces determined for the base course, the subbase course and subgrade reinforcement
layers in the pavement, which allowed the identification of the maximization region for each mechanical property analyzed. To
determine the optimal mixture within this region, optimization was performed using the desirability function of the Minitab 18
software.
To verify the suitability of soil-waste mixtures for the base course, the minimum value of 690 kPa was adopted for the UCS ac­
cording to Gautreau, Zhang and Wu [95] and T 220 [96]. For RM, the minimum values from Ref. [97] were adopted. They are 82.74
MPa for fine soils (soil sample S1) and 165.47 MPa for granular soils (soil sample S2). For the CBR index, a minimum value of 60% was
adopted in accordance with DNIT [94], which is specified for base courses of flexible pavements intended to withstand a number of
loads on the standard axis less than 5 × 106.
In order for the soil-waste mixtures to be used in the subbase course, a minimum acceptable UCS value of 690 kPa was assumed
according to Gautreau, Zhang and Wu [95]. According to Ref. [97], the minimum acceptable values for clayey and sandy soils for the
subbase course are 82.74 MPa and 165.47 MPa, respectively. For the CBR index, a minimum value of 20% was assumed according to
DNIT [94].
Minimum values for UCS, RM, and CBR index of 345 kPa, 68.95 kPa, and 2%, respectively, were assumed for the subgrade
reinforcement layer according to Gautreau, Zhang and Wu [95], AASHTO [97], and DNIT [94].
With the minimum requirements to be achieved for each structural layer, the optimum dosage of soil-waste mixtures could be
determined considering the two soil samples.
The technique of scanning electron microscopy (SEM) was used for the morphological analysis of the optimal mixture of each soil
sample. The mineralogical composition of each optimal mixture was determined by X-ray diffraction (XRD) analysis.

3. Results and discussions


3.1. Environmental characterization
Concentrations of chemical elements in the leached extracts from the EAFS and FA samples were below the maximum levels
specified in NBR 10004 [84] for solid wastes considered environmentally hazardous.
In the solubilized extract of the ground EAFS sample, the elements iron and aluminum had concentrations higher than the
maximum values for a solid waste to be considered inert, which are also recommended as maximum values in order not to endanger the
potability of water according to the FUNASA [98] regulation. This means that the EAFS sample should not be in direct contact with the
water supply. In the solubilized extract of the FA, the concentration of the chemical element chromium was found above the maximum
level that prevents direct contact with the water supply, as the concentration is above the limit of 0.05 mg/L.
In view of the results presented, the industrial wastes EAFS and FA are classified as solid wastes of class II A - non-hazardous, non-
inert. It was determined that these industrial wastes have a low impact on the environment and are not considered aggressive to the
environment and human health when used in structural layers of road pavements as soil stabilizers, except in cases where there is
direct contact with water resources.

6
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

3.2. Simplex-centroid design


3.2.1. Mechanical properties analyzed
Fig. 3.1a shows the results of the UCS tests for the soil-waste mixtures. It was found that the largest increases in UCS values
compared to the values obtained for the soil samples in the natural state occurred for mixtures M3 and M4 for soil sample S1 and M3 for
soil sample S2. For soil samples S1 and S2, there was a tendency for the UCS to increase with the increase in EAFS by mixtures M1, M2,
and M3. Later, the results of the SEM tests are presented, which showed that cementitious compounds were formed in mixture M3 for
both soil samples, explaining the increase in UCS values. The same observation was also confirmed in the studies of Ismail, Awad and
Mwafy [30] and Al-Homidy et al. [99]. The UCS values in mixtures containing FA were higher than the value of the soil sample in its
natural state. Bose [100,101] stated that the increase in UCS value in the presence of FA was due to pozzolanic reactions producing
cementitious compounds.
Fig. 3.1b shows the resilient modulus (RM) results for the soil-waste mixtures of soil samples S1 and S2. All mixtures for the two soil
samples showed better performance compared to the soil samples in the natural state (M0 mixture). For the M1, M2 and M3 mixtures, it
was observed that the addition of EAFS improved the response of soil samples S1 and S2 under cyclic loading. Mixture M3 was
analyzed by scanning electron microscopy for both soil samples. This showed the formation of cementitious compounds, which ex­
plains the improvement in the performance of the mixtures with EAFS compared to the soil samples in the natural state.
Fig. 3.1c shows the results of the CBR test. It was found that the mixtures M4 and M3 had the highest values of CBR index for the soil
samples S1 and S2, respectively. The increases compared to mixture M0 were 132.14% for soil sample S1 and 70.73% for soil sample
S2. The mixtures for soil sample S1, with the exception of M1, achieved the minimum CBR value of 20% recommended for sub-base
courses of flexible pavements according to DNIT [94]. There was a tendency for the CBR index values to increase with increasing EAFS
content, which was also observed by other authors [26,27,31,102]. The mixtures containing FA showed higher CBR index values
compared to the soil sample in its natural state. This increase could be due to the formation of cementitious compounds by pozzolanic
reactions [100,103].
Fig. 3.1d shows the CBR expansion data for soil samples S1 and S2 and their mixtures with EAFS and FA. For mixtures M1, M2, and
M3, it was observed for soil samples S1 and S2 that the increase in EAFS content resulted in a decrease in CBR expansion values. Some
studies conducted on mixtures between samples of clayey soils and EAFS reported that the decrease in expansion with the addition of
EAFS was due to the occurrence of cation exchange between the ions present in the two materials, resulting in flocculation of the
particles and a change in their structure [104]. Other authors have also observed a decrease in the expansion of soil samples with an
increase in EAFS content [27,28,30].
Taking into account all the mechanical properties analyzed, the mixtures with the highest EAFS (M3) values were those that
obtained the best results.

3.2.2. Response surfaces for UCS


The response surfaces of the UCS test found for the soil-waste mixtures with respect to the soil samples S1 and S2 are shown in
Fig. 3.2, and the equations representing these surfaces are listed in Table 3.1. The independent variables were the percentages of soil
(S), EAFS, and FA.
It was found that the mathematical models fitted by the stepwise method for the response surfaces were linear models. The
equations presented showed that EAFS was the only component of the mixtures that contributed to the increase in UCS values, since it
was the only element of the mixture whose coefficient was above the average of the 3 components for the soil-waste mixtures [105],
considering the samples of soils S1 and S2. This significant contribution of EAFS is due to the formation of cementitious compounds

Fig. 3.1. Results of the analyzed properties of the tested mixtures and soil samples: (a) UCS, (b) RM, (c) CBR index, and (d) CBR expansion.

7
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Fig. 3.2. Response surfaces for the UCS of soil-waste mixtures: (a) soil sample S1, (b) soil sample S2.

Table 3.1
Equations referring to UCS response surfaces.

Soil sample Regression equation R2

S1 278.1S + 370.0EAFS+304.8FA 78.25


S2 126.0S + 188.4EAFS+146.0FA 68.79

such as calcium silicate hydrate (C–S–H), calcium aluminate silicate hydrate (C-A-S-H), and calcium aluminate hydrate (C-A-H),
formed during the sealed curing phase in a humid chamber, and to a lesser extent to structural changes due to cation exchange, higher
values of dry unit weight and lower values of optimum moisture content, which is consistent with the explanations in the
technical-scientific literature [26,106,107].
The coefficient of determination (R2) is a goodness-of-fit measure for linear regression models. This statistic indicates the per­
centage of the variance in the dependent variable that the independent variables explain collectively. The R2 value for the regression
equation for the S1 soil sample (about 80%) was good, considering that the soil and especially the EAFS are very heterogeneous
materials. Given this variability of materials, the UCS values of a particular mixture selected within the simplex space can be predicted
with good accuracy. The R2 value for the regression equation for soil sample S2 (about 70%) indicates that the fit was only reasonable
and it is not possible to predict the estimated UCS values with high accuracy.

3.2.3. Response surfaces for RM


The response surfaces of the RM test determined by the statistical models for the soil-waste mixtures of soil samples S1 and S2 are
shown in Fig. 3.3a e 3.3b, and the equations representing these surfaces are listed in Table 3.2.
The response surfaces for the RM were linear for soil sample S1 and nonlinear for soil sample S2. Looking at the linear terms for the
S1 soil sample, the EAFS is the component most responsible for the increase in RM values, while the FA sample is the component
causing the largest decrease in RM values.
For the S2 soil sample, considering the linear terms, it was found that the EAFS is the component that contributes most to the
increase of the RM values and that the soil is the one that contributes most to the decrease of this property. As for the interaction
between the components, there was antagonism between the soil and the EAFS, and between the EAFS and FA.
The R2 value for the regression equation for soil sample S1 (about 90%) showed that the fit was very good and the estimated RM
values could be predicted with high accuracy. The R2 value for the regression equation for soil sample S2 (nearly 100%) showed that
the fit was excellent.

Fig. 3.3. Response surfaces for RM of soil-waste mixtures: (a) Soil sample S1, (b) Soil sample S2.

8
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Table 3.2
Equations referring to RM response surfaces.

Soil sample Regression equation R2

S1 63.24S + 84.49EAFS+42.84FA 87.92


S2 28.71S + 38.62EAFS+31.718FA-20.45S.EAFS-12.91EAFS.FA 97.78

3.2.4. Response surfaces for CBR


The response surfaces for soil-waste mixtures of soil samples S1 and S2, considering the CBR, are shown in Fig. 3.4, and the
corresponding mathematical models in Table 3.3.
The mathematical model of the soil-waste mixtures from the S1 soil sample shows two terms with interactions between the pseudo
components, indicating that the interactions between soil and FA and between EAFS and FA cause an increase in RM (synergy).
However, considering only the coefficients without interaction, it is clear that EAFS is the element that contributes most to the increase
of the CBR index.
For soil-waste mixtures from soil sample S2, the response surface shows a tendency to increase the CBR index with the increase of
EAFS, in addition to the indication that soil and FA are responsible for the decrease of the property.
The R2 value for the regression equation for soil sample S1 (nearly 100%) showed that the fit was excellent and the estimated CBR
values could be predicted with high accuracy. The R2 value for the regression equation for soil sample S2 (about 75%) indicates that
the fit was only good and it is not possible to predict the estimated UCS values with high accuracy.

3.2.5. Determination of optimal composition


The desirability function through the Minitab 18 software showed that the mixture of 80% soil sample, 20% EAFS and 0% FA is the
optimal composition for all considered scenarios of structural layers and for soil samples S1 and S2.
Table 3.4 shows the individual and overall desirability values obtained for the optimal mixtures with respect to the considered
structural layers and soil samples S1 and S2. The individual desirability values are calculated by the relationship between the value
obtained for each property analyzed by the mathematical model of each response surface and the minimum acceptable value for the
material to be used for each structural layer of the pavement. The overall desirability values are determined by a weighted relationship
between the individual desirability values.
The optimal mixture of soil sample S1 considering the subgrade reinforcement layer had individual desirability values of 1.00 for
the UCS, RM, and CBR. Thus, the mixture met all the minimum acceptable values according to the technical standards, which can also
be observed by the value of 1 in the overall desirability.
Although the minimum values were not achieved, the optimal mixture of 80% soil, 20% EAFS, and 0% FA for soil sample S2 was the
one that had the highest value for overall desirability for all structural layers tested. The highest value for overall desirability was
obtained for the subgrade reinforcement layer, showing the better suitability of the optimal mixture for the construction of the less
noble layers of the pavement.

3.2.6. Characterization of optimal mixtures


Additional specimens of the optimal mixture were prepared to perform the mechanical tests of UCS, RM, and CBR with the optimal
parameters of the compaction test (γd,max and wopt) cured sealed in a humid chamber for 7 days to confirm the data provided by the
optimization performed with the desirability function. Table 3.5 shows the experimental results for UCS, RM, and CBR corresponding
to these specimens, the estimated results of these properties by the mathematical models created in Minitab 18, and the variation of the
experimental results from the estimated results.
The experimental data obtained with the optimal mixture proportions showed that the prediction of the regression equations for
soil sample 1 was very good, with the maximum variation between the predicted value and the obtained value being 12.37% for the
CBR property. For the optimal mixture with soil sample S2, the predicted values were not so close to the experimentally determined
values, especially for the CBR values. Nevertheless, the prediction of the values was considered acceptable because due to the vari­
ability of soils, this variation of CBR values from 8.55% to 11.67% can occur in the laboratory. Since the mixture contains EAFS in its
composition, the variability between specimens may be even greater. Considering the variability of the studied materials and the

Fig. 3.4. CBR response surfaces of soil-waste mixtures: (a) soil sample S1, (b) soil sample S2.

9
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Table 3.3
Equations referring to CBR response surfaces.

Soil sample Regression equation R2

S1 16.79S + 24.79EAFS+21.56FA+18.3S.FA+10.3EAFS.FA 98.08


S2 7.15S + 14.15EAFS+9.55FA 74.53

Table 3.4
Desirability values related to the optimal mixtures for the analyzed properties.

Mechanical property S1 S2

Subgrade reinforcement Sub-base Base Subgrade reinforcement Sub-base Base

UCS (kPa) 1.00 0.54 0.53 0.55 0.27 0.27


RM (MPa) 1.00 1.00 1.00 0.56 0.23 0.23
CBR (%) 1.00 1.00 0.42 1.00 0.71 0.24
Overall desirability 1.00 0.81 0.61 0.67 0.36 0.25

Table 3.5
Comparison between the values estimated and obtained experimentally for the specimens.

Properties S1 S2

Estimated value Experimental average Variation (%) Estimated value Experimental average Variation (%)

UCS (kPa) 296.48 283.67 − 4.32 138.48 171.15 23.59


RM (MPa) 67.49 74.63 10.58 27.42 35.98 31.23
CBR (%) 18.39 20.67 12.37 8.55 11.67 36.45

experimental values obtained in both soils, experimental design of mixtures may be a good tool for prediction and optimization of
mixtures of soils and industrial wastes.

3.2.7. Mineralogical and microstructural influence on the strength behavior of soil-waste mixtures
The mineralogical and morphological compositions of the optimal mixtures of soil samples stabilized with 20% EAFS were
determined to investigate the role of EAFS in increasing the values of UCS, RM, and CBR.
3.2.7.1. Optimal mixture stabilization for soil sample S1. The images obtained by SEM of the optimal mixture of the soil sample S1
shown in Fig. 3.5a indicate the formation of structures in the form of a cementitious gel, which produces a denser structure than the
structure of the soil sample in the natural state and increases the material strength, as shown by the results of UCS, RM, and CBR. The
cementitious gel may consist of C–S–H, C-A-S-H, and C-A-H, which cover the surface of the soil sample particles, fill the pores, and
enhance agglomeration between the particles. These compounds were probably formed by chemical reactions between the calcium
oxide in EAFS and the silica and alumina components in soil sample S1. Similar structures of these compounds were also found by other
authors [26,106,107]. The XRD analysis of the optimal mixture corresponding to soil sample S1 is shown in Fig. 3.5b.
The improvement in the evaluated mechanical properties can also be attributed to the fact that the electronegativity of the chemical
element Ca (1.0 Pauling) present in the EAFS facilitates the formation of agglomerative compounds with the silica (1.9 Pauling)
present in the soil sample S1 due to a high electronegativity difference [26].
3.2.7.2. Optimal mixture stabilization for soil sample S2. The images obtained by SEM of the optimal mixture related to soil sample S2
(Fig. 3.6) show the formation of C–S–H, C-A-S-H, and C-A-H structures resulting from chemical reactions between the calcium oxide
present in the EAFS and the silica and alumina components present in the soil sample S2. The XRD data (Fig. 3.6) show that the
diffractogram of the optimal mixture has an iron oxide peak due to the presence of iron oxide in the EAFS, which in combination with
aluminum oxides causes an aggregation effect of the particles that reduces the pores of the material [107]. According to Al-Amoudi
et al. [26], the large difference in electronegativity between the element calcium (1.0 Pauling) and the elements iron (1.8 Pauling)
and silicon (1.9 Pauling) favors the formation of strong bonds between these elements. In this scenario, the Ca atom is located between
Fe and Si, resulting in a bonding reaction between silica, calcium hydroxide and iron oxide.

4. Conclusions
The main objective of this research project was to propose a protocol for the dosage of soil-industrial waste mixtures based on the
statistical model of mixture design of experiments in a simplex-centroid network to determine the optimal dosage of mixtures of
tropical soils, EAFS and FA take into account the optimization of the mechanical properties Unconfined Compressive Strength (UCS),
Resilient Modulus (RM) and California Bearing Ratio (CBR).
The design of experiments of mixtures was useful to perform the optimization of the analyzed mechanical properties, based on the
proportions of the components of the mixtures, with the aim of achieving the values recommended in the literature.
The statistical software Minitab 18 allowed to obtain response surfaces for each mechanical property studied as a function of the
pseudo-components of the mixtures. To determine the proportion of components that optimized the soil-waste mixtures, the

10
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

Fig. 3.5. Mineralogical and microstructural analyses of the optimal mixture for soil sample S1: (a) Images obtained by SEM ×20,000 and (b) Diffractogram obtained
by XRD.

Fig. 3.6. Mineralogical and microstructural analyses of the optimal mixture for soil sample S2: (a) Images obtained by SEM ×10,000 and (b) Diffractogram obtained
by XRD.

desirability function was used, which resulted in the proportions of 80% soil, 20% EAFS, and 0% FA for both soil samples. After
determining the optimal mixture, test specimens were prepared to compare the estimated and experimentally determined values for
the properties evaluated. This comparison showed that the use design of experiments of mixtures is a good tool for predicting and
optimizing the characteristics of stabilized soil.
The addition of EAFS has a positive effect on UCS, RM and CBR, mainly due to the formation of C–S–H, C-A-S-H, and C-A-H
structures for soil samples S1 and S2, respectively. The presence of wustite and iron oxide in the optimal mixtures of soil samples S1
and S2, respectively, was also observed, which contributed to fill the pores and increase the values of the analyzed mechanical
properties.
The addition of FA to the soil samples had a negative effect on all simulations with the two analyzed soil samples, despite the
improvement of the analyzed mechanical properties compared to the soil in its natural state.
It was observed that the CBR expansion decreased in the optimal mixtures of the studied soil samples compared to the soil samples
in its natural state, which is due to the cation exchange that changes the structure of the material and causes a higher degree of
flocculation.
Based on SEM and XRD tests, it was found that calcium silicate hydrate (C–S–H), calcium aluminate silicate hydrate (C-A-S-H), and
calcium aluminate hydrate (C-A-H) were formed, which were responsible for improving the mechanical strength.
The use of industrial wastes in the stabilization of tropical soils with low bearing capacity proved to be a technically and envi­
ronmentally viable alternative, as it improved the evaluated mechanical properties and provided an alternative for the use of these
wastes in paving works, which usually require large quantities of material.

Author contributions
- The application of a Design of Experiments (DOE) of mixtures in a simplex centroid network for dosage of mixtures of soil samples
and steel waste through the optimization of mechanical properties with the aim of their application in structural layers of pavements.
This methodology is widely used in the field of civil construction, but has not yet been applied in the field of paving, where essentially
empirical methods are used. The DOE approach allows optimization based on statistical analysis, which favors its applicability to any
other types of mixtures with different materials.
- A chemical, mineralogical and microstructural analysis of mixtures between soil and Electric Arc Furnace Slag, showing the main
compounds possibly formed and their relationships with the variations in the results of mechanical properties.

11
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

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

Data will be made available on request.

Acknowledgments
The authors would like to thank the Department of Civil Engineering and the Infratest Research Group, which supported this study.

References
[1] IAB, Relatório de sustentabilidade, Instituto Aço Brasil, 2019, 84.
[2] K. Traven, M. Češnovar, V. Ducman, Particle size manipulation as an influential parameter in the development of mechanical properties in electric arc furnace
slag-based AAM, Ceram. Int. 45 (17) (2019) 22632–22641.
[3] W. Li, R. Li, Y. Chen, et al., Comparison of two sulfate-bearing soils stabilized with reactive magnesia-activated ground granulated blast furnace slag: swelling,
strength, and Mechanism 13 (1) (2023) 230.
[4] A.S. Reddy, R.V.P. Chavali, Solidification/stabilization of copper-contaminated soil using magnesia-activated blast furnace slag, Innovative Infrastructure
Solutions 8 (2) (2023) 79, 2023/01/25.
[5] H. Soeizi, A. Ardakani, Use of concrete waste and blast furnace slag in the stabilization of low plasticity clay soil, J Journal of Transportation Infrastructure
Engineering 9 (1) (2023) 57–75.
[6] İ. Keskin, M. Salimi, E.Ö. Ateyşen, et al., Comparative study of swelling pressure in expansive soils considering different initial water contents and BOFS
stabilization, Adv. Civ. Eng. 2023 (2023), 4823843, 2023/02/02.
[7] M.J. Rezaei-Hosseinabadi, M. Bayat, B. Nadi, et al., Sustainable utilisation of steel slag as granular column for ground improvement in geotechnical projects,
Case Stud. Constr. Mater. 17 (2022), e01333, 2022/12/01/.
[8] I.Z. Yildirim, U. Balunaini, M. Prezzi, Strength-Gain Characteristics and Swelling Response of Steel Slag and Steel Slag–Fly Ash Mixtures, vol. 35, 2023,
04023223, 8.
[9] A.S. Brand, P. Singhvi, E.O. Fanijo, et al., Stabilization of a clayey soil with ladle metallurgy furnace slag fines, Materials 13 (19) (2020) 4251.
[10] B. Xu, Y. Yi, Use of ladle furnace slag containing heavy metals as a binding material in civil engineering, Sci. Total Environ. 705 (2020), 135854.
[11] E.C. Lopes, T.O. da Silva, H.N. Pitanga, et al., Stabilisation of clayey and sandy soils with ladle furnace slag fines for road construction, Road Mater. Pavement
Des. 24 (1) (2023) 247–266, 2023/01/02.
[12] M. Amran, S. Debbarma, T. Ozbakkaloglu, Fly ash-based eco-friendly geopolymer concrete: a critical review of the long-term durability properties, Construct.
Build. Mater. 270 (2021), 121857, 2021/02/08/.
[13] A.R.K. Gollakota, V. Volli, C.-M. Shu, Progressive utilisation prospects of coal fly ash: a review, Sci. Total Environ. 672 (2019) 951–989, 2019/07/01/.
[14] R. Palod, S. Deo, G. Ramtekkar, Review and suggestions on use of steel slag in concrete and its potential use as cementitious component combined with GGBS,
Int. J. Civ. Eng. Technol. 8 (4) (2017) 1026–1035.
[15] C.S.G. Penteado, B.L. Evangelista, G.C.d.S. Ferreira, et al., Use of electric arc furnace slag for producing concrete paving blocks, Ambiente Construído 19 (2)
(2019) 21–32.
[16] A. Anand, R. Sarkar, A comprehensive study on bearing behavior of cement–fly ash composites through experimental and probabilistic investigations,
Innovative Infrastructure Solutions 6 (1) (2020) 39, 2020/11/26.
[17] M. Ozturk, O. Akgol, U.K. Sevim, et al., Experimental work on mechanical, electromagnetic and microwave shielding effectiveness properties of mortar
containing electric arc furnace slag, Construct. Build. Mater. 165 (2018) 58–63, 2018/03/20/.
[18] S. Apithanyasai, N. Supakata, S. Papong, The potential of industrial waste: using foundry sand with fly ash and electric arc furnace slag for geopolymer brick
production, Heliyon 6 (3) (2020), e03697.
[19] Ç. Bağlan, O. Aydin, O. Isik et al., Development of Aerated Concrete Brick with Chemical Additives from Electric Arc Furnace Slag Waste, Building for the
Future: Durable, Sustainable, Resilient. pp. 628-637..
[20] M.f. Alsheltat, M.A. Elfigih, Effects of electric arc furnace slag powder and fly ash within ternary waste blend on performance of concrete, Open Ceramics 14
(2023), 100359, 2023/06/01/.
[21] F. Özalp, Effects of electric arc furnace (EAF) slags on mechanical and permeability properties of paving stone, kerb and concrete pipes, Construct. Build.
Mater. 329 (2022), 127159, 2022/04/25/.
[22] K. Randhawa, and R. Chauhan, Investigating Strength Behaviour of Pond Ash as a Construction Fill Material. p. 012085..
[23] P. Mikhailenko, Z. Piao, L.D. Poulikakos, Electric arc furnace slag as aggregates in semi-dense asphalt, Case Stud. Constr. Mater. 18 (2023), e02049, 2023/07/
01/.
[24] P. Sukmak, G. Sukmak, P. De Silva, et al., The potential of industrial waste: electric arc furnace slag (EAF) as recycled road construction materials, Construct.
Build. Mater. 368 (2023), 130393, 2023/03/03/.
[25] S.A. Ziaee, F. Moghadas Nejad, M. Dareyni, et al., Evaluation of rheological and mechanical properties of hot and warm mix asphalt mixtures containing
Electric Arc Furnace Slag using gyratory compactor, Construct. Build. Mater. 378 (2023), 131042, 2023/05/16/.
[26] O.S.B. Al-Amoudi, A.A. Al-Homidy, M. Maslehuddin, et al., Method and mechanisms of soil stabilization using Electric Arc Furnace dust, Sci. Rep. 7 (1) (2017),
46676, 2017/04/28.
[27] H. Aldeeky, O. Al Hattamleh, Experimental study on the utilization of fine steel slag on stabilizing high plastic subgrade soil, Adv. Civ. Eng. 2017 (2017).
[28] M.M. Zumrawi, A.A.-A.A. Babikir, Laboratory study of steel slag used for stabilizing expansive soil, University Of Khartoum Engineering Journal 6 (2) (2017).
[29] D.H. Diniz, J.M. F.d. Carvalho, J.C. Mendes, et al., Blast oxygen furnace slag as chemical soil stabilizer for use in roads, J. Mater. Civ. Eng. 29 (9) (2017),
04017118.
[30] A.I.M. Ismail, S.A. Awad, M.A.G. Mwafy, The utilization of Electric Arc furance slag in soil improvement, Geotech. Geol. Eng. 37 (1) (2019) 401–411, 2019/
01/01.
[31] N. Sebbar, A. Lahmili, L. Bahi et al., Treatment of clay soils with steel slag, in Road Engineering. p. 02017..
[32] S. Shahsavani, A.H. Vakili, M. Mokhberi, Effects of freeze-thaw cycles on the characteristics of the expansive soils treated by nanosilica and Electric Arc
Furnace (EAF) slag, Cold Reg. Sci. Technol. 182 (2021), 103216, 2021/02/01/.
[33] M.N.-T. Lam, S. Jaritngam, D.-H. Le, Roller-compacted concrete pavement made of Electric Arc Furnace slag aggregate: mix design and mechanical properties,
Construct. Build. Mater. 154 (2017) 482–495, 2017/11/15/.
[34] F.I. Shalabi, I.M. Asi, H.Y. Qasrawi, Effect of by-product steel slag on the engineering properties of clay soils, Journal of King Saud University-Engineering
Sciences 29 (4) (2017) 394–399.
[35] E.C. Lopes, T.O. da Silva, H.N. Pitanga, et al., Application of electric arc furnace slag for stabilisation of different tropical soils, Int. J. Pavement Eng. (2021)
1–12.

12
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

[36] A.K. Ram, P. Rawat, S. Mohanty, Strength performance of soil-fly ash-MSW fine layered system under different controlled loading conditions: a comparative
study, Construct. Build. Mater. 369 (2023), 130524, 2023/03/10/.
[37] A.R. Reddy, K.L. Singh, Performance of pavement subgrade using fly ash stabilized peat soil reinforced with nylon fiber, Int. J. Pavement Res. Technol. (2023),
2023/03/02 1-13.
[38] E. Lal Mohammadi, E. Khaksar Najafi, P. Zanganeh Ranjbar, et al., Recycling industrial alkaline solutions for soil stabilization by low-concentrated fly ash-
based alkali cements, Construct. Build. Mater. 393 (2023), 132083, 2023/08/22/.
[39] Y.P. Arias-Jaramillo, D. Gómez-Cano, G.I. Carvajal, et al., Evaluation of the effect of binary fly ash-lime mixture on the bearing capacity of natural soils: a
comparison with two conventional stabilizers, Lime and Portland Cement 16 (11) (2023) 3996.
[40] A.A.M.S. Mohamed, J. Yuan, M. Al-Ajamee, et al., Improvement of expansive soil characteristics stabilized with sawdust ash, high calcium fly ash and cement,
Case Stud. Constr. Mater. 18 (2023), e01894, 2023/07/01/.
[41] T. Hoang, H. Do, J. Alleman, et al., Comparative evaluation of freeze and thaw effect on strength of BEICP-stabilized silty sands and cement- and fly ash-
stabilized soils, Acta Geotechnica 18 (2) (2023) 1073–1092, 2023/02/01.
[42] B. Xu, Y. Yi, Soft clay stabilization using ladle slag-ground granulated blastfurnace slag blend, Appl. Clay Sci. 178 (2019), 105136.
[43] G. Wang, Y. Wang, Z. Gao, Use of steel slag as a granular material: volume expansion prediction and usability criteria, J. Hazard Mater. 184 (1) (2010)
555–560, 2010/12/15/.
[44] K.C. Onyelowe, G. Alaneme, D.B. Van, et al., Generalized review on EVD and constraints simplex method of materials properties optimization for civil
engineering, Civil Engineering Journal 5 (3) (2019) 729–749.
[45] D. Jiao, C. Shi, Q. Yuan, et al., Mixture design of concrete using simplex centroid design method, Cement Concr. Compos. 89 (2018) 76–88.
[46] V.S.C. Medeiros, L.G. Pedroti, B.C. Mendes, et al., Study of mixtures using simplex design for the addition of chamotte in clay bricks, Int. J. Appl. Ceram.
Technol. 16 (6) (2019) 2349–2361.
[47] Z.B. Yıldırım, M. Karacasu, V. Okur, Optimisation of Marshall Design criteria with central composite design in asphalt concrete, Int. J. Pavement Eng. 21 (5)
(2020) 666–676, 2020/04/15.
[48] A. Azevedo, J. Alexandre, M. Marvila, et al., Evaluation of Technological Properties of Soil-Cement Blocks Using Experimental Design of Mixtures,
Characterization of Minerals, Metals, and Materials, vol. 2019, Springer, 2019, pp. 647–655.
[49] M.M.S. Lopes, R.d.C.S.S.A. Alvarenga, L.G. Pedroti, et al., Influence of the incorporation of granite waste on the hiding power and abrasion resistance of soil
pigment-based paints, Construct. Build. Mater. 205 (2019) 463–474, 2019/04/30/.
[50] A. Solouki, G. Viscomi, P. Tataranni, et al., Preliminary evaluation of cement mortars containing waste silt optimized with the design of experiments method,
Materials 14 (3) (2021) 528.
[51] G. Nakkeeran, L. Krishnaraj, Prediction of cement mortar strength by replacement of hydrated lime using RSM and ANN, Asian Journal of Civil Engineering 24
(5) (2023) 1401–1410, 2023/07/01.
[52] Q. Zhang, X. Feng, X. Chen, et al., Mix design for recycled aggregate pervious concrete based on response surface methodology, Construct. Build. Mater. 259
(2020), 119776, 2020/10/30/.
[53] S. Jamil, J. Shi, M. Idrees, Effect of various parameters on carbonation treatment of recycled concrete aggregate using the design of experiment method,
Construct. Build. Mater. 382 (2023), 131339, 2023/06/13/.
[54] R. Priyanga, A. Muthadhi, Optimization of compressive strength of cementitious matrix composition of Textile Reinforced Concrete–Taguchi approach, Results
in Control and Optimization 10 (2023), 100205, 2023/03/01/.
[55] K. Onyelowe, G. Alaneme, C. Igboayaka, et al., Scheffe optimization of swelling, California bearing ratio, compressive strength, and durability potentials of
quarry dust stabilized soft clay soil, Materials Science for Energy Technologies 2 (1) (2019) 67–77.
[56] E. Emmanuel, N. Fogne Appiah, P. Agyemang, et al., Simplex lattice strength optimization of lime-micro silica stabilized coir fiber-reinforced soil, J. Nat. Fibers
(2020) 1–17.
[57] I.C. Attah, R. Kufre Etim, G. Uwadiegwu Alaneme, et al., Scheffe’s approach for single additive optimization in selected soils amelioration studies for cleaner
environment and sustainable subgrade materials, Cleaner Materials 5 (2022), 100126, 2022/09/01/.
[58] G.U. Alaneme, I.C. Attah, R.K. Etim, et al., Mechanical properties optimization of soil—cement kiln dust mixture using extreme vertex design, International
Journal of Pavement Research and Technology 15 (3) (2022) 719–750, 2022/05/01.
[59] ABNT NBR 12653, Materiais Pozolânicos - Requisitos, Associação Brasileira de Normas Técnicas, 2014.
[60] ASTM Standard C618-22, Standard Specification for Coal Fly Ash and Raw or Calcined Natural Pozzolan for Use in Concrete, ASTM International, 2022.
[61] ASTM Standard D6913/D6913M-17, Standard Test Methods for Particle-Size Distribution (Gradation) of Soils Using Sieve Analysis, ASTM International, 2017.
[62] ASTM Standard D7928-21e1, Standard Test Method for Particle-Size Distribution (Gradation) of Fine-Grained Soils Using the Sedimentation (Hydrometer)
Analysis, ASTM International, 2021.
[63] ASTM Standard D4318-17e1, Standard Test Methods for Liquid Limit, Plastic Limit, and Plasticity Index of Soils, ASTM International, 2018.
[64] ASTM Standard D854-14, Standard Test Methods for Specific Gravity of Soil Solids by Water Pycnometer, ASTM International, 2016.
[65] ASTM Standard D698-12(2021), Standard Test Methods for Laboratory Compaction Characteristics of Soil Using Standard Effort (12,400 Ft-Lbf/ft3 (600 kN-
M/m3)), ASTM International, 2021.
[66] J.M. Manso, V. Ortega-López, J.A. Polanco, et al., The use of ladle furnace slag in soil stabilization, Construct. Build. Mater. 40 (2013) 126–134.
[67] I. Akinwumi, Soil Modification by the Application of Steel Slag, Period. Polytech. Civ. Eng. 58 (4) (2014) 371–377. https://doi.org/10.3311/PPci.7239.
[68] M. Mahmudi, S. Altun, T. Eskisar, Experimental and numerical evaluation of clay soils stabilized with Electric Arc Furnace (EAF) slag, Advances in Sustainable
Construction Resource Management 144 (2021) 73.
[69] Y.-N. Sheen, D.-H. Le, M.N.-T. Lam, Performance of Self-Compacting Concrete with Stainless Steel Slag versus Fly Ash as Fillers: A Comparative Study,
Periodica Polytechnica Civil Engineering, 2021.
[70] ABNT NBR 11579, Portland Cement — Determination of Fineness Index by Means of the 75 μm Sieve (Nº 200), Associação Brasileira de Normas Técnicas,
2012.
[71] ABNT NBR 16372, Portland Cement and Other Powdered Materials — Determination of Fineness by the Air Permeability Method (Blaine Method), Associação
Brasileira de Normas Técnicas, 2019.
[72] ABNT NBR 16605, Portland Cement and Other Powdered Material — Determination of the Specific Gravity, Associação Brasileira de Normas Técnicas, 2017.
[73] ISO 13320, Particle Size Analysis — Laser Diffraction Methods, vol. 2020, International Organization for Standardization, 2020.
[74] DNER-CLA 259, Classificação de solos tropicais para finalidades rodoviárias utilizando corpos-de-prova compactados em equipamento miniatura,
Departamento Nacional de Estradas de Rodagem, 1996.
[75] AASHTO M 145-91, Standard Specification for Classification of Soils and Soil-Aggregate Mixtures for Highway Construction Purposes, vol. 2021, American
Association of State and Highway Transportation Officials, 2021.
[76] ASTM Standard D2487-17e1, Standard Practice for Classification of Soils for Engineering Purposes (Unified Soil Classification System), ASTM International,
2020.
[77] ABNT NBR NM 18, Portland Cement - Chemical Anlysis - Determination of Loss on Ignition, Associação Brasileira de Normas Técnicas, 2012.
[78] A.L. Borges Marinho, C.M. Mol Santos, J.M. F.d. Carvalho, et al., Ladle furnace slag as binder for cement-based composites, J. Mater. Civ. Eng. 29 (11) (2017),
04017207.
[79] S.A. Ziaee, K. Behnia, Evaluating the effect of electric arc furnace steel slag on dynamic and static mechanical behavior of warm mix asphalt mixtures, J. Clean.
Prod. 274 (2020), 123092, 2020/11/20/.
[80] C.-C. Li, C.-M. Lin, Y.-E. Chang, et al., Stabilization and crystal characterization of Electric Arc Furnace oxidizing slag modified with ladle furnace slag and
alumina, Metals 10 (4) (2020) 501.

13
K.H.P. Rodrigues et al. Journal of Building Engineering 78 (2023) 107761

[81] ABNT NBR NM 13, Portland Cement — Chemical Analysis — Determination of the Free Calcium Oxide by Ethylenglycol, Associação Brasileira de Normas
Técnicas, 2012.
[82] ABNT NBR NM 15, Portland Cement - Chemical Analysis - Determination of Insoluble, Associação Brasileira de Normas Técnicas, 2012.
[83] D.F. Graffitti, Avaliação do teor de cal livre em escória de aciaria elétrica, 2002.
[84] ABNT NBR 10004, Solid Waste - Classification, Associação Brasileira de Normas Técnicas, 2004.
[85] ABNT NBR 10005, Procedure for Obtention Leaching Extract of Solid Wastes, Associação Brasileira de Normas Técnicas, 2004.
[86] ABNT NBR 10006, Procedure for Obtention of Solubilized Extraction of Solid Wastes, Associação Brasileira de Normas Técnicas, 2004.
[87] J.M. Montenegro-Cooper, M. Celemín-Matachana, J. Cañizal, et al., Study of the expansive behavior of ladle furnace slag and its mixture with low quality
natural soils, Construct. Build. Mater. 203 (2019) 201–209, 2019/04/10/.
[88] E. Ficagna, A. Gava, S.B. Rossato, et al., Effect on Merlot Red Wine of Fining Agents Mixture: Application of the Simplex Centroid Design, Food Science and
Technology, 2020. AHEAD.
[89] F. Maghool, F.A. Arulrajah, S. Horpibulsuk, Y.J. Du, Geotechnical Properties of Ladle Furnace Slag in Roadwork Applications, Sixth International Conference
on Geotechnique, Construction Materials and Environment, 2016. Bangkok, Thailand.
[90] ABNT NBR 12025, Soil- Cement — Simple Compression Test of Cylindrical Specimens — Method of Test, Associação Brasileira de Normas Técnicas, 2012.
[91] DNIT ME 134, Pavimentação – Solos – Determinação do módulo de resiliência – Método de ensaio, Departamento Nacional de Infraestrutura de transportes,
2018.
[92] ABNT NBR 9895, Solo – Índice de Suporte Califórnia (ISC) - Método de ensaio, Associação Brasileira de Normas Técnicas, 2016.
[93] ABNT NBR 12253, Soil-cement — Mixture for Use in Pavement Layer — Procedure, Associação Brasileira de Normas Técnicas, 2012.
[94] Manual de Pavimentação, Departamento Nacional de Infraestrutura de Transportes, 2006, p. 274.
[95] G.P. Gautreau, Z. Zhang, Z. Wu, Accelerated Loading Evaluation of Subbase Layers in Pavement Performance, Louisiana Transportation Research Center, 2010.
[96] AASHTO T220-66, Standard Method of Test for Determination of the Strength of Soil-Lime Mixtures, vol. 2018, American Association of State and Highway
Transportation Officials, 2013.
[97] AASHTO, Mechanistic-Empirical Pavement Design Guide: A Manual of Practice, third ed., American Association of State Highway and Transportation Officials,
2020, 264, [Washington D.C.].
[98] FUNASA, Ordinance Nº 1.469/2000: Approves the Control and Monitoring of the Quality of Water for Human Consumption and its Standard of Potability,
2001.
[99] A.A. Al-Homidy, O. Al-Amoudi, M. Maslehuddin, et al., Stabilisation of dune sand using electric arc furnace dust, Int. J. Pavement Eng. 18 (6) (2017) 513–520.
[100] B. Bose, Geo engineering properties of expansive soil stabilized with fly ash, Electron. J. Geotech. Eng. 17 (1) (2012) 1339–1353.
[101] N.C. Consoli, D.N. Giese, H.B. Leon, et al., Sodium chloride as a catalyser for crushed reclaimed asphalt pavement – fly ash – carbide lime blends,
Transportation Geotechnics 15 (2018) 13–19, 2018/06/01/.
[102] A. S. S. Gunarti, and I. Raharja, Mechanical Properties Improvement of Clays Using Silica Sand Waste and Dust Sand Foundry Waste. p. 012002..
[103] N.C. Consoli, A.D. Rosa, R.B. Saldanha, Variables governing strength of compacted soil–fly ash–lime mixtures, J. Mater. Civ. Eng. 23 (4) (2011) 432–440.
[104] M. Parsaei, A.H. Vakili, M. Salimi, et al., Effect of Electric Arc and Ladle Furnace Slags on the Strength and Swelling Behavior of Cement-Stabilized Expansive
Clay, Bulletin of Engineering Geology and the Environment, 2021, 2021/06/10.
[105] D.C. Montgomery, Design and Analysis of Experiments, John wiley & sons, 2017.
[106] O. Amini, M. Ghasemi, Laboratory study of the effects of using magnesium slag on the geotechnical properties of cement stabilized soil, Construct. Build.
Mater. 223 (2019) 409–420, 2019/10/30/.
[107] J. Wu, Q. Liu, Y. Deng, et al., Expansive soil modified by waste steel slag and its application in subbase layer of highways, Soils Found. 59 (4) (2019) 955–965.

14

You might also like