You are on page 1of 15

Applied Clay Science 184 (2020) 105390

Contents lists available at ScienceDirect

Applied Clay Science


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

Research Paper

Mechanical and compressibility characteristics of a soft clay stabilized by T


slag-based mixtures and geopolymers

Mahdi Salimi, Ali Ghorbani
Department of Civil Engineering, Faculty of Engineering, University of Guilan, Rasht, Guilan, Iran

A R T I C LE I N FO A B S T R A C T

Keywords: In the present study, industrial wastes such as Granulated Blast Furnace Slag (GBFS) and Basic Oxygen Furnace
Soft clay Slag (BOFS) activated with calcium oxide (CaO) and medium reactive magnesia (MgO) are used for chemical
Stabilization stabilization of a soft clay. This environmentally friendly approach can eliminate the hazards associated with
GBFS improper waste disposal and reduce greenhouse gas emissions generated by cement production. To this end,
BOFS
various amounts of additives (ranging from 2.5 to 20%) with the activator:slag ratio of 1:3 are added to kaolinite
CaO
clay and cured at two temperatures of 20 and 45°C. A series of laboratory tests, including pH, Electrical
MgO
Geopolymer Conductivity (EC), one-dimensional consolidation, Unconfined Compressive Strength (UCS), are conducted on
the stabilized samples. The increasing temperature causes a faster formation of cementitious products and a
higher UCS value, as confirmed by the SEM micrographs and XRD analysis, particularly in the case of MgO-BOFS
(MB) and CaO-BOFS (CB) samples with the UCS values of 4 and 4.7 MPa after 90 days of curing, respectively.
Furthermore, the MB- and CB-stabilized clay samples show a better compressibility characteristic compared to
the MgO-GBFS (MG) and CaO-GBFS (CG) blends. To further enhance the activity of the additives and to prepare
a slag-clay based geopolymers, two types of alkaline solutions at various Na2SiO3:NaOH ratios are added to the
mixtures at the final step. In addition, the energy absorption capacity (Eu) and the secant modulus (E50) of the
optimum blends are determined to assess the toughness and stiffness of the samples. The results indicate that the
UCS values of the MB and CB samples increase up to 7.41 MPa and 8.44 MPa after 90 days of curing, respec-
tively, when the Na2SiO3:NaOH ratio is 80:20. Generally, the use of slag-clay based geopolymers, particularly
BOFS, is very effective to address the problems associated with the soft soil and the optimum mixtures are
successful in decreasing settlements and enhancing compressive strength of the soft soil, which can be con-
sidered for use as a pavement base material.

1. Introduction 2014). Lack of adequate access to suitable replacement materials and


the high cost of this technique have led scholars to seek alternative
Soft soils (e.g., marine and kaolin clays) typically contain a very approaches. Based on recent studies, soil stabilization practices, how-
high percentage of clay and thereby have low bearing capacity, extra- ever, have been introduced as practical techniques; so that chemical
ordinary compressibility, erodibility, high water absorption and ex- stabilization is considered to be one of the easiest and the most eco-
cessive settlement behavior (Zukri, 2013; Hamidi and Marandi, 2018). nomical approaches to improve the soil behavior (Yadu and Tripathi,
This type of soil, which is generally taken into account as a problematic 2013; Goodarzi and Salimi, 2015a; Mola-Abasi et al., 2018a; Vakili
soil in geotechnical engineering (Emmanuel et al., 2019), contributes to et al., 2018; Ghorbani and Salimzadehshooiili, 2019; Ghorbani et al.,
the underlying problems in construction projects due to its undesirable 2019). Previous studies in the field of clay stabilization with chemical
characteristics. Furthermore, slow pore water flow caused by low per- additions have shown the efficiency of this method for improving
meability of soft clays can lead to severe damages to the foundation construction materials. The chemical stabilization method can be as-
structures. Hence, the improvement of soft soils is inevitable prior to cribed to the addition of cementitious materials, such as cement, lime,
any engineering operations for achieving properties tailored to the re- gypsum, to the soil in order to enhance its properties. Previous research
quirements. One of the traditional soil improvement methods is re- has shown that the inclusion of such additives gives rise to chemical
moving the soil and substituting it with other materials (Ghobadi et al., reactions contributing to the enhancement of soil characteristics


Corresponding author.
E-mail addresses: Salimi@phd.guilan.ac.ir (M. Salimi), ghorbani@guilan.ac.ir (A. Ghorbani).

https://doi.org/10.1016/j.clay.2019.105390
Received 13 July 2019; Received in revised form 24 November 2019; Accepted 26 November 2019
Available online 30 November 2019
0169-1317/ © 2019 Elsevier B.V. All rights reserved.
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

including Atterberg limits, soil strength, soil porosity, damping and waste disposal. It should be noted that these industrial wastes are more
stiffness (Consoli et al., 2018a,b; Ghorbani and Salimzadehshooiili, cost-effective than conventional binders. In addition to the economic
2018; Mola-Abasi et al., 2018b,c; Pasupathy et al., 2018). and environmental advantages of slag-based geopolymers to lime and
However, the production of these common soil modifiers is asso- cement, the higher mechanical performance of geopolymers makes
ciated with the release of large amounts of greenhouse gases in the them more suitable alternatives (Yaghoubi et al., 2019). It is worth
atmosphere which lead to inevitable environmental consequences noting that the mechanical effectiveness of such additives is due to their
(Phummiphan et al., 2018; Afrasiabian et al., 2019; Khajeh et al., chemical compositions (in particular CaO, SiO2 and Fe2O3), for in-
2019). Likewise, cement production is an energy-intensive process and stance, the grindability of materials containing slag can be attributed to
emits about 4 billion tons of CO2 each year (Yusuf et al., 2014). the presence of Fe2O3, as well as the pozzolanic (i.e., the ability to react
Moreover, the use of lime and cement can arise some limitations and with calcium-based materials and produce cementitious compounds)
lead to volume instability in post-stabilization phase due to the devel- and hydraulic characteristics are due to SiO2 and CaO contents, re-
opment of expansive minerals (e.g. thaumasite and ettringite). In recent spectively (Pan et al., 2016).
years, apart from the aforementioned drawbacks, the increase in costs To activate the slag and to accelerate pozzolanic reactions, there is a
associated with the application of traditional additives, forces re- need for alkali products such as lime. The addition of lime to the GBFS
searchers to replace them with environmentally friendly alternatives. and BOFS resulted in an increase in the pH to about 12.4, which sig-
Kim and Palomino (2009) evaluated fabric development in kao- nificantly increases the formation of calcium aluminate hydrate (CAH)
lin–polyacrylamide systems using sedimentation tests, viscosity and li- and calcium silicate hydrate (CSH) gels as illustrated in Eqs. 1–4
quid limit measurements, and SEM analysis. They indicated that kaolin (Ghorbani and Hasanzadehshooiili, 2018; Jahandari et al., 2019) and
clay fabric can be successfully treated with the application of poly- consequently, enhance the strength of the lime-slag blends.
acrylamide. Soltani et al. (2018) observed that the rubber powder in-
CaO + H2 O → Ca(OH)2 Hydration of quicklime (1)
clusion and polymer treatment could be enhanced the inferior en-
gineering properties of a highly expansive soil. Mirzababaei et al.
Ca(OH)2 → Ca2 + + 2(OH)− Ionization of calcium hydroxide;
(2018) investigated the simultaneous effects of short propylene fibers
and poly vinyl alcohol (PVA) as a nontraditional polymeron on the pH rises to about 12.4 (2)
mechanical characteristics of a high plasticity clay. Butane Tetra Car-
boxylic Acid (BTCA) was also used for making PVA hydrogels insoluble Ca2 + + 2(OH)− + SiO2 → 3CaO. 2SiO2 . 3H2 O
in water. They showed that the UCS and ductility of the clay could be Calcium Silicate Hydrate (CSH) (3)
significantly improved with the combined use of fiber reinforcement,
PVA and BTCA. Soltani et al. (2019) conducted a set of experimental Ca2 + + 2(OH)− + Al2O3 → 3CaO. Al2O3 . 3H2 O
tests to evaluate the effect of sulphonated oil (SO) agent for stabiliza-
Calcium Aluminate Hydrate (CAH) (4)
tion of a highly expansive soil and found that the addition of SO at
optimum concentration could increase the strength and reduce the Lime incorporation with slag has been successfully used for geo-
swell–shrink related volume changes, even after wetting–drying cycles. technical applications by other researchers (Obuzor et al., 2012; Celik
Estabragh et al. (2010) evaluated the mechanical behavior of soil-ce- and Nalbantoglu, 2013; Yi et al., 2015; Phanikumar and Nagaraju,
ment mixtures with different amount of acrylic resin and indicated that 2018; Salimi et al., 2018). Apart from the lime, the addition of the
the strength of soil-cement was increased significantly with the addition magnesium oxide (MgO) to the slag has recently been increasingly
of acrylic resin. Alazigha et al. (2016) showed that lignosulfonate (LS) taken into account for improving the properties of clayey soils (Yi et al.,
substantial impact on the swelling potential of expansive soils and 2014a,b; Ghorbani et al., 2015; Burciaga-Díaz and Betancourt-Castillo,
could be an economical and eco-friendly replacement for traditional 2018; Zheng et al., 2019). Jin et al. (2013) assessed the hydration of
additives. MgO-activated GGBS paste by means of different reactive MgOs and
Geopolymer is a green cementing agent produced from a geological found that the MgOs reactivity and their chemical composition had an
origin or an industrial waste (Khan et al., 2016); an alkali-activated effect on the hydration kinetics without any changes in the hydration
aluminosilicate material rich in silica (SiO2) and alumina (Al2O3) products. Yi et al. (2014b) conducted an experimental study to evaluate
(Huang et al., 2018; Phummiphan et al., 2018). There are many liquid the properties of two types of soils with GGBS-MgO and GGBS-lime
activators, among them are sodium silicate (Na2SiO3) and sodium hy- blends over a range of ratios, and Portland cement alone (control). They
droxide (NaOH), which are widely applied for the preparation of geo- found that although reactive MgO and hydrated lime activated GGBS
polymer (Arulrajah et al., 2016; Phetchuay et al., 2016; Poltue et al., had a similar strength ranges in the long run, the reactive MgO was
2019; Sukmak et al., 2019). Sukmak et al. (2013) evaluated the effect of more efficient and reacted at a higher rate than that of lime provides
silty clay and fly ash as fine aggregates and pozzolanic material, re- reacts in the short term. Conducting a set of unconfined compressive
spectively, on the strength of the clay–fly ash geopolymer using a liquid strength (UCS) tests, Yi et al. (2016) investigated the magnesia re-
alkaline activator containing a mixture of Na2SiO3 and NaOH. Leong activity on the activating efficacy for the GGBS-stabilized soft clay and
et al. (2018) evaluated the strength development of residual soil–fly ash the results proved the positive contribution of MgO (especially the high
geopolymer using Na2SiO3/KOH (or NaOH) as the alkali activator. reactive one) for the aforementioned mixture by increasing the UCS
Recently, two types of by-products from the steel industry, values significantly over time. Hoy et al. (2018) studied the strength
Granulated Blast Furnace Slag (GBFS) and Basic Oxygen Furnace Slag and microstructural development of Recycled Asphalt Pavement (RAP)
(BOFS), have attracted increasing interest among researchers due to using slag-based geopolymer. Their results showed that the strength of
their pozzolanic properties (Goodarzi and Salimi, 2015b; Sharma and the RAP-slag blend enhanced remarkably with the increase of ground
Sivapullaiah, 2016; Ge et al., 2018; Sekhar and Nayak, 2018). These granulated slag content and/or curing time due to the slag geopoly-
high‑calcium materials can be considered as a major source to produce merization. Mohammadinia et al. (2016) assessed the geotechnical
geopolymers worldwide due to their ease of availability. The reuse of engineering and strength characteristics of fly ash–slag geopolymer-
such industrial by-products in the synthesis of geopolymers can greatly stabilized construction and demolition (C&D) materials to determine
limit the application of traditional additives, resulting in lower green- their efficiency for pavement base/subbase applications. They found
house gas emissions such as CO2 (Arulrajah et al., 2017b; that slag-based geopolymer had a higher compressive strength in
Mohammadinia et al., 2018). Furthermore, incorporating these mate- comparison with fly ash–based geopolymer. Kua et al. (2016) evaluated
rials into chemical reactions and preventing their direct exposure to the the UCS variations of coffee grounds-slag-fly ash geopolymer to achieve
environment also eliminates the hazards associated with improper a maximum strength. Hoy et al. (2016) assessed the possibility of using

2
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

fly ash-geopolymer to stabilized recycled asphalt pavement as a sus- Table 1


tainable road construction material. Arulrajah et al. (2019) enhanced Engineering and geo-environmental properties of the kaolin clay sample used in
the strength characteristics of coffee ground geopolymers using re- the study.
cycled glass as a supplementary filler material. Aziz et al. (2019) per- Soil property Value Standard Designation
formed an investigation on ground-granulated blast furnace slag
(GGBFS) geopolymers to examine the thermal efficiency and strength Cation exchange capacity (CEC), cmol/ 11.2 Hendershot and Duquette
kg (1986)
development.
Specific surface area (SSA), m2/kg 25,000 Eltantawy and Arnold (1973)
Despite extensive studies on geopolymers and their different com- Clay fraction, % 68 ASTM D422
pounds, research on the use of lime and magnesium oxide as promoters Liquid Limit, % 38.2 ASTM D4318
in the clay-slag based geopolymer has not been performed. Hence, in Plasticity Index, % 19 ASTM D4318
Specific gravity, GS 2.69 ASTM D854
the current study, physical-chemical properties and mechanical beha-
Maximum dry density, kg/m3 1560 ASTM D698
vior of the samples were investigated by performing a set of tests in- Optimum moisture content, % 28.5 ASTM D698
cluding EC, pH, UCS and one-dimensional consolidation with along X- Soil classification CL ASTM D2487
ray diffraction (XRD) and scanning electron microscope (SEM) analysis Unconfined compressive strength⁎, 0.18 ASTM D2166
on soft soils stabilized with lime, magnesium oxide and slag. Then, MPa
pH 8.82 Ouhadi et al. (2006)
incorporating alkaline solutions to the optimum mixtures, the perfor-
Electrical conductivity (EC), mS/cm 0.15 EPA manual (1983))
mance of lime and magnesium oxide as promoters on clay-slag based

geopolymer was observed. It is worth noting that the optimum blends The soil sample was mixed with the optimum moisture content and com-
can be used as a pavement base material, where enhancement of me- pacted at the maximum dry density.
chanical characteristics and reduction of settlement are needed.
(1986). The specific surface area (SSA) of the soil samples was de-
termined using the ethylene glycol-monoethyl ether (EGME) method
2. Materials and methods employed by Eltantawy and Arnold (1973). The mineralogical compo-
sition of the soil sample with X-ray diffraction technique showed that it
2.1. Materials contained high amount of kaolinite mineral (about 70%), followed by
quartz (about 30%). Moreover, the Optimum Moisture Content (OMC)
Kaolinite is characterized as a soft clay with low cation exchange and Maximum Dry Density (MDD) were determined to be 28.5% and
capacity (CEC) of about 3–15 meq per 100 g dry clay (Ghosh and 1560 kg/m3, respectively, in accordance with the standard Proctor
Bhattacharyya, 2002), relatively low surface charge and surface ac- compaction test with a compaction energy of 592 kJ/m3, which is de-
tivity. In the present study, kaolinite was selected as a soft soil obtained signated by ASTM D698 (ASTM, 2006).
from north of Iran (Jirandeh, Guilan - see Fig. 1), which engineering Apart from the kaolinite clay, several additives were applied to
and geo-environmental characteristics were measured according to stabilize the soil samples. Two types of slags including GBFS and BOFS
ASTM methods (ASTM, 2006) and the Environmental Protection were prepared from Esfahan Steel Co., Esfahan, Iran. It should be noted
Agency (EPA) manual (EPA, 1983), respectively, as shown in Table 1. that the kaolinite clay, GBFS and BOFS were oven-dried at 105 °C for
The pH and EC of the soil sample were measured in a 1:20 soil-water 24 h to remove any adsorbed water and passed through a No. 200 sieve
ratio. The CEC of the kaolinite clay was obtained using the BaCl2 so- in order to obtain uniform dry powders as shown in Fig. 2. Particle size
lution replacement method, as proposed by Hendershot and Duquette

Fig. 1. Location of Jirandeh (Map data © 2019 Google, Imagery © 2019 TerraMetrics.)

3
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 2. A view of GBFS, BOFS and Kaolinite clay used in this study.

100

90

80

70
Sieve #200
Percent finer (%)

60

50

40

30
Kaolinite
20
BOFS
10
GBFS
0
0.001 0.01 0.1 1
Particle Size (mm)

Fig. 3. Particle size distribution curves of the studied kaolinite clay, GBFS and BOFS.

Table 2
Chemical compositions of the kaolinite clay sample, BOFS and GBFS.
Oxide (wt%) SiO2 Al2O3 Fe2O3 CaO MgO Na2O K2O P2O5 SO3 TiO2 LOI⁎

Kaolin clay 58.26 29.43 1.14 0.89 0.16 < 0.1 0.51 – – < 0.1 9.23
BOFS 15.6 8.2 20.9 48 3.4 0.2 < 0.1 0.4 0.3 1.3 1.65
GBFS 32.1 16 6.4 27.1 8.8 0.6 0.8 < 0.1 3.4 2.9 1.85


Loss on ignition.

distribution curves of the studied clay and slags were determined by 2.2. Mix designs
wet sieving and hydrometer methods as per ASTM D422 (Fig. 3). The
laboratory grade calcium oxide (CaO) and medium reactive magnesia The kaolinite clay was treated with different amounts of GBFS and
(MgO) were obtained, as promoters of slags, from Merck Co. Darmstadt, BOFS using two types of activators including quick lime and magnesium
Germany. Medium type of MgO was selected because of its low cost and oxide (Table 3). For this purpose, the amounts of 2.5%, 5%, 10% and
its capability to activate slags (Jin et al., 2015; Du et al., 2016). 20% by weight of the additives in dry mass were added to the kaolinite.
Moreover, due to the widespread application of sodium silicate It should be noted that the range of 0–20% additives has been applied
(Na2SiO3) and sodium hydroxide (NaOH) to produce geopolymer, they in several previous studies (Jin et al., 2015; Goodarzi et al., 2016;
were employed as alkaline activators, which were obtained from Merck Salimi et al., 2018) to achieve the maximum possible strength for soil
Co. Darmstadt, Germany. The main chemical compositions of the used samples. Moreover, following Goodarzi and Movahedrad (2017) and Gu
materials were analysed using the X-ray fluorescence (XRF) method and et al. (2015), the activator to slag ratio was set as 1:3. Physical and
are given in Table 2. As it can be seen, the used by-products (i.e, GBFS mechanical characteristics of the prepared
and BOFS) have a high content of multivalent cations that can be ex- samples including consistency limits and compaction properties are
changed with the cations having lower valence on the clay surfaces. listed in Table 4. The clay-slag-activator mixtures were stirred with the
OMC for each sample to form a completely homogeneous composite

4
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Table 3 and SD for the UCS and consolidation tests = 0.01–0.4].


The mix proportion of agents and alkaline activators.
Test Total amount of Activator:slag = 1:3 (CG, CB, NaOH:Na2SiO3 2.3. Testing program and methods
group no. additives (%) MG, MB)⁎
The mechanical properties of the samples were assessed by con-
CaO or MgO GBFS or BOFS
(wt%) (wt%) ducting UCS test; so that the wet homogenized blends of amendments
and clay were statically placed into cylindrical steel molds with height
1 2.5 0.63 1.87 – and diameter of 70 mm and 35 mm, respectively, to achieve the max-
2 5 1.25 3.75 –
imum dry density. Having removed the samples from the mold, they
3 10 2.5 7.5 –
4 20 5 15 –
were cured similar to those described earlier in the sample preparation
5 20 5 15 100:0 section, and then UCS test was conducted on the stabilized samples
6 20 5 15 80:20 according to ASTM D2166 at a constant strain rate of 1.2 mm/min.
7 20 5 15 60:40 Performing a set of one-dimensional consolidation tests at the tem-
8 20 5 15 40:60
peratures of 20 and 45 °C for a 28-day curing time, according to ASTM
9 20 5 15 20:80
10 20 5 15 0:100 D2435, compression behavior of the untreated and treated kaolinite
clay was evaluated. According to the previous studies (Latifi et al.,

CG: CaO-GBFS; CB: CaO-BOFS; MG: MgO-GBFS; MB: MgO-BOFS. 2015, 2016), the increase in consolidation properties (i.e., CC and CS)
up to 28 days of curing is much more intense than for the ages above
and were placed in air-tight plastic bags. Then, they were cured in a 28 days, and therefore, it was chosen for this experiment as an optimum
chamber at different temperatures of 20 and 45 °C with a relative hu- curing time. The samples were first prepared similar to that for UCS test
midity of 85%. The temperature of 45 °C was chosen to simulate the hot and then were compacted in the oedometer ring with a diameter and
regions in the world. The soils in many areas are frequently exposed to height of 50 mm and 20 mm, respectively, at MDD and OMC of each
such a high temperature during summer (Arulrajah et al., 2017a) and it mix combination. After that, all the samples were initially consolidated
reflects the high end of observed field air temperatures (Webster et al., with a stress of 6.25 kPa that regularly increased by an increment ratio
2018). For each curing time (i.e. 7, 28 and 90 days), the tests were of 1 to a maximum pressure of 1600 kPa. It is important to note that the
carried out to assess the effects of slag-activator blend on the me- samples were kept saturated for 24 h in the first step of the loading
chanical and compressibility parameters of the kaolinite clay. (6.25 kPa) prior to the continuous of the other loading steps to absorb
In this study, two types of liquid alkaline activators including 10 M the water thoroughly. During the test, saturation will be provided by
sodium hydroxide (NaOH) and sodium silicate (Na2SiO3) containing topping up the water in the confining consolidation rings. A set of batch
Na2O, SiO2, and H2O with the amounts of 8.5%, 28.5% and 63% by equilibrium tests (i.e., pH and EC tests) was conducted so as to assess
weight, respectively, were used to produce clay-slag based geopolymer. the chemical reactions of amendments with the kaolinite clay particles
As mentioned above, GBFS, BOFS, CaO and MgO with the specified according to the EPA (1983). For this purpose, a suspension was pre-
amounts were dry mixed and then, alkaline solutions of sodium hy- pared with a ratio of clay-additive slurries to distilled water of 1:20 for
droxide and sodium silicate were slowly sprayed onto the mixtures. The each sample with a specified weight percentage and then was equili-
Na₂SiO₃:NaOH ratios studied were 100:0, 80:20, 60:40, 40:60, 20:80 brated on a horizontal shaker. Afterwards, the pH and EC of the ad-
and 0:100. In line with Rattanasak and Chindaprasirt (2009), and ditives-clay samples were measured after 1, 28 and 90 days of curing.
Somna et al. (2011), the 10 M NaOH solution was chosen for the cur- Microstructural analysis of natural kaolinite and treated samples with
rent study due to achieving the desired strength. The mix proportion of 20% additives (20% was chosen as an optimal amount in this study)
additives and alkaline activators of the current study are given in after 28 days of curing at 45 °C were performed employing SEM and
Table 3. During the experiments, distilled water (EC = 3.6 μS/cm; XRD tests. Air-dried pieces of the selected samples taken from post-test
pH = 6.9) was used to prepare the NaOH solution. It should be noted UCS samples were used for the micro-level investigation by following
that all tests were repeated in triplicate and the mean values were Dash and Hussain (2011), Hoy et al. (2016), Li et al. (2016), and
calculated to minimize variations and ensure consistency. However, Siddiqua and Barreto (2018). The samples were fractured to produce
very little difference was found between the results of repeated mea- freshly exposed surfaces for the test. Images of the samples were
sures of each test because of the accuracy in preparing and testing the magnified 5000 times using a scanning electron microscope modeled
samples [standard deviation (SD) for the EC and pH tests = 0.1–0.5, VEGA3-TESCAN. In line with previous research (Ghorbani and

Table 4
Physical and mechanical characteristics of the prepared samples.
Type of sample Amount of additives (%) LL (%) PL (%) PI (%) ωopt (%) γopt (kg/m3) Pc (kPa)

Kaolinite 0 38.2 19.2 19 28.5 1560 205


MB-stabilized 2.5 36.84 19.43 17.41 27.35 1582 381
5 35.45 19.94 15.51 26.27 1647 460
10 32.73 21.11 11.62 24.83 1672 552
20 29.37 21.76 7.61 22.39 1737 628
CB-stabilized 2.5 36.41 19.51 16.9 27.41 1595 485
5 34.72 20.08 14.64 26.32 1651 571
10 31.19 21.51 9.68 24.66 1683 646
20 27.32 22.72 4.6 21.87 1754 712
MG-stabilized 2.5 37.47 19.35 18.12 27.58 1564 232
5 36.65 19.51 17.14 26.61 1623 241
10 33.51 20.18 13.33 25.24 1639 325
20 30.69 20.44 10.25 22.86 1677 404
CG-stabilized 2.5 37.03 19.41 17.62 27.29 1573 370
5 36.41 19.64 16.77 26.58 1645 496
10 33.13 20.46 12.67 24.93 1656 544
20 29.84 20.84 9 22.76 1687 571

5
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

8
2.5% 5% 10% 20%
7

EC (mS/cm)
4

1
Kaolinite
0
MB CB MG CG MB CB MG CG MB CB MG CG
1 day of curing 28 days of curing 90 days of curing

a
7
2.5% 5% 10% 20%
6

4
EC (mS/cm)

1
Kaolinite
0
MB CB MG CG MB CB MG CG MB CB MG CG
1 day of curing 28 days of curing 90 days of curing

b
Fig. 4. Electrical conductivity of treated samples with additives after 1, 28 and 90 days of curing periods at a) 20 and b) 45 °C.

Salimzadehshooiili, 2019; Ghorbani et al., 2019), Microstructural a decrease in EC value with the passage of time up to 90 days as shown
Image Processing (MIP) was performed on SEM images of the treated in Fig. 4a. Such trends might be due to the decreased ion concentration
samples in order to determine the area of the voids using ImageJ in the mixtures owing to hydration and/or pozzolanic reactions, which
software. In the case of XRD, the samples were initially milled to pre- would entail the consumption of additives and a reduction in EC level.
pare homogeneous powders and then were scanned in the 2θ range of 4 Comparing Fig. 4a and b, a further decrease in EC value was found
to 60° for their XRD spectra using a Philips PW-1730 diffractometer in stabilized samples which is contributed to the progression of che-
(Step Size = 0.05°, Scan Step Time = 1 s, Anode Material = Cu, mical reactions between clay surfaces and additives, and to more con-
Kα = 1.5406 Å, Generator Settings = 40 kV and 30 mA). sumption of ions in the system. The results indicate that EC values of
samples cured at higher temperature (45 °C) experienced lower rates
3. Results and discussions compared to those of 20 °C. It is obvious that increasing ambient
temperature can increase hydration and pozzolanic reactions leading to
3.1. Physical and chemical properties decrease in the ion concentration.

3.1.1. Effect of additives (MB, CB, MG and CG) on the EC 3.1.2. Effect of additives (MB, CB, MG and CG) on the pH
Fig. 4 shows variations in the EC of treated samples with additives The results in Fig. 5 show that the pH values of MB- and CB-treated
cured at two different ambient temperatures after 1, 28 and 90 days of samples being slightly higher than those of MG and CG samples. So that
curing. It is evident that the EC values are considerably increased after in the presence of 5% MB or CB, the pH increased to about 12 and
the addition of the stabilizers, especially in MB and CB samples. Such attained about 12.5 with the more addition of stabilizers. Whereas the
that the EC of the samples cured at 20 °C is increased from 0.15 mS/cm pH value of 12 was recorded in the MG or CG contents up to 10%. This
for natural kaolinite to 6.1, 6.7, 2.51 and 3.39 mS/cm by adding 20% means that the pH condition in order to begin pozzolanic reactions is
MB, CB, MG and CG, respectively, after 1 day of curing at 20 °C further provided in the MB- or CB-treated samples that can improve the
(Fig. 4a). It is easily perceived that such an increase is caused by rising performance of additives. In addition, the rate of reduction in pH values
concentrations of ions such as Ca2+ and OH– from solubility or partial of MB and CB samples after 90 days of curing is significantly higher
solubility of the additives. As the EC level increases in clay-additive than those of MG and CG samples. This finding indicates the further
mixtures, the osmotic potential shows an increasing trend which re- pozzolanic activity of MB and CB combinations. Actually, activators can
duces the repulsive forces acting among the clay surfaces, and conse- be partly consumed by BOFS through pozzolanic reactions. The pre-
quently leads to the production of larger particles and can allow for sence of less unreacted MgO and CaO in MB and CB samples results in a
higher flocculation to take place. It is worthwhile to note that there was further reduction in the pH of the system. These results are strongly

6
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

13
2.5% 5% 10% 20%

12

11

pH
10

9
Kaolinite

8
MB CB MG CG MB CB MG CG MB CB MG CG
1 day of curing 28 days of curing 90 days of curing

Fig. 5. pH of treated samples with additives after 1, 28 and 90 days of curing periods at 20 °C.

consistent with the UCS tests and support that MB and CB have a in close agreement with the results of EC tests in Fig. 4 that presents an
greater activity compared to MG and CG. increase in the curing time has a minor influence on the performance of
samples with lower amounts of additives. The results of UCS tests also
3.2. Mechanical properties show that the type of activator has a significant effect on the behavior
of stabilized clayey soil. As can be seen in Fig. 7, samples containing
3.2.1. Effect of additives (MB, CB, MG and CG) on the UCS BOFS resulted in a higher level of strength than those prepared with
Performing UCS tests, the strength of treated samples was measured GBFS. Such that the UCS values of the samples with 20% MB and CB
to evaluate the impact of the additives on the mechanical strength of after 90 days of curing at 45 °C were measured 4.1 and 4.73 MPa, re-
kaolinite clay. To evaluate the strength improvement of treated sam- spectively, while these values for the same level of binder related to the
ples, a non-dimensional parameter is presented in Eq. (5) as follows MG and CG were about 2.54 and 3 MPa, respectively. Hence, BOFS
(Saride et al., 2013): exhibited a higher activity than GBFS for the production of pozzolanic
compounds, which increased the long-term strength. According to the
Strength Improvement Factor (SIF) EC tests, the improvement of soil behavior owing to such reactions can
UCS of treated sample cured at n days remarkably occur in the MB- and CB-treated samples, which is in line
= with the results of UCS tests. Based on Fig. 7, the ductility of the treated
UCS of untreated samople cured at n days (5)
samples was dependent on the type of additives. The addition of the
Fig. 6 shows the results of UCS tests for the samples cured for 7, 28 GBFS to the samples led to a more ductile behavior as well as decreased
and 90 days at 20 and 45 °C. The clayey soil demonstrated an increasing UCS value.
trend of UCS values in the presence of additives and the time pro- In this study, MB- and CB-stabilized samples with the content
gression. Based on the previous studies (Dash and Hussain, 2011; of > 10% after 7 days of curing at 45 °C met Malaysian's criteria with a
Coudert et al., 2019; Wang et al., 2019), such behavior can be largely minimum UCS of 0.8 MPa (or SIF > 4.44 based on the UCS value of
attributed to the development of short-term (i.e. cation exchange and untreated kaolinite equal to 0.18 MPa) for low-volume-traffic roads
flocculation–agglomeration) and long-term (i.e. pozzolanic activity) (Arulrajah et al., 2017a). After 90 days of curing, all samples cured at
reactions as well as the production of more cementitious compounds 20 and 50 °C met Malaysian's criteria except samples containing 2.5%
such as CSH and CAH gels which occupy the void space. The formation MG and 2.5% CG.
of these new materials is associated with the interlocking of the clay The area under the stress-strain curves up to the maximum stress
particles and consequently, gains a significant strength. It should be was calculated from Fig. 7, which is defined as peak strain energy or
noted that with the extension of curing time, a higher rate of strength energy absorption capacity (Eu) and regarded as a measure of toughness
development was gained ascribed to the fact that the strength of ma- of the material (Zhang et al., 2019; Zhao et al., 2019). Besides, the
trices containing chemical additives (e.g., CaO, MgO, BOFS, and GBFS) secant modulus (E50) of the samples were determined from Fig. 7,
is increased with time as far as alkaline conditions are present. In other which is characterized the stiffness or elasticity of the samples and
words, the progression of hydration and pozzolanic reactions may defined as the slope of a straight line from the origin to the 50% of
continue for a long time as long as adequate reactive clay minerals, free failure point in the stress-strain curve (Tsuchida et al., 2001). Fig. 8
multivalent ions (e.g., Ca) and sufficient water are available (Cherian presents the variations of Eu and E50 against 20% additives. It can be
and Arnepalli, 2015). seen that both modulus of Eu and E50 showed the same trend as found
As it can be seen in Fig. 6, following the rise in curing temperature, for the compressive strength of the samples. The Eu values for CB- and
a considerable increase in the SIF values was obtained in all samples, MB-stabilized samples were about 18 and 17 kJ/m3, respectively,
which can be explained by acceleration of the pozzolanic reactions in which were larger compared to those of GBFS-stabilized samples. In the
higher temperature. Moreover, the SIF of the samples cured at 45 °C case of E50, samples containing 20% CB and MB yielded values of 207
after 7 days of curing was higher than that of 20 °C. The development of and 190 MPa, respectively. While these values for CG and MG samples
strength at elevated temperatures is ascribed to a fast production of were equal to 126 and 119, respectively.
cementitious materials. These products are associated with better ri-
gidity of the clay particles and cause greater strength values.
It is important to note that slight changes happened in the strength 3.2.2. Effect of geopolymers on the UCS
of the samples containing lower than 5% stabilizers irrespective of the In this section, alkaline solutions act as activators, and CaO and
curing period. It implies that a small amount of any of the additives is MgO act as promoters on the clay-slag based geopolymers. Generally,
largely consumed in the short-term reactions and the formation of ce- the geopolymerization degree of slag-clay based geopolymer system is
mentitious products can be achieved in the higher percentages. This is small. Thus, the use of such promoters can improve compressive

7
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 6. SIF values of a) MB, b) CB, c) MG and d) CG samples cured for 7, 28 and 90 days of treatment at 20 and 45 °C.

8
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 10. Variations of peak strain energy (Eu) and secant modulus (E50) against
20% additives after 90 days of treatment at 45 °C.

Fig. 7. Stress-strain variation of treated samples with additives after 90 days of respectively.
treatment at 45 °C. Fig. 10 shows the variations of the Eu and E50 for geopolymer
samples containing 20% additives with an alkaline solution ratio of
80:0 (as optimum value). It is clear that the results are in good agree-
ment with the UCS value of the samples presented in Fig. 9. The de-
velopment of Eu and E50 is in favor of BOFS-containing samples. Such
that the Eu value of the CB sample was about 22% and 37% higher than
that of CG and MG samples, respectively. In other words, the required
energy to deform the BOFS-containing samples was greater than that of
GBFS. Similar results were obtained for the modulus of elasticity, which
indicated that the behavior of the MB and CG samples was improved
significantly. So that the E50 value of the CB sample compared with that
of MG and CG samples increased by 40 and 50%, respectively. These
findings are highly consistent with the results of UCS tests and confirm
that MB and CB have a greater activity compared to MG and CG.

Fig. 8. Variations of peak strain energy (Eu) and secant modulus (E50) against
20% additives. 3.2.3. Effect of additives (MB, CB, MG and CG) on the compressibility
properties
In this study, performing a set of one-dimensional consolidation
tests, the behavior of volume change of treated samples with various
additives was determined. To this end, the compression index (CC) was
measured by the e–log σ'v compression curves. The consolidation results
for MB, CB, MG and CG samples cured for 28 days of curing at 45 °C is
shown in Fig. 11. The results show that the increase of additive contents
leads to improved one-dimensional compression behavior. The samples
containing BOFS has much more effect on chemical reactions than
GBFS ones, such that the void ratio of the samples with 20% additives
including MB, CB, MG and CG was about 1.28, 1.26, 1.12 and 1.24,
respectively. The variations of the CC of the treated samples with var-
ious amounts of additives after 28 days of curing at 20 and 45 °C are
shown in Fig. 12. The results demonstrated that the addition of che-
mical additives to the clayey soil is accompanied by a decrease in
compression index, so that these changes are significantly greater in
Fig. 9. SIF variations of kaolinite clay with 20% additives after 90 days of
treated samples at higher temperature. Moreover, it is clear that the
treatment at 45 °C.
type and amount of amendments have a strong influence on the com-
pressibility of the samples. Reducing the compression rate of samples
strength of mortars. Fig. 9 shows the effect of different ratios of alkaline can be attributed to cation exchange and pozzolanic reactions. On the
solutions on the strength properties of kaolinite clay with 20% additives one hand, the replacement of multivalent cations from the additives by
(as an optimal amount) after 90 days of curing at 45 °C. According to monovalent ones of clay can give rise to a flocculation of the clay
the Fig. 9, the UCS development of treated samples is strongly depen- particles; a chemically induced pre-consolidation; and, as a con-
dent on the type of promoter and the ratio of Na2SiO3:NaOH. The sequence, a reduction in the compression index. On the other hand, the
lowest and highest SIF values were observed at ratios of 0:100 and development of pozzolanic reactions between additives and soil parti-
80:20. In other words, when the activation solution was NaOH, the cles leads to improved soil compressibility. As shown in Fig. 12, the
UCSs had a minimal value, whereas the increase of Na2SiO3 content highest and lowest reductions in the CC values belong to CB and MG
resulted in further increase of UCS. Furthermore, similar to the previous samples, respectively, which are in good agreement with other results
results, alkaline solutions in the 80:20 ratio also had the highest impact obtained from the pH and UCS tests.
on CB samples, which increased the strength by 42-fold (8.4 MPa) Furthermore, the pre-consolidation pressure was measured from the
compared to untreated soil. Moreover, SIF value for MB, CG, and MG void ratio-pressure curve employing the method suggested by
samples was obtained 37 (7.4 MPa), 32 (6.4 MPa) and 29 (5.8 MPa), Casagrande. It is worth noting that the increase of additive contents was

9
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

associated with an increase of pre-consolidation pressure. So that the


pre-consolidation pressure increased from 200 kPa for kaolinite clay to
628, 712, 404 and 571 kPa for stabilized clay with 20% of MB, CB, MG
and CG, respectively, as shown in Table 4. Such behavior can be at-
tributed to the pozzolanic reaction that forms calcium silicate hydrate
gel and results in increased strength. This means that the stabilized soil
is able to tolerate increased pressure without the occurrence of settle-
ment at higher additive contents.

3.3. Microstructural properties

3.3.1. XRD analysis


To better assess the interparticle reactions between the kaolinite
clay and other additives in various mixtures, the XRD and SEM analysis
were carried out. Fig. 13 depicts the XRD patterns for the natural
kaolinite and treated samples with 20% additives after 28 days of
curing at 45 °C. In all the treated samples, a significant reduction in the
intensities of the kaolinite (K) reflections (i.e. d001 = 3.56 and 7.13 Å)
accompanied by the presence of some new reflections of crystalline
products including CAH and CSH, confirming microstructural changes
in the treated samples. The decline in the kaolinite peak intensity by
addition of additives can be traced to the production of cementitious
materials and consequently, flocculated structures owing to cation ex-
change reactions. Such aggregated particles exhibits reduced intensities
due to the reduction in the reflection of incident ray compared with
their initial oriented kaolinite structure. Furthermore, the occurrence of
pozzolanic reactions can be dissolved the clay fractions, which reduces
the reflections of kaolinite mineral. It is obvious in Fig. 13 that in the
presence of a certain amount of each additive (20%), the CB- and MB-
treated samples represent a further formation of CSH and CAH gels
compared to that of CG and MG samples. This indicates that a re-
markable pozzolanic activity would occur in the former samples. The
findings can be interpreted by the fact that a greater presence of the
soluble ions in samples containing BOFS results in a higher increment in
the soil pH and EC values (see Figs. 4 and 5), a more dissolution of the
kaolinite clay fractions in the CB and MB mixtures, and as a con-
sequence, a higher decrease in peak intensities of the kaolinite mineral
(Fig. 13).
The natural kaolinite and treated samples were also investigated
using scanning electron microscopy to further support the finding in-
dicating that better efficiency of BOFS samples is largely due to the
developments of cementitious products (i.e., CSH and CAH gels).

3.3.2. SEM analysis


Fig. 14 displays the SEM and microstructural image processing
(MIP) of the untreated and treated samples cured for 28 days at 45 °C.
As it can be seen, the clay fabric in the presence of additives is fully
different from that of untreated sample. The arrangement of the clay
particles transforms from a dispersed form (Fig. 14a) to a more ag-
gregated nature and some new cementitious materials are produced by
adding additives, especially MB and CB samples (Fig. 14(b-e)). This can
be ascribed to the cation exchange reactions and substitution of mul-
tivalent ions from the additives by cations of lower valency in the clay
particles. The flocculated structure and patches of cementitious pro-
ducts were formed by the addition of stabilizers as shown in Fig. 14.
This is more significantly demonstrated in samples containing BOFS
(i.e., MB and CB) indicating higher activity of BOFS compared to GBFS,
as was also supported by tests provided in the proceeding sections. In
other words, the soil-additive matrix appears to have a less-strength-
ened structure paste in the MG- and CG-treated samples compared with
MB and CB samples, as shown in Fig. 14(f-i). It is worth noting that
particles are not fully bonded to each other, and there are large voids in
them, which are indicated in red (Fig. 14g and i). The development of
voids in the samples can be ascribed to the imperfect hydration process
Fig. 11. One-dimensional consolidation test results for a) MB, b) CB, c) MG and and the lack of appropriate packing of particles. The incidence of each
d) CG samples cured for 28 days of treatment at 45 °C. of these two cases interferes with the continuous structure of the

10
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 12. CC variations of the treated samples with various amounts of additives after 28 days of treatment at 20 and 45 °C.

Fig. 13. XRD patterns for the natural kaolinite and treated samples with 20% additives after 28 days of curing at 45 °C.

particles and might be associated with less increase in strength. The give rise to the faster formation of cementitious products, particu-
hydration and pozzolanic reactions lead to the production of cementi- larly in MB and CB samples; so that the UCS value of these two
tious products occupying the cavities. samples at the content of 20% after 90 days of curing attains 4 and
4.7 MPa, respectively.
4. Conclusions (3) The clay-BOFS based geopolymers, CB and MB samples, show the
highest increase on the UCSs of the samples in the Na2SiO3:NaOH
The current research investigated the potential use and efficiency of ratio of 80:20. Moreover, the sole NaOH solution yields a lowest
GBFS- and BOFS-treated kaolinite clay with various activators (e.g., UCS value for all samples, such that as the Na2SiO3:NaOH ratio
CaO and MgO) at different temperatures and curing times. Moreover, to increases to 80:20, the strength of the treated samples increases.
evaluate the effect of clay-slag based geopolymer on the behavior of the Therefore, the alkaline solution including sodium silicate and so-
soft clay, various Na2SiO3:NaOH ratios as alkaline solutions were added dium hydroxide can strongly improve the strength, and conse-
to the optimal amount (20%) of the mixtures. Throughout the study, the quently, stabilization using geopolymer prepared with the slag will
implications are as follows: be an efficient and cost-effective method to enhance the mechanical
properties of soft clay.
(1) Soft clays can be satisfactorily treated when GBFS and BOFS, as (4) Both modulus of Eu and E50 exhibit the same trend as found for the
waste materials, are used and are activated by CaO and MgO. Such compressive strength of the samples. The toughness and stiffness of
a technique can overcome the instability of the clays and enhance the BOFS-containing samples, particularly in the case of CB, were
the mechanical properties, especially in the case of samples pre- greater than that of GBFS-based samples.
pared by BOFS. (5) The MB- and CB-treated samples tend to achieve a higher pH value
(2) The addition of all the studied stabilizers, up to 20%, provides a and represent a better compressibility compared to the MG and CG
minor effect on the UCSs of clay samples at the temperature of treatment. These findings can be justified by the enhanced pozzo-
20 °C, and the curing time shows no remarkable effect in the lanic activity in the former compositions, as confirmed by the XRD
strength. Increasing temperature from 20 to 45 °C, however, can patterns and the SEM analyses.

11
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 14. SEM analysis and MIP results: a) Kaolinite clay SEM; b) kaolinite +20% CB SEM; c) kaolinite +20% CB MIP; d) kaolinite +20% MB SEM; e) kaolinite
+20% MB MIP; f) kaolinite +20% CG SEM; g) kaolinite +20% CG MIP; h) kaolinite +20% MG SEM; i) kaolinite +20% MG MIP.

12
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Fig. 14. (continued)

Funding ASTM, 2006. Annual book of ASTM standards vol 04.08. In: American Society for Testing
and Materials.
Aziz, I.H., Abdullah, M.M.A.B., Heah, C.-Y., Liew, Y.-M., 2019. Behaviour changes of
This research did not receive any specific grant from funding ground granulated blast furnace slag geopolymers at high temperature. Adv. Cem.
agencies in the public, commercial, or not-for-profit sectors. Res. 1–28.
Burciaga-Díaz, O., Betancourt-Castillo, I., 2018. Characterization of novel blast-furnace
slag cement pastes and mortars activated with a reactive mixture of MgO-NaOH.
References Cem. Concr. Res. 105, 54–63.
Celik, E., Nalbantoglu, Z., 2013. Effects of ground granulated blastfurnace slag (GGBS) on
Afrasiabian, A., Salimi, M., Movahedrad, M., Vakili, A.H., 2019. Assessing the impact of the swelling properties of lime-stabilized sulfate-bearing soils. Eng. Geol. 163, 20–25.
GBFS on mechanical behaviour and microstructure of soft clay. Int. J. Geotech. Eng. Cherian, C., Arnepalli, D.N., 2015. A critical Appraisal of the Role of Clay Mineralogy in
1–11. Lime Stabilization. Int. J. Geosynth. Gr. Eng. 1, 8. https://doi.org/10.1007/s40891-
Alazigha, D.P., Indraratna, B., Vinod, J.S., Ezeajugh, L.E., 2016. The Swelling Behaviour 015-0009-3.
Of Lignosulfonate-Treated Expansive Soil. Consoli, N.C., Bittar Marin, E.J., Quiñónez Samaniego, R.A., Heineck, K.S., Johann,
Arulrajah, A., Mohammadinia, A., Phummiphan, I., Horpibulsuk, S., Samingthong, W., A.D.R., 2018a. Use of Sustainable Binders in Soil Stabilization. J. Mater. Civ. Eng. 31,
2016. Stabilization of recycled demolition aggregates by geopolymers comprising 6018023.
calcium carbide residue, fly ash and slag precursors. Constr. Build. Mater. 114, Consoli, N.C., Quiñónez Samaniego, R.A., González, L.E., Bittar, E.J., Cuisinier, O., 2018b.
864–873. Impact of Severe climate Conditions on loss of Mass, Strength, and Stiffness of
Arulrajah, A., Kua, T.-A., Suksiripattanapong, C., Horpibulsuk, S., Shen, J.S., 2017a. Compacted Fine-Grained Soils–Portland Cement Blends. J. Mater. Civ. Eng. 30,
Compressive strength and microstructural properties of spent coffee grounds-bagasse 4018174.
ash based geopolymers with slag supplements. J. Clean. Prod. 162, 1491–1501. Coudert, E., Paris, M., Deneele, D., Russo, G., Tarantino, A., 2019. Use of alkali activated
Arulrajah, A., Mohammadinia, A., D’Amico, A., Horpibulsuk, S., 2017b. Effect of lime kiln high-calcium fly ash binder for kaolin clay soil stabilisation: Physicochemical evo-
dust as an alternative binder in the stabilization of construction and demolition lution. Constr. Build. Mater. 201, 539–552.
materials. Constr. Build. Mater. 152, 999–1007. Dash, S.K., Hussain, M., 2011. Lime stabilization of soils: reappraisal. J. Mater. Civ. Eng.
Arulrajah, A., Kua, T.-A., Suksiripattanapong, C., Horpibulsuk, S., 2019. Stiffness and 24, 707–714.
strength properties of spent coffee grounds-recycled glass geopolymers. Road Mater. Du, Y.J., Bo, Y.L., Jin, F., Liu, C.Y., 2016. Durability of reactive magnesia-activated slag-
Pavement Des. 20, 623–638. stabilized low plasticity clay subjected to drying-wetting cycle. Eur. J. Environ. Civ.

13
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Eng. 20, 215–230. https://doi.org/10.1080/19648189.2015.1030088. Mater. Civ. Eng. 30, 4018171.


Eltantawy, I.M., Arnold, P.W., 1973. Reappraisal of ethylene glycol mono ethyl ether Li, M., Chai, S., Du, H., Wang, C., 2016. Effect of chlorine salt on the physical and me-
(EGME) method for surface area estimations of clays. J. Soil Sci. 24, 232–238. chanical properties of inshore saline soil treated with lime. Soils Found. 56, 327–335.
Emmanuel, E., Lau, C.C., Anggraini, V., Pasbakhsh, P., 2019. Stabilization of a soft marine Mohammadinia, A., Arulrajah, A., Sanjayan, J., Disfani, M.M., Win Bo, M., Darmawan, S.,
clay using halloysite nanotubes: a multi-scale approach. Appl. Clay Sci. 173, 65–78. 2016. Stabilization of demolition materials for pavement base/subbase applications
EPA, 1983. Process Design Manual: Land Application Of Municipal Sludge. Res. Lab. EPA- using fly ash and slag geopolymers. J. Mater. Civ. Eng. 28, 4016033.
625/1–83-016. Mohammadinia, A., Arulrajah, A., D’Amico, A., Horpibulsuk, S., 2018. Alkali-activation
Estabragh, A.R., Beytolahpour, I., Javadi, A.A., 2010. Effect of resin on the strength of of fly ash and cement kiln dust mixtures for stabilization of demolition aggregates.
soil-cement mixture. J. Mater. Civ. Eng. 23, 969–976. Constr. Build. Mater. 186, 71–78.
Ge, L., Wang, C.-C., Hung, C.-W., Liao, W.-C., Zhao, H., 2018. Assessment of strength Mola-Abasi, H., Khajeh, A., Naderi Semsani, S., 2018a. Variables controlling tensile
development of slag cement stabilized kaolinite. Constr. Build. Mater. 184, 492–501. strength of stabilized sand with cement and zeolite. J. Adhes. Sci. Technol. 32,
Ghobadi, M.H., Abdilor, Y., Babazadeh, R., 2014. Stabilization of clay soils using lime and 947–962.
effect of pH variations on shear strength parameters. Bull. Eng. Geol. Environ. 73, Mola-Abasi, H., Khajeh, A., Naderi Semsani, S., 2018b. Effect of the Ratio between
611–619. https://doi.org/10.1007/s10064-013-0563-7. Porosity and SiO 2 and Al 2 O 3 on Tensile Strength of Zeolite-Cemented Sands. J.
Ghorbani, A., Hasanzadehshooiili, H., 2018. Prediction of UCS and CBR of microsilica- Mater. Civ. Eng. 30, 4018028.
lime stabilized sulfate silty sand using ANN and EPR models; application to the deep Mola-Abasi, H., Khajeh, A., Semsani, S.N.S., 2018c. Porosity/(SiO2 and Al2O3 Particles)
soil mixing. Soils Found. 58, 34–49. Ratio Controlling Compressive Strength of Zeolite-Cemented Sands. Geotech. Geol.
Ghorbani, A., Salimzadehshooiili, M., 2018. Evaluation of Strength behaviour of Cement- Eng. 1–10.
RHA Stabilized and Polypropylene Fiber Reinforced Clay-Sand Mixtures. Civ. Eng. J. Obuzor, G.N., Kinuthia, J.M., Robinson, R.B., 2012. Soil stabilisation with lime-activated-
4, 2628–2641. GGBS-A mitigation to flooding effects on road structural layers/embankments con-
Ghorbani, A., Salimzadehshooiili, M., 2019. Dynamic Characterization of Sand Stabilized structed on floodplains. Eng. Geol. 151, 112–119. https://doi.org/10.1016/j.enggeo.
with Cement and RHA and Reinforced with Polypropylene Fiber. J. Mater. Civ. Eng. 2012.09.010.
31, 4019095. Ouhadi, V.R., Yong, R.N., Sedighi, M., 2006. Desorption response and degradation of
Ghorbani, A., Hasanzadehshooiili, H., Karimi, M., Daghigh, Y., Medzvieckas, J., 2015. buffering capability of bentonite, subjected to heavy metal contaminants. Eng. Geol.
Stabilization of problematic silty sands using microsilica and lime. Balt. J. Road 85 (1-2), 102–110.
Bridg. Eng. 10. Pan, S.-Y., Adhikari, R., Chen, Y.-H., Li, P., Chiang, P.-C., 2016. Integrated and innovative
Ghorbani, A., Hasanzadehshooiili, H., Mohammadi, M., Sianati, F., Salimi, M., Sadowski, steel slag utilization for iron reclamation, green material production and CO2 fixation
L., Szymanowski, J., 2019. Effect of selected Nanospheres on the Mechanical Strength via accelerated carbonation. J. Clean. Prod. 137, 617–631.
of Lime-Stabilized High-Plasticity Clay Soils. Adv. Civ. Eng. 2019, 1–11. https://doi. Pasupathy, K., Berndt, M., Sanjayan, J., Rajeev, P., Cheema, D.S., 2018. Durability
org/10.1155/2019/4257530. Article ID 4257530. Performance of Precast Fly Ash–based Geopolymer Concrete under Atmospheric
Ghosh, D., Bhattacharyya, K.G., 2002. Adsorption of methylene blue on kaolinite. Appl. Exposure Conditions. J. Mater. Civ. Eng. 30, 4018007.
Clay Sci. 20, 295–300. Phanikumar, B.R., Nagaraju, T.V., 2018. Swell and Compressibility of GGBS–Clay Mixes
Goodarzi, A.R., Movahedrad, M., 2017. Stabilization/solidification of zinc-contaminated in Lumps and Powders: effect of 4% Lime. Indian Geotech. J. 1–9.
kaolin clay using ground granulated blast-furnace slag and different types of acti- Phetchuay, C., Horpibulsuk, S., Arulrajah, A., Suksiripattanapong, C., Udomchai, A.,
vators. Appl. Geochem. 81, 155–165. 2016. Strength development in soft marine clay stabilized by fly ash and calcium
Goodarzi, A.R., Salimi, M., 2015a. Stabilization treatment of a dispersive clayey soil using carbide residue based geopolymer. Appl. Clay Sci. 127, 134–142.
granulated blast furnace slag and basic oxygen furnace slag. Appl. Clay Sci. 108, Phummiphan, I., Horpibulsuk, S., Rachan, R., Arulrajah, A., Shen, S.-L., Chindaprasirt, P.,
61–69. 2018. High calcium fly ash geopolymer stabilized lateritic soil and granulated blast
Goodarzi, A.R., Salimi, M., 2015b. Effect of iron industry slags on the geotechnical furnace slag blends as a pavement base material. J. Hazard. Mater. 341, 257–267.
properties and mineralogy characteristics of expansive clayey soils. Modares. J. Civ. Poltue, T., Suddeepong, A., Horpibulsuk, S., Samingthong, W., Arulrajah, A., Rashid,
Eng. 15, 161–203. A.S.A., 2019. Strength development of recycled concrete aggregate stabilized with fly
Goodarzi, A.R., Akbari, H.R., Salimi, M., 2016. Enhanced stabilization of highly expansive ash-rice husk ash based geopolymer as pavement base material. Road Mater.
clays by mixing cement and silica fume. Appl. Clay Sci. 132–133, 675–684. https:// Pavement Des. 1–12.
doi.org/10.1016/j.clay.2016.08.023. Rattanasak, U., Chindaprasirt, P., 2009. Influence of NaOH solution on the synthesis of fly
Gu, K., Jin, F., Al-Tabbaa, A., Shi, B., Liu, C., Gao, L., 2015. Incorporation of reactive ash geopolymer. Miner. Eng. 22, 1073–1078.
magnesia and quicklime in sustainable binders for soil stabilisation. Eng. Geol. 195, Salimi, M., Ilkhani, M., Vakili, A.H., 2018. Stabilization treatment of Na-montmorillonite
53–62. with binary mixtures of lime and steelmaking slag. Int. J. Geotech. Eng. 1–7. https://
Hamidi, S., Marandi, S.M., 2018. Clay concrete and effect of clay minerals types on sta- doi.org/10.1080/19386362.2018.1439294.
bilized soft clay soils by epoxy resin. Appl. Clay Sci. 151, 92–101. Saride, S., Puppala, A.J., Chikyala, S.R., 2013. Swell-shrink and strength behaviors of lime
Hendershot, W.H., Duquette, M., 1986. A simple barium chloride method for determining and cement stabilized expansive organic clays. Appl. Clay Sci. 85, 39–45.
cation exchange capacity and exchangeable cations 1. Soil Sci. Soc. Am. J. 50, Sekhar, D.C., Nayak, S., 2018. Utilization of granulated blast furnace slag and cement in
605–608. the manufacture of compressed stabilized earth blocks. Constr. Build. Mater. 166,
Hoy, M., Horpibulsuk, S., Arulrajah, A., 2016. Strength development of Recycled Asphalt 531–536.
Pavement–Fly ash geopolymer as a road construction material. Constr. Build. Mater. Sharma, A.K., Sivapullaiah, P.V., 2016. Ground granulated blast furnace slag amended fly
117, 209–219. ash as an expansive soil stabilizer. Soils Found. 56, 205–212.
Hoy, M., Horpibulsuk, S., Arulrajah, A., Mohajerani, A., 2018. Strength and micro- Siddiqua, S., Barreto, P.N.M., 2018. Chemical stabilization of rammed earth using cal-
structural study of recycled asphalt pavement: Slag geopolymer as a pavement base cium carbide residue and fly ash. Constr. Build. Mater. 169, 364–371.
material. J. Mater. Civ. Eng. 30, 4018177. Soltani, A., Deng, A., Taheri, A., Mirzababaei, M., 2018. Rubber powder–polymer com-
Huang, G., Ji, Y., Li, J., Hou, Z., Dong, Z., 2018. Improving strength of calcinated coal bined stabilization of South Australian expansive soils. Geosynth. Int. 25, 304–321.
gangue geopolymer mortars via increasing calcium content. Constr. Build. Mater. Soltani, A., Deng, A., Taheri, A., Mirzababaei, M., 2019. A sulphonated oil for stabilisation
166, 760–768. of expansive soils. Int. J. Pavement Eng. 20, 1285–1298.
Jahandari, S., Saberian, M., Zivari, F., Li, J., Ghasemi, M., Vali, R., 2019. Experimental Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., Chindaprasirt, P., 2011. NaOH-acti-
study of the effects of curing time on geotechnical properties of stabilized clay with vated ground fly ash geopolymer cured at ambient temperature. Fuel 90, 2118–2124.
lime and geogrid. Int. J. Geotech. Eng. 13, 172–183. Sukmak, P., Horpibulsuk, S., Shen, S.-L., 2013. Strength development in clay–fly ash
Jin, F., Gu, K., Abdollahzadeh, A., Al-Tabbaa, A., 2013. Effects of different reactive MgOs geopolymer. Constr. Build. Mater. 40, 566–574.
on the hydration of MgO-activated GGBS paste. J. Mater. Civ. Eng. 27, B4014001. Sukmak, P., Sukmak, G., Horpibulsuk, S., Setkit, M., Kassawat, S., Arulrajah, A., 2019.
Jin, F., Gu, K., Al-Tabbaa, A., 2015. Strength and hydration properties of reactive MgO- Palm oil fuel ash-soft soil geopolymer for subgrade applications: strength and mi-
activated ground granulated blastfurnace slag paste. Cem. Concr. Compos. 57, 8–16. crostructural evaluation. Road Mater. Pavement Des. 20, 110–131.
Khajeh, A., Chenari, R.J., Payan, M., 2019. A simple Review of Cemented Non-conven- Tsuchida, T., Porbaha, A., Yamane, N., 2001. Development of a geomaterial from dredged
tional Materials: Soil Composites. Geotech. Geol. Eng. 1–22. bay mud. J. Mater. Civ. Eng. 13, 152–160.
Khan, M.Z.N., Hao, Y., Hao, H., et al., 2016. Synthesis of high strength ambient cured Vakili, A.H., Ghasemi, J., bin Selamat, M.R., Salimi, M., Farhadi, M.S., 2018. Internal
geopolymer composite by using low calcium fly ash. Constr. Build. Mater. 125, erosional behaviour of dispersive clay stabilized with lignosulfonate and reinforced
809–820. with polypropylene fiber. Constr. Build. Mater. 193, 405–415. https://doi.org/10.
Kim, S., Palomino, A.M., 2009. Polyacrylamide-treated kaolin: a fabric study. Appl. Clay 1016/j.conbuildmat.2018.10.213.
Sci. 45, 270–279. Wang, Y., Cui, Y.-J., Benahmed, N., Tang, A.M., Duc, M., 2019. Changes of small strain
Kua, T.-A., Arulrajah, A., Horpibulsuk, S., Du, Y.-J., Shen, S.-L., 2016. Strength assessment shear modulus and suction for a lime-treated silt during curing. Géotechnique 1–19.
of spent coffee grounds-geopolymer cement utilizing slag and fly ash precursors. Webster, A.J., Groffman, P.M., Cadenasso, M.L., 2018. Controls on denitrification po-
Constr. Build. Mater. 115, 565–575. tential in nitrate rich waterways and riparian zones of an irrigated agricultural set-
Latifi, N., Marto, A., Eisazadeh, A., 2015. Analysis of strength development in non-tra- ting. Ecol. Appl. 28, 1055–1067.
ditional liquid additive-stabilized laterite soil from macro-and micro-structural con- Yadu, L., Tripathi, R.K., 2013. Stabilization of soft soil with granulated blast furnace slag
siderations. Environ. Earth Sci. 73, 1133–1141. and fly ash. Int. J. Res. Eng. Technol. 2, 115–119.
Latifi, N., Horpibulsuk, S., Meehan, C.L., Abd Majid, M.Z., Tahir, M.M., Mohamad, E.T., Yaghoubi, M., Arulrajah, A., Disfani, M.M., Horpibulsuk, S., Darmawan, S., Wang, J.,
2016. Improvement of problematic soils with biopolymer—an environmentally 2019. Impact of field conditions on the strength development of a geopolymer sta-
friendly soil stabilizer. J. Mater. Civ. Eng. 29, 4016204. bilized marine clay. Appl. Clay Sci. 167, 33–42.
Leong, H.Y., Ong, D.E.L., Sanjayan, J.G., Nazari, A., 2018. Strength Development of Yi, Y., Liska, M., Al-Tabbaa, A., 2014a. Properties and microstructure of GGBS–magnesia
Soil–Fly Ash Geopolymer: Assessment of Soil, Fly Ash, Alkali Activators, and Water. J. pastes. Adv. Cem. Res. 26, 114–122.

14
M. Salimi and A. Ghorbani Applied Clay Science 184 (2020) 105390

Yi, Y., Liska, M., Al-Tabbaa, A., 2014b. Properties of two Model Soils Stabilized with ground blast-furnace slag. Constr. Build. Mater. 52, 504–510.
Different Blends and Contents of GGBS, MgO, Lime, and PC. J. Mater. Civ. Eng. 26, Zhang, J., Soltani, A., Deng, A., Jaksa, M.B., 2019. Mechanical performance of jute fiber-
267–274. https://doi.org/10.1061/(ASCE)MT.1943-5533.0000806. reinforced micaceous clay composites treated with ground-granulated blast-furnace
Yi, Y., Gu, L., Liu, S., 2015. Microstructural and mechanical properties of marine soft clay slag. Materials (Basel). 12, 576.
stabilized by lime-activated ground granulated blastfurnace slag. Appl. Clay Sci. 103, Zhao, Y., Soltani, A., Taheri, A., Karakus, M., Deng, A., 2019. Application of slag–cement
71–76. and fly ash for strength development in cemented paste backfills. Minerals 9, 22.
Yi, Y., Gu, L., Liu, S., Jin, F., 2016. Magnesia reactivity on activating efficacy for ground Zheng, J., Sun, X., Guo, L., Zhang, S., Chen, J., 2019. Strength and hydration products of
granulated blastfurnace slag for soft clay stabilisation. Appl. Clay Sci. 126, 57–62. cemented paste backfill from sulphide-rich tailings using reactive MgO-activated slag
https://doi.org/10.1016/j.clay.2016.02.033. as a binder. Constr. Build. Mater. 203, 111–119.
Yusuf, M.O., Johari, M.A.M., Ahmad, Z.A., Maslehuddin, M., 2014. Strength and micro- Zukri, A., 2013. Pekan soft clay treated with hydrated lime as a method of soil stabilizer.
structure of alkali-activated binary blended binder containing palm oil fuel ash and Procedia Eng. 53, 37–41.

15

You might also like