You are on page 1of 16

Case Studies in Construction Materials 16 (2022) e00988

Contents lists available at ScienceDirect

Case Studies in Construction Materials


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

Strength, Durability, and Microstructures characterization of


sustainable geopolymer improved clayey soil
Noor Aamer Odeh a, Alaa H.J. Al-Rkaby b, *
a
MSc Candidate/ Thi Qar University, College of Engineering, Iraq
b
Asst. Professor/ Thi Qar University, College of Engineering, Iraq

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

Keywords: Geopolymer has been emerged as an innovative and eco-friendly alternative to traditional soil
Geopolymer improvement products such as Ordinary Portland Cement (OPC) and lime, which have negative
Clay soil environmental consequences. Geopolymer has been employed in various geotechnical applica­
Coal-fired fly ash
tions as alternative lightweight backfill in highway barriers and/or behind retaining walls,
Sulfuric acid
Sodium chloride
providing technical, environmental, and economic benefits. The present study aims to evaluate
Microstructure the mechanical, durability, and Microstructure properties of clayey soil (with different percent­
ages of sand) stabilized using coal-fired fly and a sodium hydroxide/sodium silicate solution alkali
activator. Stabilized soils exhibited high strength of 1–8 MPa compared with less than 0.3 MPa for
untreated soils. Moreover, adding sand to clay soil reduced the optimum ratio of alkaline acti­
vator from 0.8 for clayey soil with 5% sand to 0.6 for clayey soil with 10, 20, 30% sand. The
geopolymer treated samples also showed high resistance to the acidic environments and chlo­
rides. Finally, SEM results revealed a clay fabric modification caused by inter-particle contacts
and the resulting bonding caused by gel formation and hardening. According to the outcomes of
this study, it can be concluded that using a fly ash geopolymer binder for soil stabilization is a
viable alternative to cement in geotechnical applications.

1. Introduction

Chemical stabilization is a technique used to enhance soft soil’s durability and strength, and it has considerable economic and
engineering advantages [1]. Different chemical agents such as lime, cement, pozzolana, or industrial waste (fly ash (FA) and ground
granulated blast furnace slag (GGBFS)) are used for stabilization. Lime and (OPC) are the most common binders for chemical soil
stabilization [2]. However, a significant environmental concern of massive carbon emissions is manufacturing cement and lime. For
example, the cement industry contributes about 7% of (CO2) emissions.
Moreover, producing one ton of OPC generates approximately 0.8–1.0 tons of (CO2) into the atmosphere [3]. Consequently, there
has been a significant emphasis on cement alternatives in research. One of the substitutes is the alkali-activated binder that produces
60–80% less CO2 and uses 60% less energy during production than OPC [4]. Davidovits (1994) refers to these binders as "geo­
polymers", consist of solid aluminosilicate precursors and an alkaline activator produced from waste or byproducts possessing high
amorphous Si and Al resources such as fly ash, (GGBFS), rice husk ask (RHA), and metakaolin (MK) [5–8]. In brief, geopolymerization
is a quick chemical process that includes four stages: dissolution of Si Al and ions, ion diffusion, gel formation by polycondensation of

* Corresponding author.
E-mail addresses: alaa.astm@gmail.com, alaa.al-rakaby@utq.edu.iq (A.H.J. Al-Rkaby).

https://doi.org/10.1016/j.cscm.2022.e00988
Received 18 November 2021; Received in revised form 16 February 2022; Accepted 25 February 2022
Available online 28 February 2022
2214-5095/© 2022 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 1. The Particle size distribution curve of soils.

Al and Si complexes with an alkaline activator, and gel hardening, which results in a hardened geopolymeric product [9].
Geopolymers are a potential future alternative to conventional OPC as a sustainable building material because of many benefits
such as good early strength [11], rapid hardness [12], low shrinkage [10], low permeability [13], automated humidity and tem­
perature, high resistance to chemical corrosion [14], and fire resistance[15]. There are some disadvantages of geopolymer materials,
such as the loss of workability due to the alkaline solution [16,17], the harmful nature of the materials used in these solutions [18,19],
also problems such as the high alkalinity of the material [20,21].
During the last few years, some studies have been done on the applications of geopolymers as soil stabilizers [4,9,10,22–25].
Geopolymer binder assisted soil particles in forming a more dense microstructure, enhancing the volume stability and mechanical
properties of the soil [10]. Therefore, it is possible to employ the geopolymer as a shallow-depth soil stabilizer (e.g., base or subbase in
the pavement, embankment, shallow foundation, airport construction, etc.) and deep soil mixing applications [23,26].
Researchers focused on developing geopolymers from various aluminum silicate sources, such as metakaolin, fly ash, blast furnace
slag, rice husk ash [27–30]. Cristelo et al. [31] investigated the application of alkaline activation of fly ash to stabilize soft soils. They
found that activated fly ash could be used to substitute cement for soil stabilization. Unconfined compressive strength (UCS) of cement
and geopolymer samples cured for 28 days were similar. Similar behavior was recently observed by [24]. Zhu et al. [30] reported that
metakaolin-based geopolymer stabilization was an effective method for treating clayey soil as UCS increases with the increase in the
geopolymer percentage. Stabilizing by geopolymer also increases the resilient modulus, particularly by increasing the amount of alkali
activator [33,34]. Kang et al. [33] investigated the different fly ash-geopolymer mixture’s resilience modulus and permanent defor­
mation. It was observed that soil with various fly ash additives had similar resilience modulus even though they had different per­
manent deformation.
In terms of shear strength characteristics, the internal friction angles for geopolymer stabilized soils were higher than 50◦ , while the
cohesions were more significant than 250 kPa [4]. Similar results were obtained by [22]. It is reported that the mixing soil with
geopolymer results in changing the initial properties of remolded clays from quasi-over-consolidated to remarkably over-consolidated,
increasing the surface of good yields and shear strength parameters [24]. SEM images have been used to study the interaction between
soil particles and industrial waste, the response of the geopolymer to the changes in temperature, and the percentage and concen­
tration of the geopolymer. The microstructural of soil stabilized by geopolymer indicates that the dense structures of calcium
aluminate hydrate (C-A-H) and calcium silicate hydrate (C-S-H) are the reason for the increased strength [26]. Until far, research on fly
ash geopolymer has relied on a precursor produced from class F fly ash. Despite their energy efficiency and environmental friendliness,
geopolymers based on Class F fly ash have practical issues such as curing temperature and the use of a high activator content, which
currently restrict their use in soil stabilization in the site.
It is observed that geopolymers may have superior mechanical characteristics. However, little attention has been paid to the long-
term durability of geopolymer-stabilized soils. There is no clear indication in the research for the ideal quantity of geopolymer that can
provide the durability as specified for OPC- soils. A lack of extensive study on the durability performance of geopolymer-stabilized soil
is one of the major factors restricting the broad implementation of this promising approach for ground improvement. There is little
information on the wetting-drying durabilty achieving sustainability of geopolymer soil mixtures. Among the limited research,
Adhikari et al. [29] reported that geopolymer-treated soil outperforms cement-treated soil in dry and wet durability testing. Similar
behavior was recently observed by Sreelakshmi et al. [32] reported that geopolymer samples passed twelve wet and dry cycles
exhibited minimal deformation. Moreover, geopolymer enhanced soil resilience to weather changes substantially. Moreover, there is
no literature on the effects of geopolymer on the performance of treated soil and strength performance after exposure to acids and
chlorides. the studies of the effect of sulfuric acid on soils is important after the spread of manufactories and air pollution with SO2 gas,
which turns into sulfuric acid by oxidation and then interaction with water [35]. Gaseous, solid, and liquid emissions emitted by
industrial plants are closely related to environmental pollution, but the pollution may reach the ground and groundwater. The problem
is complicated by the arrival of the effects of pollution to the bearing foundations of the factory facilities, and they work to weaken
them and sometimes lead to collapse as a result of cavities. The current study investigates the possibility of using (high-calcium Class C
fly ash) -based geopolymer to stabilize the clay soil at ambient temperature. This study investigates the effects of fly ash ratio, activator

2
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Table 1
The physical properties of clay soils.
Soil property Standard value

Clay Soil-1 Soil- 2 Soil- 3

Liquid Limit (LL)% ASTM D 4318 53.5 51.1 48.5 46.9


Plastic Limit (PL)% 24.6 23 22.6 21.4
Plasticity index (PI) % 28.9 28.1 25.9 25.5
Sand content % ASTM D 422 5 10 20 30
Clay content % 56.39 52.87 46.99 41.12
Silt content % 38.61 37.13 33.01 28.88
%Passing sieve no. 200 95 90 80 70
Soil classification ASTM D 2487 CH CH CL CL
Optimum moisture content% ASTMD 1557 14.2 13.36 12.9 11.6
Max dry density (gm/cm3) 1.761 1.774 1.812 1.836

Table 2
Chemical compositions of the fly ash and clay soil using EDS.
Elt Concentration by % weight

C O F Na Mg Al Si S K Ca Fe

Clay soil 12.17 41.7 0 1.74 5.31 9.91 18.09 1.82 7.69 1.71 1.46
Fly ash 1.01 30.71 0 1.72 2.5 14.18 26.65 1.3 1.5 17.11 6.09

Fig. 2. The Particle size distribution curve of fly ash.

Fig. 3. SEM image of (a)clay soil (b) fly ash particles.

3
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Table 3
Chemical composition and ratios of elements in the clay-geopolymer specimens.
Sample no. Symbol FA% AC/FA UCS (MPa) Concentration by weight %

Si/Al Na Fe

1 untreated … … 0.22 1.82 1.74 1.4


2 FA15A0.6 15 0.8 4.08 2.09 5.66 2.12
3 FA20A0.6 20 0.6 5.18 2.45 5.42 2.9
4 FA20A0.8 20 0.8 6.4 2.6 6.82 2.3

to fly ash ratio, and clay content. Moreover, The durability of geopolymer samples was evaluated by it immersing in sulfuric acids and
sodium chloride. In order to better understanding the function of variables influencing strength development, SEM analysis was used to
examine the microstructural development of soil-FA geopolymer.

2. Experimental program

2.1. Materials

2.1.1. Soils
The soil used in the current study is high plastic clayey (CH) with sand content (5%). The disturbed sample was taken at a depth of
approximately 1–1.5 m. The grain size distribution of soil is shown in Fig. 1. Table 1 displays some properties of such soil. The
elemental analysis using energy dispersive spectroscopy (EDS) is also shown in Table 2. Scanning electron microscopy (SEM)
demonstrated that particles had an irregular shape, as illustrated in Fig. 3. (a). To study the effect of geopolymer on such clay soil that
has different content of sand, the clay mentioned above was mixed with varying proportions of sand (5%, 15%, and 25%). And thus,
three different soils were prepared, soil- 1, soil − 2, and soil − 3, respectively. The gradient curve of these soils and the physical
parameters are shown in Fig. 1 and Table 2.

2.1.2. Coal-fired fly ash


Fly ash (FA) was supplied from local coal-fired electrical generating plants. Its properties comply with high calcium fly ash (Class C)
in ASTM standard C618. The particle distribution analyzed through the hydrometer test is shown in Fig. 2. The average particle size is
0.03 mm. In addition, chemical compositions obtained by EDS of fly ash are presented Table 3. Fig. 3
(b) shows high-resolution pictures of fly ash obtained using SEM. The micrograph shows that the fly ash comprises different particle
sizes with spherical shapes.

2.2. Alkali activators

A solution of sodium silicate (Na2SiO3) and sodium hydroxide (NaOH) in powder form was employed in this research. The sodium-
based alkaline activator solution was chosen because it was less expensive and readily available than potassium-based solutions.
Furthermore, NaOH has been shown to have a more excellent capability for liberating silicate and aluminate monomers. Before
combining with Sodium silicate, NaOH was dissolved in distilled water for at least 24 h to obtain the solution’s necessary Molar (M)
concentration. Throughout the research, the molarity of the NaOH solution remained constant at 10 M. To produce a 10 M solution,
400 g of NaOH pellets were dissolved in one liter of distilled water using molarity calculations (10 *40 = 400 g, where 40 denotes
NaOH’s molecular weight). The mass ratio of Na2SiO3/NaOH was equal to 2.0 to generate a high alkaline environment and maximize
the early strength.

2.3. Samples Preparation

To prepare samples for all the tests, three different proportions (10, 15, 20%) of the source material (FA) were mixed as a partial
substitute (by weight) with dry soil for 5 min to ensure mixture uniformity. Alkali activator was then prepared by mixing adequate
amounts of NaOH and Na2SiO3 based on the mixtures’ alkaline ratios for 5 min and then left at room temperature for another
5–10 min. The other step was mixing the alkaline solution in different proportions (AC/FA(0.4,0.6,0.8)) with free water at the required
liquid content and then added gradually to the dry mixture. The various ingredients were mixed for an additional 3–5 min until a
homogeneous mixture was obtained. The mixture was then compacted in cylindrical mold) PVC(, 45 mm in diameter and 90 in height
mm. The compaction was achieved in controlled weight/thickness layers for each specimen to produce the desired density. After the
compaction, the sample of geopolymers was maintained for 24 h before immersing in water for curing. The period of 28 days was
chosen as an average curing time.

2.4. Tests conducted

2.4.1. Unconfined compressive strength


To investigate the sample’s strength development, unconfined compressive strength (UCS) tests were performed at a curing time of

4
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 4. UCS test device.

Fig. 5. samples treated with acid and chlorides.

28 days. The UCS tests samples were prepared using cylindrical split tubes from (PVC) of 45 mm diameter and 90 mm height, Whereas
an aspect ratio of height-to-diameter of 2:1 was achieved. According to (ASTM D1633-00, [36]), procedures performed the
Compressive strength test for soil samples stabilized with geopolymer using the uniaxial machine, with a loading capacity of 50 kN.
The applied load and resultant displacements were determined using a load cell and a Linear Variable Displacement Transducer
(LVDT), respectively, as shown in Fig. 4. All UCS testing was carried out using a displacement rate of 0.1 mm per minute.

5
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 6. Effect of fly ash content on the UCS of different soil-geopolymer mixtures.

Fig. 7. Strength development index of geopolymer soils at different fly ash ratio and AC/FA 0.4.

2.4.2. Microstructural analysis


Microstructural characteristics were studied using SEM/EDS for pieces resulting from fractured samples tested by the uniaxial
machine were used to examine the microstructural formation of gel structure and the modification in soil fabric during stabilization.

2.4.3. Durability
For the durability test, after 28 days of curing, the soil-geopolymer samples were submerged in a 5% wt of NaCl solution and 5% vol
H2SO4 solution for 28 days, as shown in Fig. 5. Then, the unconfined compressive strength (UCS) of such exposed samples was
investigated.

6
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 8. The influence of the alkaline ratio (AC/FA) on the UCS of soil-geopolymer.

3. Results and discussion

3.1. Unconfined compressive strength (UCS)

3.1.1. Effect of fly ash content


The influence of various amounts of fly ash as the primary source material on the strength of soil-geopolymer mixtures was studied
by considering different proportions of fly ash FA (10%, 15%, and 20%) and activated to fly ash ratio AC/FA (0.4, 0.6, and 0.8). From
Fig. 6, It has been shown that increasing the fly ash content led in a significant difference in the unconfined compression strength (UCS)
values for the geopolymer- soil for all activator ratios. Such an increase can be attributed to more geopolymer binders in the structures.
The difference in strength between treated and untreated soil divided by the strength of untreated soil was used in this part and
termed as the strength improvement index SII. Fig. 7 illustrates the SII at AC/FA of 0.4. For all soils, the rise in SII is more evident with
increasing fly-ah concentration. For example, SII were 2.27, 2.38, 2.46, and 2.77 for clay soil, soil-1,soil-2, and soil-3, respectively, for
10% fly ash. When the fly-ash percentage was raised to 20%, the SII increased dramatically to 6.8, 7.7, and 7.5 for clay soil, soil-1, soil-
2, and soil-3, respectively. The behavior is consistent with the results mentioned in the literature [9,23,37]. Such improvement is
mainly linked to the cementitious properties of activated fly ash through the generated geopolymer gel.
Based on the laboratory test results, there is no optimum percentage of fly ash. The UCS increases with the increase of the rate of fly
ash for all tested soils. Similar findings were reported by Adhikari et al. [29],The UCS increased by increasing the fly ash to 25% and
30% for clay soils. Sreelakshmi et al. [32] reported that the optimum fly ash ratio is 30% for weak soil.

3.1.2. Effect of alkaline ratio


Many variables influence of the strength of the fly ash- geopolymer samples; one of them is the alkaline activator. This study’s
alkaline activator is a mixture of sodium silicate solution (Na2SiO3) and sodium hydroxide solution (NaOH). The sodium hydroxide
(NaOH) solution leaches the silicon and aluminum from the amorphous phase of FA, while the sodium silicate solution (Na2SiO3)
serves as a binder. The optimal activator ratios for the clay–geopolymer specimens are different to achieve maximum strengths. Fig. 8
illustrates the alkaline ratio’s effect on the UCS of clay–FA geopolymer specimens for varying fly ash content and AC/FA ratios. It can
be observed (Fig. 8, a) that the UCS of all mixtures increased with increasing the alkaline ratio (from 0.4 to 0.8) and fly ash content in

7
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 9. Strength improvement index of geopolymer soils at (20% FA).

Fig. 10. Typical stress-strain behavior of soils treated with AC/FA 0.6.

clay soil(5% sand). The UCS at FA 20% increased from 2.9 to 5.1 then 6.4 MPa when the AC/FA increased from 0.4 to 0.6,0.8. The
reason is that the increase in the activator content led to an increase in the leaching of silicon and aluminum from the amorphous phase
of the fly ash. Based on the laboratory test results, the optimum alkaline activator ratio for clay soil is 0.8. This result is similar to that
made by [38].
On the other hand, they were increasing sand content from 5% to 10%, 20%, and 30% in soil-1, soil 2, and soil 3, respectively. There
was a significant increase in UCS with an increase in the alkalinity from o.4 to 0.6 for different proportions of fly ash, as shown in Fig. 8
(b, c, d). However, with a further increase of the alkalinity to 0.8, an apparent decrease in strength was observed. This decrease in
strength is attributed to the high activator content. The high silica (silica generated from the increase in the amount of sand, silica
produced from fly ash, and increased alkaline activator) led to less amount of OH− , and this will reduce the efficiency of the

8
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 11. SEM-EDS analysis for untreated clay soil a) SEM; b) EDS.

geopolymerization process.
Similar behavior was reported by [39], who mentioned that the high increase in the silicon content Si could negatively affect the
solubility of the fly ash particles, resulting in a lower strength. Fig. 9 shows the comparison results between the strength improvement
index of three soil at 20% of fly ash. It is clear that for mixtures containing AC/FA= 0.6, SII is superior (15.01,15.88,15.98) for soil- 1,
soil − 2, and soil- 3, respectively. When the percentage of the alkaline activator was increased from 0.6 to 0.8, an SII decreased from
(15.01,15.88,15.98) to (14.3, 13.1, 11.2) at the soil- 1, soil − 2, and soil − 3, respectively. It can also note that the value of the SII
decreased in soil- 3 is more than those of soil − 2 and soil − 1. The increase in the proportion of sand, which causes a higher proportion
of silicon, resulted in higher decrease in the SII.

3.1.3. Stress-strain behavior of geopolymer-treated clays


The stress-strain responses of untreated and geopolymer treated soil specimens, at fly ash contents of 10%,15%, and 20% and fly
ash to activator ratio fixed are 0.6, for clay soil,soil-1,soil-2, and soil-3 are displayed in Fig. 10 a, b, c, and d, respectively. In general
irrespective of the sand content, it was found that the geopolymer-treated soils exhibited brittle yielding, with the stress reaching the
peak before an abrupt failure. As the fly ash ratio increases from 10% to 15% and 20%, yielding was associated with a stiffer response
(i.e., higher unconfined strength and low axial strain), Similar behavior to the results presented by [24,40]. Although all soils (clay,
soil-1, soil-2, soil-3) exhibited a typical brittle stress-strain response, these soils varied in the axial strain and peak stress, particularly
for 20% geopolymer. For example, in Fig. 10 c,d, fly ash had shown maximum stress without post-peak response at 20%, indicating an
extremely brittle response.
In all cases, both quantitative and qualitative changes in the stress-strain response of the treated clay can be attributed to the
formation of cementitious products. Another possible explanation for the disparity in stress-strain behavior between the variously
treated clay soils is a difference in plasticity, as shown in Table 1, affecting the ability of the soil to mix with the geopolymer com­
ponents and produce a homogeneous matrix-free of agglomerates. This, in turn, impacts the geopolymer soil’s strength.

3.2. Microstructural analysis

The composition of the fly ash-based geopolymer is mainly derived from the decomposition of aluminum silicate in the fly ash by
alkaline solutions, which is generated by polycondensation. When an alkaline activator comes into contact with a source of alumi­
nosilicate (FA), it breaks the bonds of aluminosilicate in fly ash to liberation active Si4+ and AL3+ . These active Si4+ and AL3+ compose
nuclei and aluminosilicate oligomers form AlO4 and SiO 4 tetrahedral, aluminum silicate chains highly depend on the ratio Si/Al. In
other words, the geopolymer product (N-A-S-H gel) generates from the leaching Si4+ and AL3+ of reactions fly ash and activator, which
harden with time and cement the soil particles [41,42]. The rate of geopolymerization of fly ash can be observed in geopolymer - fly ash
by etching on fly ash surfaces, detected by SEM analysis [43]. SEM /EDS analysis was performed on untreated and geopolymer -soil
samples considering various fly ash and alkaline solution percentages. Fig. 11 shows the SEM image and EDS spectra of untreated clay
soil. The SEM image indicates isolated soil particles, which have a weak and flaky structure with some arrangement among them.
Fig. 12 shows the Microstructure of the geopolymer treated soil samples with different fly ash and AC/FA ratios. The cementitious
products on the fly ash surfaces are observed, indicating geopolymerization reaction and accentual the development of UCS. The
etched holes in fly ash surfaces caused by the decomposition of silica and aluminum are mostly filled with smaller particles and
cementitious products, resulting in a dense matrix. They caused a high strength of the treated soils, as shown in Fig. 12 c and 13. When
comparing Fig. 12(a) at (FA 15%, AC/FA 0.6) with Fig. 12 (b, c) at (FA20%, AC/FA 0.6 and 0.8), it can notice some partially reacted fly

9
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 12. SEM-EDS analysis for clay soil-geopolymer treated samples with different fly ash and alkaline solution ratio: a) FA 15, AC/FA 0.6; b) FA 20,
AC/FA 0.6; c) FA 20, AC/FA 0.8 sample.

ash particles in (FA15, AC/FA 0.6) sample, confirming that the geopolymerization reaction is less than (FA20%, AC/FA 0.6 and 0.8).
Increasing the percentage of fly ash and alkaline substances corresponding with more sodium hydroxide NaOH and sodium silicate
Na2SiO3 resulted in the high rate of geopolymerzation in samples with 20% fly ash, as shown in Fig. 12(b, c).
The EDS analysis showed the geopolymer samples contain higher peaks and a higher proportion of O, Si, Al, Fe, and Na than
untreated clay, indicating the presence of N-A-S-H, the main cementing product responsible for strength gain.
Table 3 shows significant compositions and proportions of the mixtures at 28 days. The primary ratio of Si/Al is widely considered

10
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 13. SEM of geopolymer treated soil: Geopolymer-cementitious bond.

when the EDS technique is applied. In general, the elemental ratios of OPC and geopolymer are 2.00–3.00 for Si/Al. Enhancing the Si/
Al ratio promotes the creation of more dense, compact, and homogenous microstructures. Geopolymer product with a ratio SiAl = 2.6
(as sample F20A0.8) is more homogeneous than the ratio of SiAl = 2.09, 2.54(as samples F15A0.6,F20A0.6). This might be owing to
the insoluble particles of fly ash. The insoluble particles produced interface connections with the binder, a sensitive location that
displays failure in compressive strength testing. Thus, the higher ratio Si / Al caused the formation of Si-O-Si structure, and silicate
species are more stable. This stability is generated by a subsequent geopolymerization process that results in a more complicated
network and homogenous geopolymer, resulting in increased strength.
With less amount of calcium content, the leading geopolymeric gel was an interfered of (Na)-poly(sialate-disiloxo) or Nan-(-Si-O-
Al-O-Si-O-Si-O-)n- [44]. The Na+ ion (from alkaline activators) and Ca2+ ion (dissociated from CaO in fly ash) were able to link with
central Al- or O - (dissociated from Al2O3 in fly ash), producing N-A-S-H or a combination of (C, N)-A-S-H, which provides early
strength to geopolymer mixes. Typically, Hydrated Portland cement’s principal product is an aluminate-substituted calcium silicate
hydrated (C-A-S-H) gel. In contrast, the principal impact of geopolymers is sodium alumina-silicate hydrated (N-A-S-H) gel for fly
ash-based geopolymers and (C-A-S-H) gel for slag-based geopolymers [45]. With the contribution of calcium source from high calcium
fly ash in the system, the formation of calcium-sodium alumina-silicate hydrated (C, N)-A-S-H gel was therefore promoted, as shown in
EDS analysis of the (FA20A 0.8) sample (see Fig. 12c).
Furthermore, Table 3 shows that the concentration of Fe in geopolymer samples ranges between (2.12–2.3), which may result in the
formation of Na-polyferrosialate (Si-O-(Fe)Al) binder phases, as previously reported by [21,46], as a result of the dissolution of the
reactive process (containing Al, Si, and Fe compounds), which allowed for the improvement of a dense structure, and improved
mechanical characteristics. This showed better performance compared with other types of stabilizers [47–49]. Fig. 13.

3.3. Durability (acid and salt resistance)

Depending on the results obtained previously from the unconfined compressive strength test of the samples, it was found that the
alkaline activator ratio of 0.6 is almost the optimum ratio for all soils so, the durability of the geopolymer samples with different fly ash
ratios will be investigated at 0.6 alkaline activator. The durability of geopolymer was presented in terms of the strength after im­
mersion 28 days in H2SO4 (5%vl) and NaCl (5%wt).
In comparison with the non-exposed samples, it is noted that the stress-strain response is still brittle with stress reaching its peak
before abrupt failure, i.e., there is not a significant change in the behavior after the exposure to the chemical. However, lower un­
confined strength and higher axial strain are noted for such exposed samples. Fig. 17 shows the failure pattern in samples before and
after the immersion in the chemicals.
The mechanism which may occur during the soaking in the sulfuric acid includes an acid attack on the geopolymer matrix’s

11
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 14. Stress-strain relationship to the test specimens immersed in H2SO4 solution.

aluminosilicate bonding. This may cause degradation of the geopolymeric network, especially the aluminosilicate bonding, increasing
the production of Al–OH and Si-OH, and consequently, a loss of compressive strength. However, a slight loss of strength was observed
at high fly ash content. The percentage of reduction in UCS was (12%) at 20% fly ash, as shown in Fig. 14. The geopolymer treated
samples showed good durability for acidic resistance. This resistance can be attributed to the type of gel matrix formed and may be
closely related to the intact geopolymer network structure of Si–O and Al–O in the matrix. Similar findings were observed by [50], who
found that fly ash-based geopolymer cement is more acid-resistant than slag-based geopolymer. A large amount of N-A-S-H gel formed
less sensitive to sulfuric acid attack from the C-A-S-H gel. Fig. 15 presents the results of the compressive strengths of samples immersed
in chlorides. The change in the strength after immersion was not exceeded (10%) at 10% of fly ash, while the differences were slight of
higher ratios. In general, the geopolymer treated samples showed good durability for acidic and chlorides environments. Fig. 16 shows
the residual compressive strength ratio, which is defined as the ratio of compressive strength after immersion in solutions to
compressive strength before immersion. The maximum residual strength ratio was reported when the fly ash was 20%. This indicates
that as fly ash content increased, the pore volume decreased, and the matrix became dense, and consequently, the resistance of the
binder to an acid and salt attack was improved. It is essential to mention that samples in this study were immersed in chemical solutions
for the short term, only 28 days.

12
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 15. Stress-strain relationship to the test specimens immersed in NaCl solution.

4. Conclusions

The original motivation of this study comes from the following points:

1. With the prevalence of weak soils and demand to create structures due to civilization’s acceleration, the need for soil stabilization
has grown.
2. The need to search for environmentally friendly and sustainable alternatives to reduce the environmental impacts of traditional soil
stabilization materials (e.g., OPC and lime), exemplified by the high carbon footprint and other negative environmental effects
associated with sourcing and excessive exploration of non-renewable raw materials.
3. Extensive studies are needed on the effectiveness of geopolymer as a sustainable and environmentally friendly technology for
stabilizing soft soils and determining the optimum quantity of geopolymer ingredients that can improve the mechanical and
durability properties.
4. Solutions are needed to overcome some practical problems of geopolymers such as processing temperature and high content, which
limit their application in the field.

13
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

Fig. 16. The residual compressive strength ratio of geopolymer samples immersed in H2SO4 and NaCl solutions.

Fig. 17. Failure form for samples:a) Without immersion in solutions; b) immersion in H2SO4;c) immersion in NaCl.

To decrease the desired quantity of activator content (i.e., enhance cost-effectiveness) while retaining efficient ambient tem­
perature curing, this research has concentrated on increasing the reactivity of the geopolymer by employing fly ash with high
calcium concentration (high-calcium Class C fly ash). Clay soils with different contents of sand, clay (5% sand), soil-1 (10% sand),
soil-2 (20% sand), and soil-3 (30% sand) were stabilized using different ratios for fly ash (10,15% and 20%), considering different
alkaline ratios (0.4, 0.6, and 0.8). SEM analysis was used to examine the microstructural development of soil- geopolymer.
Moreover, the durability of geopolymer-treated clay against severe chemical attacks in the short term for 28 days (acid and salt
resistance) was evaluated. Some conclusions can be drawn as follows:
1. Using geopolymer based on high-calcium Class C fly ash promoted geopolymers’ prospective usage as an effective and envi­
ronmentally sustainable alternative binder to OPC for soil stabilization. The fly-ash FA content increased the unconfined
compression strength (UCS) for the geopolymer-treated soils significantly for all activator ratios. For fly-ash FA content of 20%
and activator ratio of 0.4, the strength improvement index reached 6.8, 7, 7.2, and 7.5 for clay soil, soil-1,soil-2, and soil-3,
respectively.
2. Increasing the sand (silica) in clay soil proportion decreased the optimum activator ratio by about 25%. The content of the
alkaline activator plays an essential role in the polymerization. Irrespective of the fly-ash FA content, the optimum activator
ratio was 0.8 for clay soil, and 0.6 for soil-1, soil-2, and soil-3.
3. The dominant stress-strain response for geopolymer-treated soils was found to be a brittle yield, with the stress peaking before a
sudden failure. When the geopolymer ratio increases, the response becomes stiffer (i.e., higher unconfined force and lower axial
strain).
4. The geopolymer treated samples (activator of 0.6) showed high durability against acidic and chlorides environments. For the
low content of fly-ash FA, the reduction in UCS was not exceeded 10% and 38% against chloride (NaCl) and acid (H2SO4),
respectively.
5. Results of SEM/EDS imply that the high performance of the stabilized soil owing to the reactive phase that was dissolved
resulted in two types of binder; Na-polysialate (Si-O-Al) and Na-polyferrosialate (Si-O-(Fe)Al).

This investigation demonstrated that soil geopolymer mixtures performed satisfactorily for acid and salt durability tests, in addition

14
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

to the development of strength and microstructure. This clearly promotes prospective geopolymers as an effective and environmentally
sustainable alternative binder to OPC for soil stabilization.

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.

References

[1] R.H. Karol, Chemical Grouting and Soil Stabilization, Revised and Expanded, vol. 12, Crc Press, 2003.
[2] G.P. Makusa, Soil Stabilization Methods and Materials in Engineering Practice, J0urnal, vol. 1, pp. 1–35, 2012.
[3] A.A. Amer, S. El-Hoseny, Properties and performance of metakaolin pozzolanic cement pastes, J. Therm. Anal. Calorim. 129 (1) (2017) 33–44.
[4] S. Rios, C. Ramos, A. Viana da Fonseca, N. Cruz, C. Rodrigues, Mechanical and durability properties of a soil stabilised with an alkali-activated cement, Eur. J.
Environ. Civ. Eng. 23 (2) (2017) 245–267, https://doi.org/10.1080/19648189.2016.1275987.
[5] L. Vickers, A. van Riessen, and W. Rickard, Fire-resistant Geopolymers: Role of Fibres and Fillers to Enhance Thermal Properties, SpringerBriefs in Materials. pp.
17–18, 2015, [Online]. Available: 〈http://link.springer.com/content/pdf/〉10.1007/978–981-287–311-8.pdf.
[6] A.A. Yusuf, F.L. Inambao, Characterization of Ugandan biomass wastes as the potential candidates towards bioenergy production, Renew. Sustain. Energy Rev.
117 (2020), 109477.
[7] K.A. Buyondo, P.W. Olupot, J.B. Kirabira, A.A. Yusuf, Optimization of production parameters for rice husk ash-based geopolymer cement using response surface
methodology, Case Stud. Constr. Mater. 13 (2020), e00461.
[8] A.A. Yusuf, F.L. Inambao, A.S. Hassan, S.S. Nura, V. Karthickeyan, Comparative study on pyrolysis and combustion behavior of untreated Matooke biomass
wastes in East Africa via TGA, SEM, and EDXS, Int. J. Energy Environ. Eng. 11 (2) (2020) 265–273.
[9] P. Sargent, P.N. Hughes, M. Rouainia, A new low carbon cementitious binder for stabilising weak ground conditions through deep soil mixing, Soils Found. 56
(6) (2016) 1021–1034, https://doi.org/10.1016/j.sandf.2016.11.007.
[10] M. Zhang, H. Guo, T. El-Korchi, G. Zhang, M. Tao, Experimental feasibility study of geopolymer as the next-generation soil stabilizer, Constr. Build. Mater. 47
(2013) 1468–1478.
[11] J. Davidovits, Geopolymers and geopolymeric materials, J. Therm. Anal. 35 (2) (1989) 429–441.
[12] W.K.W. Lee, J.S.J. Van Deventer, The effect of ionic contaminants on the early-age properties of alkali-activated fly ash-based cements, Cem. Concr. Res. 32 (4)
(2002) 577–584.
[13] M. Yaghoubi, A. Arulrajah, M. Miri Disfani, S. Horpibulsuk, M. Leong, Compressibility and strength development of geopolymer stabilized columns cured under
stress, Soils Found. 60 (5) (2020) 1241–1250, https://doi.org/10.1016/j.sandf.2020.07.005.
[14] A. Palomo, M.T. Blanco-Varela, M.L. Granizo, F. Puertas, T. Vazquez, M.W. Grutzeck, Chemical stability of cementitious materials based on metakaolin, Cem.
Concr. Res. 29 (7) (1999) 997–1004.
[15] T.-W. Cheng, J.P. Chiu, Fire-resistant geopolymer produced by granulated blast furnace slag, Miner. Eng. 16 (3) (2003) 205–210.
[16] M.T. Marvila, A.R.G. de Azevedo, P.R. de Matos, S.N. Monteiro, C.M.F. Vieira, Rheological and the fresh state properties of alkali-activated mortars by blast
furnace slag, Materials 14 (8) (2021) 2069.
[17] M.T. Marvila, A.R.G. de Azevedo, L.B. de Oliveira, G. de Castro Xavier, C.M.F. Vieira, Mechanical, physical and durability properties of activated alkali cement
based on blast furnace slag as a function of% Na2O, Case Stud. Constr. Mater. 15 (2021), e00723.
[18] A.R.G. de Azevedo, M.T. Marvila, H.A. Rocha, L.R. Cruz, C.M.F. Vieira, Use of glass polishing waste in the development of ecological ceramic roof tiles by the
geopolymerization process, Int. J. Appl. Ceram. Technol. 17 (6) (2020) 2649–2658.
[19] A.R.G. de Azevedo, et al., Circular economy and durability in geopolymers ceramics pieces obtained from glass polishing waste, Int. J. Appl. Ceram. Technol.
(2021).
[20] M.T. Marvila, A.R.G. Azevedo, G.C.G. Delaqua, B.C. Mendes, L.G. Pedroti, C.M.F. Vieira, Performance of geopolymer tiles in high temperature and saturation
conditions, Constr. Build. Mater. 286 (2021), 122994.
[21] M.T. Marvila, A.R.G. de Azevedo, C.M.F. Vieira, Mecanismos de reação de materiais álcali ativados, Rev. IBRACON Estrut. Mater. 14 (3) (2021).
[22] A.H.J. Al-Rkaby, Evaluating shear strength of sand-GGBFS based geopolymer composite material, Acta Polytech. 59 (4) (2019) 305–311, https://doi.org/
10.14311/AP.2019.59.0305.
[23] I. Phummiphan, S. Horpibulsuk, P. Sukmak, A. Chinkulkijniwat, A. Arulrajah, S.-L. Shen, Stabilisation of marginal lateritic soil using high calcium fly ash-based
geopolymer, Road. Mater. Pavement Des. 17 (4) (2016) 877–891.
[24] H.H. Abdullah, M.A. Shahin, M.L. Walske, Geo-mechanical behavior of clay soils stabilized at ambient temperature with fly-ash geopolymer-incorporated
granulated slag, Soils Found. 59 (6) (2019) 1906–1920, https://doi.org/10.1016/j.sandf.2019.08.005.
[25] S. Rios, N. Cristelo, A. Viana da Fonseca, C. Ferreira, Structural performance of alkali-activated soil ash versus soil cement, J. Mater. Civ. Eng. 28 (2) (2016)
4015125.
[26] C. Teerawattanasuk, P. Voottipruex, Comparison between cement and fly ash geopolymer for stabilized marginal lateritic soil as road material, Int. J. Pavement
Eng. 20 (11) (2019) 1264–1274, https://doi.org/10.1080/10298436.2017.1402593.
[27] A. Fernandez-Jimenez, I. Garcia-Lodeiro, A. Palomo, Development of new cementitious caterials by alkaline activating industrial by-products, IOP Conf. Ser.
Mater. Sci. Eng. 96 (1) (2015) 12005.
[28] I. Garcia-Lodeiro, A. Palomo, A. Fernández-Jiménez, An Overview of the Chemistry of Alkali-Activated Cement-based Binders, Woodhead Publishing Limited,
2015.
[29] B. Adhikari, M.J. Khattak, S. Adhikari, Mechanical and durability characteristics of flyash-based soil-geopolymer mixtures for pavement base and subbase
layers, Int. J. Pavement Eng. (2019) 1–20.
[30] Y. Zhu, R. Chen, and H. Lai, Stabilizing Soft Ground Using Geopolymer: An Experimental Study, in CICTP 2020, 2020, pp. 1144–1155.
[31] N. Cristelo, S. Glendinning, L. Fernandes, A.T. Pinto, Effects of alkaline-activated fly ash and Portland cement on soft soil stabilisation, Acta Geotech. 8 (4)
(2013) 395–405.
[32] S. Sreelakshmi, B.K. Huchegowda, S. Chithaloori, G.K. Kumar, Strength and durability behaviour of geopolymer-stabilized soil. Geotechnical Characterization
and Modelling, Springer, 2020, pp. 585–591.
[33] X. Kang, G.-C. Kang, K.-T. Chang, L. Ge, Chemically stabilized soft clays for road-base construction, J. Mater. Civ. Eng. 27 (7) (2015) 4014199.
[34] A. Mohammadinia, A. Arulrajah, J. Sanjayan, M.M. Disfani, M. Win Bo, S. Darmawan, Stabilization of demolition materials for pavement base/subbase
applications using fly ash and slag geopolymers, J. Mater. Civ. Eng. 28 (7) (2016) 4016033.
[35] N.M. Johnson, R.C. Reynolds, G.E. Likens, Atmospheric sulfur: its effect on the chemical weathering of New England, Science vol. 177 (4048) (1972) 514–516.
[36] ASTM International, ASTM D1633 - 00(2007) Standard Test Methods for Compressive Strength of Molded Soil-Cement Cylinders, ASTM Int. West
Conshohocken, PA, USA, pp. 1–15, 2007, [Online]. Available: https://www.astm.org/DATABASE.CART/HISTORICAL/D1633–00.htm.
[37] H.H. Abdullah, M.A. Shahin, and P. Sarker, Stabilisation of clay with fly-ash geopolymer incorporating GGBFS, World Congr. Civil, Struct. Environ. Eng., pp.
1–8, 2017, doi: 10.11159/icgre17.141.
[38] H.H. Abdullah. An Experimental Investigation on Stabilisation of Clay Soils with Fly-ash Based Geopolymer, Curtin University, 2020.

15
N.A. Odeh and A.H.J. Al-Rkaby Case Studies in Construction Materials 16 (2022) e00988

[39] A.M. Mustafa Al Bakri, M.T. Muhammad Faheem, A.V. Sandu, A. Alida, M.A.A. Salleh, C.M. Ruzaidi, Microstructure studies on different types of geopolymer
materials, Appl. Mech. Mater. 421 (February 2014) (2013) 384–389, https://doi.org/10.4028/www.scientific.net/AMM.421.384.
[40] S. Wang, Q. Xue, Y. Zhu, G. Li, Z. Wu, K. Zhao, Experimental study on material ratio and strength performance of geopolymer-improved soil, Constr. Build.
Mater. 267 (xxxx) (2020), 120469, https://doi.org/10.1016/j.conbuildmat.2020.120469.
[41] X.Y. Zhuang, et al., Fly ash-based geopolymer: clean production, properties and applications, J. Clean. Prod. 125 (2016) 253–267.
[42] J.S.J. Van Deventer, J.L. Provis, P. Duxson, G.C. Lukey, Reaction mechanisms in the geopolymeric conversion of inorganic waste to useful products, J. Hazard.
Mater. 139 (3) (2007) 506–513.
[43] A. Fernández-Jiménez, A. Palomo, I. Sobrados, J. Sanz, The role played by the reactive alumina content in the alkaline activation of fly ashes, Microporous
Mesoporous Mater. 91 (1–3) (2006) 111–119.
[44] X. Guo, H. Shi, W.A. Dick, Compressive strength and microstructural characteristics of class C fly ash geopolymer, Cem. Concr. Compos. 32 (2) (2010) 142–147,
https://doi.org/10.1016/j.cemconcomp.2009.11.003.
[45] F. Pacheco-Torgal, J. Labrincha, C. Leonelli, A. Palomo, P. Chindaprasit, Handbook of Alkali-Activated Cements, Mortars and Concretes, Elsevier, 2014.
[46] C.R. Kaze, et al., Characterization and performance evaluation of laterite based geopolymer binder cured at different temperatures, Constr. Build. Mater. 270
(2021), 121443.
[47] A.H.J. Al-Rkaby, Z.M.S. Alafandi, Size effect on the unconfined compressive strength and Modulus of elasticity of limestone rock, Electron. J. Geotech. Eng. 20
(12) (2015) 1393–1401.
[48] A.H.J. Al-Rkaby, A. Chegenizadeh, H.R. Nikraz, An experimental study on the cyclic settlement of sand and cemented sand under different inclinations of the
bedding angle and loading amplitudes, Eur. J. Environ. Civ. Eng. 23 (8) (2019) 971–986, https://doi.org/10.1080/19648189.2017.1327891.
[49] A.H.J. Al-Rkaby, Strength and deformation of sand-tire rubber mixtures (STRM): an experimental study, Stud. Geotech. Mech. 41 (2) (2019) 74–80, https://doi.
org/10.2478/sgem-2019-0007.
[50] N.K. Lee, H.K. Lee, Influence of the slag content on the chloride and sulfuric acid resistances of alkali-activated fly ash/slag paste, Cem. Concr. Compos. 72
(2016) 168–179, https://doi.org/10.1016/j.cemconcomp.2016.06.004.

16

You might also like