You are on page 1of 34

Journal Pre-proofs

Use of Secondary Additives in Fly ash based Soil Stabilization for Soft Sub-
grades

Hadi Karami, Jaspreet Pooni, Dilan Robert, Susanga Costa, Jie. Li, Sujeeva.
Setunge

PII: S2214-3912(21)00075-1
DOI: https://doi.org/10.1016/j.trgeo.2021.100585
Reference: TRGEO 100585

To appear in: Transportation Geotechnics

Received Date: 29 January 2021


Revised Date: 4 May 2021
Accepted Date: 16 May 2021

Please cite this article as: H. Karami, J. Pooni, D. Robert, S. Costa, Jie. Li, Sujeeva. Setunge, Use of Secondary
Additives in Fly ash based Soil Stabilization for Soft Subgrades, Transportation Geotechnics (2021), doi: https://
doi.org/10.1016/j.trgeo.2021.100585

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2021 Elsevier Ltd. All rights reserved.


Use of Secondary Additives in Fly ash based Soil Stabilization for Soft
Subgrades

Hadi Karami1, Jaspreet Pooni2, Dilan Robert3*, Susanga Costa4, Jie. Li5 and Sujeeva. Setunge6
1Ph.D.Scholar, Civil Engineering Department, School of Engineering, RMIT University, Melbourne,
VIC 3001, Australia. Email: s3642429@student.rmit.edu.au
2 Ph.D.
Scholar, Civil Engineering Department, School of Engineering, RMIT University, Melbourne,
VIC 3001, Australia ((corresponding author). ORCID: https://orcid.org/0000-0001-8444-6955 Email:
s3436956@student.rmit.edu.au
3 Senior
Lecturer, Civil Engineering Department, School of Engineering, RMIT University,
Melbourne, VIC 3001, Australia (corresponding author). ORCID: https://orcid.org/0000-0002-5686-
7055 Email: dilan.robert@rmit.edu.au
4 Senior Lecturer, School of Engineering, Faculty of Science, Engineering & Built Environment,
Deakin University, Waurn Ponds, VIC 3216, Australia (corresponding author). ORCID:
https://orcid.org/0000-0001-5194-7542 Email: susanga.costa@deakin.edu.au
5 Professor,Civil Engineering Department, School of Engineering, RMIT University, Melbourne, VIC
3001, Australia (corresponding author). ORCID: https://orcid.org/0000-0002-0344-2707 Email:
jie.li@rmit.edu.au
6Professor,Civil Engineering Department, School of Engineering, RMIT University, Melbourne, VIC
3001, Australia. ORCID: https://orcid.org/0000-0001-5958-2189 Email: sujeeva.setunge@rmit.edu.au

Abstract

Expansive soils are widespread in many parts of the world. Due to its low strength, high compressibility,

and massive volumetric changes, these soils are a potential origin of damage to roads, buildings,

foundations and other geo-infrastructure. Extensive research has been conducted on the utilisation of

fly ash to stabilize expansive soils. This paper describes how the efficiency of fly ash based soil

stabilization can be improved using secondary additives. Class F fly ash, an industrial by-product, was

used as the base additive. Lime, CSA cement, enzyme and polymers were utilized as secondary

additives. A series of mechanical and microscopic tests (CBR, compaction test, SEM, XRD, FTIR and

TGA) was carried out on different combinations of additives. The results indicate that secondary

additives can be effectively used to improve the efficiency of fly ash based soil stabilization. Soil-fly

ash-lime-enzyme was identified as an optimum combination to enhance bearing capacity while soil-fly

ash-lime and soil-fly ash-enzyme also showed substantial improvements in subgrade performance.

Findings from laboratory investigations were verified applying into 3-D numerical modelling to
1
evaluate the pavement performance which revealed substantial benefits of pavement thickness

reduction when treated using secondary additives.

Highlights:

 Demonstrated the ability of secondary additives to enhance fly ash based soil stabilization.

 Identified soil-fly ash-lime-enzyme as the best combination to improve bearing capacity.

 Demonstrated that pavement thickness can be reduced when secondary additives are used.

 Revealed that addition of polymer has a negative impact on bearing capacity.

Keywords: Fly ash, expansive soils, stabilization, microstructure, strength, secondary additives,

pavement

1 Introduction

Moisture ingress into road pavements can weaken the road system, impacting its service life and

sustainability. Settlement of the pavements due to seasonal moisture change can cause significant

longitudinal cracking for expansive or soft subgrades in roads and flexible pavements (Armstrong and

Zornberg, 2018, Zornberg et al., 2017). Presence of soft subgrade (i.e soils with a high amount of fine

material) could result excessive costs as required to replace them with better quality soils such as new

crushed rock for durable road performance (Trzebiatowski et al., 2005, Fauzi et al., 2013). Instead,

improving ground conditions by means of cost effective measures are widely recognised and practised

for to modify ground conditions (Correia et al., 2016). One of the common soil stabilization methods

widely used in the construction industry is jet grouting which can cope with various geotechnical

challenges such as fly ash soil stabilization (Shen et al., 2021). Another applicable and economical

method to dealing with soft subgrades is stabilizing the in-situ soil using waste material such as fly ash

with pozzolanic behaviour as a base additive mix with other stabilizers such as lime, Calcium

Sulfoaluminate (CSA) cement, enzyme and polymer. This research focuses on the use of fly ash and

secondary additives to stabilize soft subgrade soils for improving its behaviour and long term

performance.

2
Fly ash can be an ideal application to treat expansive soil, taking advantage of its pozzolanic properties,

as evidenced in the literature. Expansive soils are highly susceptible to moisture induced volume change

fluctuations, and in the presence of a new silt-sized and non-plastic additive such as fly ash, it assists in

reducing the swelling behaviour and increasing the bearing capacity in terms of ability to support

construction traffic on the soil to form a sturdy working platform for road construction. Cation exchange

increases the flocculation of clayey soils in the presence of fly ash, which is an excellent combination

of silicate, aluminium, and iron oxides (Cokca, 2001, Zhang et al., 2019). Fly ash reduces the surface

area of the soil and, consequently, the water dependency, as a result reducing the swell potential, swell

pressure, and plasticity of the soils (Anand J. Puppala, 2003). Moreover, the pozzolanic nature of fly

ash causes a cementation process to take place, and the new cementitious materials help to prevent

further deduction in swelling character of expansive soils (Thomas, 1999). The amount of calcium in

the fly ash is an essential ingredient to indicate the fly ash behaviour in strength improvement, however,

the remaining alkalis (Na2O and K2O), carbon (LOI), and sulphate (SO3)) only marginally can affect

the performance (Thomas, 1999). Phanikumar (2007) has shown that with the addition of fly ash to

expansive soils, the pore orientation of the soil can be modified, significantly improving its compaction

behaviour. Hence, increasing the fly ash content in clay soil modification shifts the maximum dry

density of the soil towards a lower water percentage, and the maximum dry density unit (MDD)

increases. This modified soil might be a suitable material for the road embankments, subgrades, and

backfill material (Phanikumar, 2007, Wang et al., 2019). Fly ash in soil treatment can act as a hydraulic

barrier due to the low coefficient of permeability in the order of 10-5 cm/s (Raymond, 1961, Dayal,

1988, Martin, 1990, Chen, 1992, Singh, 1994, Singh, 1996). Numerous literature has reported that

decreasing the strength is related to increasing fly ash content, however by incorporating an activator,

the pozzolanic strength of fly ash is significantly enhanced (Sivapullaiah, 1998, J.P. Prashantha, 2015).

Similarly, soil stabilized with alkali-activated fly ash exhibited a significant increase in stiffness, with

curing duration (Rios et al., 2017). Owing to the pozzolanic nature of fly ash, strength development

can vary due to its physical properties like grain size or the chemical contents such as reactive silica,

carbon, and iron (J.P. Prashantha, 2015).

3
Number of studies were reported in literature using fly ash as a pavement material stabilizing agent.

Application of fly ash class F as base layers in highways was investigated by Arora and Aydilek (2005),

who concluded that because of absence in self-cementing behaviour of the fly ash, it cannot be applied

to the soil without an activator such as Portland cement or lime. They further noted that the optimum

mixture needs to be obtained considering the adequate strength and durability, and should consider the

environmental concerns Kumar and Patil (2006) indicated the analytical ways for saving resources in

the embankments, granular sub-bases, water-bound macadams and pavement quality concrete layers of

fly ash based road formation. Hoy et al. (2016) investigated the strength development of recycled

asphalt pavement with fly ash geopolymer as a road construction material. (Eskioglou and Oikonomou,

2008) investigated the possibility of clay soil and sand stabilization with fly ash. They found that

stabilization with fly ash improves the natural and mechanical characteristics of soils. (Mishra and

Gupta, 2018) investigated the effect of engineering properties of soil by the inclusion of recycled

polyethylene terephthalate (PET) fibres in combination with the fly ash in the subgrade soil. They

showed an improvement in the shear strength, California Bearing Ratio (CBR) value, and a decrease in

the plasticity index for clayey soils. In order to enhance the efficiency of fly ash stabilized soils, several

research works have reported the use of secondary additives such as lime, fibre and cement in

conjunction with fly ash for effective soil improvement (Kumar et al.,2007; Mahvash et al., 2017;

Kaniraj and Havanagi, 1999; Janz and Johnsson, 2002. However, there is a need to investigate the use

of more sustainable materials to be incorporated as secondary additives in fly ash soil stabilization. The

current study aims to investigate the use of novel sustainable stabilizers in improving the fly ash based

soil stabilization with a validation analysis of verified 3-D numerical modelling to ascertain increased

efficiency of adopted method for sustainable soil stabilization.

This paper demonstrates the results of laboratory experiences on fly ash soil mixture for stabilization

incorporating traditional and non-traditional stabilizers on expansive clay subgrades soil. It investigates

the use of traditional (lime) and novel additives (i.e. CSA cement, polymer, and enzyme) considering

durability, cost, workability, and environment to propose a natural and eco-friendly approach for

stabilization of expansive soils using fly ash on unsealed pavements. The first phase involved a number
4
of laboratory experiments, varying from physical, mechanical and microscopic tests have been carried

out to analyse the effects of fly ash stabilization of expansive soils in conjunction with secondary

additives. CBR values indicate how to design the base and sub-base layer for the road and pavement

construction (Fauzi et al., 2013). The mechanical behaviour observed from the addition of secondary

additives in fly ash stabilized soils were explained using mineralogical and micrographic techniques.

Microstructural changes were assessed using the Scanning Electron Microscope (SEM),

Thermogravimetric Analysis (TGA), Fourier Transform Infrared (FTIR) and X-Ray Diffraction (XRD).

Three-dimensional numerical simulations were also performed to evaluate the efficiency of secondary

additives based fly ash soil stabilization. The results show that the incorporation of the selected

secondary additives has the potential to change the soil properties in the way of improving the

mechanical properties.

2 Research Significance

Fly ash has been proven to be a supplementary cementitious material which has been in use since last

century due to its pozzolanic behaviour and low cost. However, its use has not been optimised as

evidenced by millions of tons going to landfill every year instead of been recycled into roads and

pavements construction. This study investigates on how its use can be enhanced by improving its

efficacy for subgrades improvement, specifically soft subgrades, in combination with sustainable

additives. A number of studies reported in the literature for devising the combined benefits of fly ash

soil stabilization using common activators such as cement and lime, which at times resulted in

substantial cost and environmental concerns. The current study proposes the use of fly ash with cost-

effective and green secondary additives for introducing soft subgrade improvement methodologies for

sustainable pavements. The results from the study promote utilising coal combustion waste products to

stabilize expansive clay soils in soft subgrades in conjunction with secondary additives for sustainable

ground and pavement improvements.

3 Research Methodology

3.1 Materials

3.1.1 Soil

5
The soil for the experimental program was collected from a land excavation site in the Melbourne region

at ~ 0-2m depth. According to the Unified Soil Classification System (USCS) and the results from the

lab tests, the soil is categorised as a fine-grained soil (over 50% passing the sieve No.200) with a liquid

limit less than 50 and classified CL (ASTM-D2487, 2017). Based on the plasticity index, the soil is

classified with a high degree of expansivity (Holtz and Gibbs, 1956, Chen, 1988). Table 1 and Figure

1 show the physical properties of the soil and the particle size distribution curve, respectively.

Table 1: Physical Properties of the Soil

Figure 1: Particle size distributions of the fly ash (Leea et al., 2002, Dale P. Bentz, 2011) and soil

3.1.1 Fly ash

In this study, the fly ash class F is used as the primary and base additive. According USCS, fly ash is

classified as non-plastic fine silt. Fly ash used in the current study was sourced from local supplier. The

physical and mechanical properties for the applicable fly ash are shown in Table 2. Like other fly ashes,

this fly ash contains abundant silica and alumina. Table 3 shows the chemical composition of the fly

ash as a percentage of the weight. The particle size distribution for all sorts of fly ashes is similar to

having particles of silt size. Figure 1 gives the particle size distribution of representing fly ash used in

the current study (Leea et al., 2002, Dale P. Bentz, 2011).

Table 2: Physical Properties of Fly Ash (Kumar et al., 2007)

Table 3: Chemical composition of Fly Ash

3.1.2 Lime

In this study, hydrated lime, which is the most concentrated form of lime, was used as one of the

secondary additives. Hydrated lime was produced and supplied by an Australian material supplier.

Typical physical and engineering properties of hydrated lime can be found in Kumar et al. (2007). are

shown in Table 4.

Table 4: Physical Properties of Hydrated Lime (Kumar et al., 2007)

3.1.3 Enzyme

6
The enzyme-based stabilizer Eko Soil was also used as a secondary additive in this study. It is a non-

toxic material and a production of water and plant proteins. Eko Soil increases the soil compaction

capability and, consequently, the bearing capacity and strength of clayey soils. The optimised

application specifications of this enzyme are by dilution mass ratio (DMR) of 1:500 and application

mass ratio (AMR) of 1% (Pooni et al., 2019, Renjith et al., 2017). Moreover, Renjith et al. (2020)

conducted a detailed investigation on the stabilization effects of enzyme-based additive, and identified

that DMR 1:500 at 1% provides the optimum strength for the tested fine grained soil. CBR of soil

increased up to 500%, 425%, and 200% from the stabilization based on DMRs 1:100, 1:500, and 1:900,

respectively. The peak strength was observed at 1% AMR for DMR 1:500 (Renjith et al., 2020).

Typically, dilution mass ratio (DMR) of 1:500 and application mass ratio (AMR) of 1% is specified in

industry applications (EkoSoil, 2015). The Properties of Eko-Soil are shown in Table 5.

Table 5: Properties of Eko-Soil (EkoSoil, 2015)

3.1.4 CSA Cement

The third secondary additive used in this study is calcium sulfoaluminate (CSA) cement which is a

specific class of rapid-setting types of cement with a specific gravity of 3.1. Some characteristics such

as fast strength tracking, quick chemical reaction, and linking and less shrinkage, make the CSA cement

different from Portland cement (Dachtar, 2004, Bluey, 2019, Péra and Ambroise, 2004, Pooni et al.,

2020a). Chemical components and physical properties of CSA cement are shown in Table 6.

Table 6: Chemical composition and physical properties of CSA cement (Pooni et al., 2020b)

3.1.5 Polymer

The fourth secondary additive used in this study is a polymer which is based on water-soluble multi-

purpose and for dry mixes compounds. Mineral fillers are the main content of the anti-blocking unit in

this polymer, which has been produced without using organic solvents, plasticisers, and film-forming

agents. The polymer which has particle size above 400 microns has low emissions for formulating

components (WackerChemie, 2019).

7
3.2 Methods

Two variations of 7.5% and 15% of fly were selected for this research based on previous studies

(Phanikumar, 2007, Kumar et al., 2007) . The dry density, optimum water content, and bearing capacity

were determined by comparing stabilized soils and untreated samples. Based on literature, secondary

additives were used in following quantities; lime 3% (Kumar et al., 2007), enzyme DMR = (1:500) and

AMR = (1%) (Renjith et al., 2017, Pooni et al., 2019, Rintu Renjith, 2017), CSA cement 3% (Pooni et

al., 2020b), and polymer 3% (Kolay et al., 2016). The tests are described in detail in the following

section.

3.3 Experiments

California Bearing Ratio (CBR) and standard proctor compaction tests were performed to investigate

the densification and cementation effects of fly ash and other additives. The tests were performed as

per AS 1289.1.1:2001 for sampling and preparation of the soils (StandardsAustralia, 2001), AS

1289.5.1.1:2017 for standard Proctor compaction tests (StandardsAustralia, 2017) and AS 1289.6.1.1

for California Bearing Ratio test (StandardsAustralia, 2014). The experimental plan is summarised in

Table 7 and elaborated below.

Table 7: Summary of tests conducted

3.4 Sample preparation and testing program

For all test specimens of different mixtures, the soil and additives were mixed and moisturised to the

optimum water content using a standard proctor compaction test for each mixture. The soil was dried

for 24 hours in an oven at 110° C and sieved by 4.75mm sieve size. For each mixture of material, two

sets of CBR specimens were prepared using standard laboratory compaction followed by four days of

curing. Each specimen was compacted in a 2000 cm3 CBR mould in three layers applying 53 blows per

layer using a 2.5 kg rammer. For strength tests, the optimum quantities of the lime, enzyme, CSA

cement, and polymer were mixed with dry soil with various amounts of fly ash. Then according to the

standard compaction curve test results, the required amount of water was added to optimum moisture

content and samples were thoroughly mixed with a mechanical mixer. The samples were then
8
compacted according to Australian Standard 1289.6.1.1:2014, to obtain 117 mm in height, and 152 mm

in diameter. The California Bearing Ratio (CBR) tests were carried out after four days of soaked curing

in a water bath at a temperature ranging between 18 and 20о C. The tests were performed by static

tensile Shimadzu Autograph Testing Machine 50 kN, with 1mm per minute speed.

It is to be noted that the time dependent development is crucial in the understanding of strength

characteristics of soils treated with stabilizers as the structure of such materials evolves with time due

to continuing hydration/pozzolanic reactions (Salahudeen et al., 2014, Peethamparan and Olek, 2008).

The reported studies in literature have revealed that the stabilized soil mixtures using cementitious

admixtures have the potential for a time-dependent increase in strength and therefore with additional

curing time, further strength may develop (Amadi and Osu, 2018). Further studies are required to

understand the effect of curing on the new stabilized soil mix as proposed in the current study.

3.4.1 Soil–fly ash mixtures

To study the effects of other additives on fly ash stabilized soil, different mixtures were made. Type F

fly ash was added at 7.5% and 15% to the dry soil and then moisturised to their respective OMC as

obtained by standard compaction curve test. Before compacting the samples in the mould, they were

cured in sealed bags for 24 hours to a full equilibrium under room temperature and then well mixed by

mechanical mixer before compacting in the mould.

3.4.2 Soil-fly ash-lime mixtures

The soil-fly ash mixture was dry mixed with 3% of the dry soil weight of lime, and then water was

added to the optimum moisture content percentage obtained by the standard proctor compaction test

result for the control soil. The soil-fly ash-lime and water were thoroughly mixed by a mechanical

laboratory planetary mixer. Then the prepared soil mixture was cured in sealed and clear plastic bags

for 24 hours to permit moisture equilibration. After the curing duration, the CBR tests sample

preparation was carried out on the soil sample in order to obtain the bearing capacity of the soil.

3.4.3 Soil-fly ash-CSA cement mixtures

9
In the CSA cement mixing procedure, the fly ash was added to the dry soil and mixed with water to 3%

below the optimum moisture content obtained by the standard compaction curve and cured sealed for

24 hrs in clear plastic bags at room temperature. The CSA cement was added as a solution with the 3%

retained water to the cured soil-fly ash mix and mixed well with mechanical lab mixer. Since the CSA

cement was added to the soil, the reaction is expected to commence in a short time frame. Hence the

samples were compacted immediately after adding the CSA cement to the soil mix.

3.4.4 Soil-fly ash-lime-enzyme mixtures

For the soil-fly ash-lime-enzyme mixture, the lime, fly ash, and soil were mixed before adding water,

and water was added at 1% less than the optimum moisture content. Subsequently, samples were cured

for 24 hours in sealed clear plastic bags. After curing and equilibration of the mixture, the diluted

enzyme was added to the mixture and blended mechanically. Then the soil, fly ash, lime, and enzyme

mixture were cured for one hour in a sealed plastic bag (Pooni et al., 2019) and then compacted in the

CBR mould.

3.4.5 Soil-fly ash-CSA-enzyme mixtures

In this mixture, the fly ash and soil were mixed by hand in a tray with water at 3% lower than optimum

water percentage attained by standard compaction test (AustralianStandard, 2006) and cured for 24

hours at room temperature in sealed clear plastic bags. After equilibration, a 1% diluted enzyme was

added and mixed through with a mechanical mixer for 2-3 minutes and cured for another hour in sealed

plastic bags (Pooni et al., 2019). The CSA material added to the 2% remaining water, dissolved and

mixed well with the soil mixture for 3 to 5 minutes before compaction, by the mechanical mixer.

3.4.6 Soil-fly ash-polymer mixtures

In the soil-fly ash-polymer mixture, all the contents were added together and homogeneously mixed in

dry conditions. Subsequently, the optimum water content was mixed in and cured for 24 hours in sealed

clear plastic bags. The polymer, which added to the dry soil-fly ash mixture, is 3% of the dry soil weight.

3.4.7 Soil-fly ash-polymer-enzyme mixtures

10
Soil, fly ash, and the polymer was thoroughly mixed in a dry state; subsequently, water was added to

the mixture at 1% below the optimum moisture content obtained and cured in sealed and clear plastic

bags for 24 hours to allow equilibration. After 24 hours, 1% enzyme was added to the mixture and

blended with a lab mechanical mixer for 3 minutes. This mix cured for another hour before compacting

in a CBR mould before further testing.

3.4.8 Microscopic test preparation

Samples for microstructural testing were obtained from mechanically tested specimens and dried in the

110°C oven for 24 hours. For XRD and FTIR analyses, dried soil samples were grounded into fine

particles and sieved using no. 200 sieve to attain a fine powder for testing. Perkin Elmer Frontier FTIR

spectrometer was used to record the FTIR spectra. The fine powder (2 mg) was mixed with dry

Potassium bromide powder (KBr) (400 mg), forming a homogenous test sample that was placed into

the pellet and tested with 16 scans in the mid-infrared region of 4000 cm-1 to 400 cm-1 at a spectral

resolution of 4 cm-1. TGA curves were recorded using the TGA8000 apparatus by heating a 5 mg

samples at a rate of 10 °C/min to 850°C in a N2 atmosphere. Bruker D4 Endeavor diffractometer

apparatus equipped with a copper anode, 35 mA current, 40.0 kV input voltage and Cu-Kα radiation

were scanned using a 0.02° step size and a counting time of 21 min in the range of 6° to 90° determine

the XRD pattern. For SEM analysis, oven-dried soil fragments were filtered using the No.200 sieve and

subsequently placed on a 20 mm sample holder and carbon-coated. Samples were placed in the FEI

Quanta 200 and tested at a working distance of 10 mm, magnification of 7000 with 20 and 30 kV of

energy, and a spot size of 5.0.

4 Results

The results and analysis of the experimental tests are presented herein. Compaction behaviour of the

soil-fly ash mixtures with secondary additives is firstly reported, followed by the results from

mechanical tests. Subsequently, the results from the microstructural investigation are presented showing

the modification in the micro and mineralogical structure by secondary additives in soil-fly ash

mixtures.

11
4.1 Effect of fly ash based soil stabilization

4.1.1 Mechanical Behaviour

The compaction characteristics of expansive clayey soils are influenced by fly ash. It can be observed

from Figure 2 that with the addition of fly ash, there is a decrease in optimum moisture content (OMC)

and an increase in the maximum dry density (MDD) of soil-fly ash mixes compared to the untreated

soil. For example, at 7.5% fly ash content OMC and MDD are 19% and 1.76 g/cm3, respectively. The

decrease in the optimum moisture content of the soil-fly ash mix can be attributed to the impact of fly

ash in adhering soil particles together like cement clinkers. As fly ash content increases from 7.5% to

15%, this behaviour becomes more prominent. The addition of fly ash to the raw soil decreases the CBR

result (Figure 3) of the untreated soil by 6.6% and 10.7% for 7.5% and 15% fly ash contents,

respectively. The reduction in the mechanical characteristics probably is due to replacing the plastic

fines particles in the expansive soil by non-plastic fines (fly ash), which provide no friction or cohesive

strength (Kate, 2005, Bell, 1996). Previous research work has also shown that the suction properties

would be reduced in soil-fly ash, and the mechanical properties and swelling behaviour could be

reduced (Phanikumar, 2007).

Figure 2: Compaction characteristics of soil-fly ash mix against fly ash content, OMC (left) and MDD

(right)

Figure 3: CBR of soil-fly ash mix against fly ash content

4.1.2 Microscopic Behaviour

SEM images of untreated soil and soil-fly ash mixed are presented in Figure 4 (a-b). Untreated soil

consists of the following clay minerals: illite, quartz, montmorillonite, and kaolinite, as determined from

XRD and FTIR analysis. SEM imaging reveals that the untreated soil sample has a dispersed soil

structure consisting of a number of flake-like particles with small aggregates. The presence of a large

number of connected pores is likely to result in low strength. In contrast, fly ash treated soils shows

slight flocculation accompanied by the integration of smooth particles within the voids of the soils

12
(Figure 4 a and b). Figure 5 shows the TGA results for control and fly ash treated soil. The drop in

weight loss occurs from the addition of fly ash which decomposes.

Figure 4: Microanalysis report of (a) untreated soil (b) soil and fly ash (c) soil, fly ash and enzyme

mixture, (d) soil, fly ash, lime mixtures, (e) soil, fly ash, CSA cement mixture

Figure 5: TGA results for control and stabilized soils.

4.2 Effect of enzyme on fly ash stabilized soils

4.2.1 Mechanical Behaviour

The enzyme additive in the soil-fly ash mixture marginally impacts the compaction characteristics.

From the addition of 1% enzyme additive to the soil-fly ash mixture, there is an average 2.1% change

in OMC and an average decrease of 0.9% in MDD as can be seen in figure 6. This change in OMC is

consistent with the clay’s reduced affinity for water (Pooni et al., 2019).

The addition of the enzyme additive to expansive clays leads to an increase in mechanical strength,

consequently, an improved CBR providing sufficient strength for a reduced pavement layer (Pooni et

al., 2019). The CBR results of enzyme treated mixtures are higher than those of soil-fly ash samples.

For soil-fly ash mixture treated with enzyme additive, CBR results increase by 80% and 37% in the

presence of 7.5% and 15% fly ash, respectively (Figure 7). The improvement in bearing capacity can

be associated with the reduced affinity for water driven by the addition of the enzyme additive in

neutralising the negatively charged clay molecules.

Figure 6: Compaction characteristics of soil-fly ash-enzyme mix against fly ash content, OMC (left)

and MDD (right)

Figure 7: CBR of soil-fly ash-enzyme mix against fly ash content

4.2.2 Microscopic Behaviour

The influence of the enzyme on XRD analysis of fly ash treated soils are depicted in Figure 8. There

are no noticeable increases in the intensity of the peaks of the expansive clay minerals with the addition

of the enzyme stabilizer to fly ash treated soil mixtures. Moreover, the XRD spectra show there is an
13
absence of new peaks, highlighting the lack of pozzolanic products from the cementitious effect from

fly ash. This behaviour indicates that the enzyme soil stabilizer does not result in the formation of

cementitious products similar to that of other non-traditional stabilizers (Latifi et al., 2018). Moreover,

the enzyme stabilizer is not an activator for fly ash in forming pozzolanic bonds between soil particles

to increase soil strength but rather serves to facilitate compaction in improving density, as seen in the

competition plot.

Figure 9 compares the FTIR spectra of soil-fly ash mix with and without enzyme. Treatment of enzyme

showed a marginal reduction in the intensity of the interlamellar water region (3200 cm-1 - 400 cm-1 and

1640 cm-1) due to the reduced affinity by the enzyme soil stabilizer. The FTIR spectra do not show

changes that indicate the formation of pozzolanic products. Similarly, this evident in TGA results

(Figure 5) which reveal insignificant changes at 115-150°C compared to soil-fly ash samples. Hence

this further confirms that strength enhancement from enzyme treatment is not due to any pozzolanic

reactions between soil and fly ash particles, instead attributed to the enzyme stabilization mechanism.

Figure 4c shows the SEM image of soil-fly ash mix with the addition of the enzyme. The enzyme

modifies the soil texture, forming aggregations with sharp edges of soil and fly ash particles results in

soils that are less ‘flaky’ in nature. Soil-fly ash with enzyme mix gives more strength compared to soil-

fly ash mixtures (Figure 4b).

Figure 8: XRD analyses results for soil-fly ash-enzyme mix.

Figure 9: FTIR results for soil-fly ash-enzyme mix

4.3 Effect of lime on fly ash stabilized soils

4.3.1 Mechanical Behaviour

A considerable change in the compaction characteristics was observed for soil-fly ash mixtures

combined with lime. An increase in MDD is accompanied by a decrease in OMC is compared to soil-

fly ash mixtures. From the addition of 3% lime to the soil-fly ash mixture, there is a decrease in MDD

by 7.7% and 9.2% and an increase in OMC by 21.1% and 19.7% for 7.5% and 15% fly ash mixtures,

14
respectively. The significant reduction in density occurs instantaneously from flocculation of the soil

particles, leading to an increase in the void ratio of the mix and subsequently causing a reduced

maximum dry density of the mixture (Arvind Kumar, 2007). The increase in the optimum moisture

content of the soil mixture is a result of the increased capacity to hold water within the flocs that arise

from the flocculation of soil due to the addition of lime within the mix. As shown in Figure 10, the

addition of 1% enzyme to the soil-fly ash lime mixture facilitates the compaction process through the

clay’s reduced affinity for water; as a result, there is an average 2.9% decrease in OMC and an average

4.2% increase in MDD.

The addition of lime increased the CBR value of the soil-fly ash mix by 789% and 799% for 7.5% and

15% fly ash, respectively (Figure 11). From compaction characteristics, it can be observed that there is

a reduction in MDD, despite this CBR results show a significant improvement indicating lime performs

as an activator for fly ash resulting in pozzolanic reactions that contribute to strength enhancements.

Although the optimum moisture content is very high, the sample appears dry and friable after four days

soaked curing. This reflects more water is needed for saturation due to the low saturation coefficient,

which indicates a large number of unfilled voids in the lime mix. Moreover, the addition of 1% enzyme

solution to the soil-fly ash-lime mixture, decreases the OMC about 12% and increase the MDD by 3.2%

and 5.2% for 7.5% and 15% of fly ash contents, respectively (Figure 10) and increases the CBR result

(Figure 11) by 12% and 9.2%. Figure 10 and 11 show that by the addition of 1% enzyme, there is a

noticeable effect on the moisture content of the treated soil. The moisture content affects the mechanical

and engineering properties of soils by which the increment in water content reduces the mechanical

properties of the soil structures.

Figure 10: Compaction characteristics of soil-fly ash-lime mix against fly ash content, OMC (left) and

MDD (right)

Figure 11: CBR of soil-fly ash-lime mix against fly ash content

4.3.2 Microscopic Behaviour

15
Figure 12 presents the XRD spectra of soil-fly ash-lime mixtures. The addition of lime modified the

soil-fly ash with the formation of pozzolanic products as indicated by the appearance and modification

of peaks. Calcium from lime and fly ash react with alumina and silica from clay in the presence of water

to produce calcium silicate hydrate (CSH), calcium aluminate hydrate (CAH), and calcium aluminium

silicate hydrate (CASH). Peaks containing cementitious compounds such as CSH (at about 21.3°, 24.8°,

27.3° 27.9°, 29.4° 39.4°, 40.2°), CAH (at about 35.9°) and CASH (at about 29.3°) develop with the

addition of lime. Similar observations have been made by previous researchers as well (Sharma et al.,

2012, Jiang et al., 2015, Jha and Sivapullaiah, 2015). Moreover, the broad peak at about 8.96° becomes

sharper with the addition of fly ash and lime, indicating modification of the montmorillonite structure

(Sharma et al., 2012), amending its expansive nature. The presence of lime enables fly ash in the soil to

react, forming cementitious compounds.

According to FTIR results shown in Figure 13, the addition of lime in the soil-fly ash mixture indicates

the presence of the pozzolanic products. The strong shoulder peak at ~ 970 cm-1 is characteristic of CSH

formations in the presence of lime. Moreover, there is a change in intensity of the peak at ~ 900 cm-1

associated with silica as the intensity reduces. This is accompanied by an increase in peak intensities

towards ~ 1000 – 1100 cm-1 (Ylmén et al., 2009). The peak at ~ 875 cm-1 and the broadening peaks in

the range of ~ 1400 – 1500 cm-1 represent the bending of CO3 stretching of CO3. Moreover, TGA results

(Figure 5) show a sharp peak drop indicating an increased presence and decomposition of CSH and or

CAH. These results show the influence of lime in the formation of pozzolanic products.

The texture of the fly ash soil mixture with the addition of lime captured by SEM imaging is shown in

Figure 4d. The sharp edges observed on soil particles are characteristic of lime. In the presence of lime,

fly ash soil mixtures show an aggregation of soil particles with a ‘rough’ surface. The combination of

lime and fly ash within the soil mix consists of varying particle sizes, which poorly fills the voids within

the soil-forming cementitious formations binding clay particles. This explains the improvement in CBR

and a decrease in MDD.

16
Moreover, from the addition of enzyme additive to the soil-fly ash-lime mixture, there is a small shift

in peak value of montmorillonite (at about 8.96°), with a slight initial increase in d-spacing is followed

by a decrease in d-spacing (to the original, untreated d-spacing) due to lattice tightening caused by the

enzyme stabilizer (Rauch et al., 2003, Scholen, 1992). The FTIR spectrum of combined soil stabilizers

on fly ash treated soils is showed in Figure 13. The combination of lime and enzyme treatment on fly

ash soil mixtures showed FTIR, and TGA (Figure 5) changes similar to that of fly ash soil mixtures.

The enzyme stabilizer reduces the affinity for water, as shown form the reduced intensity and

broadening of the interlamellar water region (3200 cm-1 - 400 cm-1 and 1640 cm-1). The peak

characteristics observe are consistent with CSH formations (at ~ 970 cm-1, ~ 1000 – 1100 cm-1) from

lime.

Figure 12: XRD analyses results of soil, fly ash, lime and enzyme mixtures. a) Whole XRD, b) at 7-10

degrees, c) at 20-23 degrees, d) at 24-27 degrees e) at 27-30 degrees, f) at 30-33 degrees g) at 33-36

degrees h) at 38.5-41.5 degrees

Figure 13: FTIR results for of soil, fly ash, lime and enzyme mixtures

4.4 Effect of CSA cement on fly ash stabilized soils

4.4.1 Mechanical Behaviour

Marginal change in the compaction curve is observed with the addition of CSA cement to the soil-fly

ash mix (Figure 14). CSA cement additive to the soil-fly ash mixture results in a 2% and 1.5% reduction

in MDD and 1.1% increase in the OMC for soil-fly ash mixtures with 7.5% and 15% fly ash,

respectively. The relatively high bulk density of 1000-1300 kg/m3 and a high specific gravity of 3.1 of

CSA cement compared soil and fly ash has led to a significant increase in MDD. However, when

enzyme additive is incorporated into the soil-fly ash-CSA cement mixture, there is an average 4.2%

decrease in OMC and an average 3.8% increase in MDD from the reduced affinity for water and

improved workability of clay soil.

Figure 15 shows that the addition of 3% CSA cement considerably increased CBR of the soil-fly ash

mix by about 621% and 384% for 7.5% and 15% fly ash contents, respectively. On the other hand, in
17
the presence of the enzyme CBR results of soil-fly ash CSA cement mixtures decrease from 21.5% and

13.22% to 19% and 11% to for 7.5% and 15% fly ash contents, respectively. The CBR reduction from

enzyme additive in the soil-fly ash-CSA cement samples is likely from chemical reactions between

CSA and enzyme and changing the pH due to the lack of free lime in CSA cement (Antoun, 2019).

Figure 14: Compaction characteristics of soil, fly ash, CSA cement and enzyme mix against fly ash

content, OMC (left) and MDD (right)

Figure 15: CBR of soil, fly ash, CSA cement and enzyme mix against fly ash content

4.4.2 Microscopic Behaviour

XRD results for CSA added to soil - fly ash mixtures are presented in Figure 16. The addition of CSA

cement modified the fly ash treated soil with the formation of cementitious products as indicated by the

appearance and modification of peaks. The CSA cement and fly ash react with alumina and silica from

clay and fly ash in the presence of water to produce calcium silicate hydrate (CSH), calcium aluminate

hydrate (CAH), ettringite (AFt), monosulfate (AFm) and strätlingite. Peaks containing CSH develop at

about 21.3°, 24.8°, 27.3°, 29.4°, and 40.2°) and CAH peaks form at about 35.9°. Moreover, strätlingite

and AFm phases are present at two-theta of about 31.2° and 32.1°, respectively. AFt phases are

extensively present (22.1°, 37.4°, 42.5°, 46.0°, 46.5°, 51.1° 53.8° 67.4° 73.3°) in the CSA treated fly

ash soil mixtures indicating its presence as a significant hydration product. Furthermore, the increased

decomposition as evidence in TGA results from the addition of CSA cement to fly ash soil mixtures

confirming the formation of cementitious and pozzolanic products.

Figure 17 shows a comparison of the FTIR spectrum of CSA treated soil-fly ash mixtures. The addition

of CSA cement produced a number of changes in the FTIR spectrum, illustrating the presence of

cementitious formations. The doublet at the ~ 3696 cm-1 and ~ 3620 cm-1 experience a subtle reduction

in intensity due to strengthening of the peak at 3650 cm-1, indicating the stretching vibrations of O-H

free present in AFt formations. The peak at ~ 970 cm-1, which is characteristic of CSH formations, does

not strengthen with CSA cement addition. The vibration bands between ~ 440 – 690 cm-1 (which is

attributed to Si–O–Si or Si–O–Al bonds) (Murmu et al., 2019) increases in intensity with CSA cement,
18
hence, highlighting the increased silica content and cementitious and pozzolanic formations as depicted

in XRD studies.

The texture of soil when the CSA cement is mixed with soil-fly ash is shown in Figure 4e. The surfaces

consist of sharp particles integrated with needle-shaped structures. The reaction of CSA cement in soil

demonstrates flocculation and agglomeration, followed by its hydration producing a number of sharp

space-filling ettringite needles between the soil particles, forming large interconnected soil particles.

Compared to fly ash, lime, and enzyme, the CSA cement treatment has more compacted particles; hence

soil-fly ash combined with CSA cement provides additional strength and density compared to soil-fly

ash mixtures with fewer voids in the mixture. As a result, this improves the bearing capacity and

strength.

The addition of enzyme additive to the soil-fly ash-CSA cement mixture modified the XRD pattern.

From the addition of the enzyme soil stabilizer, there is a slight initial increase in d-spacing, followed

by a decrease in d-spacing due lattice tightening (Rauch et al., 2003, Scholen, 1992). As a result, this

causes a slight shift in the peak value of montmorillonite (at about 8.96°) highlights a change in the

expansive nature. Moreover, the TGA results and FTIR spectrum shows similar trends to that of soil-

fly ash soil mixtures. The enzyme stabilizer reduces the affinity for water, as indicated by the reduced

intensity and broadening of the interlamellar water region (3200 cm-1 - 400 cm-1 and 1640 cm-1). The

vibrational bands observed are consistent with AFt formations (3650 cm-1) from CSA cement in the soil

mixture.

Figure 16: XRD analyses results for soil, soil, fly ash, CSA cement and enzyme mixture. a) whole XRD,

b) at 7-10 degrees, c) at 20-23 degrees, d) at 24-27 degrees e) at 27-30 degrees, f) at 30-33 degrees g)

at 33-36 degrees, h) at 38.5-41.5 degrees, i) at 42-45 degrees, j) at 45-48 degrees, k) at 51-54 degrees,

l) at 66-69 degrees, m) at 71-74 degrees

Figure 17: FTIR results for soil, fly ash, CSA cement and enzyme combinations

4.5 Effect of polymer on fly ash stabilized soils

19
4.5.1 Mechanical Behaviour

Compaction characteristics experience a change in MDD and OMC from the addition of the polymer

to soil-fly ash mixtures (Figure 18). Polymer additive to the soil-fly ash mixture results in an average

4.8% reduction in MDD and an average 2.8% increase in the OMC for soil-fly ash mixtures. The low

bulk density of 490-590 kg/m3 of polymer compared to soil and fly ash has led to a decrease in MDD.

With the addition of the enzyme additive to soil-fly ash polymer mixture, there is an average 4.7%

decrease in OMC and an average 3.4% increase in MDD from the reduced affinity for water and

improved workability of clay soil.

Polymer additive in the soil-fly ash mixture results in a reduction CBR. Figure 19 shows that CBR

decreased by 23.15% and 27.37% for 7.5% and 15% fly ash, respectively. The polymer-fly ash mix,

however, performed better with the addition of enzyme, resulting in a significant increase in CBR by

202% and 153% for 7.5% and 15% fly ash contents, respectively. Due to the insignificant mechanical

strength change observations due to addition of the polymer to the soil-fly ash mixture, a detailed

microscopic analysis was not performed.

Figure 18: Compaction characteristics of soil, fly ash, polymer and enzyme mix against fly ash content,

OMC (left) and MDD (right)

Figure 19: CBR of soil, fly ash, polymer and enzyme mix against fly ash content

4.6 Numerical modelling description

Numerical simulations were performed to evaluate the performance of unsealed roads stabilized with

fly ash mixes enhanced with novel additives. The numerical model evaluates the response and

deformation of stabilized unsealed roads subjected to static traffic loading. Such an approach compares

the novel additives in fly ash mixtures, highlighting the impact on the application to soil stabilization

for unsealed roads.

4.6.1 Pavement properties

20
A minor rural road, consisting of a formed pavement with a CBR of 3% and a typical daily traffic of

20-100 vehicle (1.3 × 104 Equivalent Standard Axles, ESA) was simulated following Austroads

empirical design guidelines (Austroads, 2009). The assumed roadway consisted of two lanes with a

total width of 6.2 m, and subgrade depth of 5 m, however only half the road with was simulated due to

the symmetrical nature. Typical stabilization depths of 150 mm were implemented as per design

manuals and guidelines (Austroads, 2009). The soil elastic properties were obtained from the extensive

laboratory testing conducted in this study. A typical base material detailed in Austroads for unsealed

roads was selected with a CBR of 20%, and density of 2.24g/cm3. The pavement layer thicknesses for

the base material were determined from empirical design charts (Austroads, 2009). Table 8 shows the

pavement design thickness of the base layer for each stabilization approach. The Poisson’s ratio of the

pavement materials was taken as 0.35.

Table 8: Pavement Design Thickness

4.6.2 Numerical simulation

Numerical modelling of the road pavements was carried out using ABAQUS. The pavement was

represented by a 3-D model with elastic material properties using 8-noded brick, with reduced

integration elements (C3D8R). Soil boundaries were assumed to be smooth and located far from

traffic loads to eliminate boundary effects. Mesh distribution towards the static 80 kN single-axle-

dual-tyre (SADT) loading was fine, whereas mesh towards the model boundaries was coarse. Static

loading was distributed over an equivalent circular area. The adopted geometry and mesh

discretization are showed in Figure 20.

Figure 20. Geometry and mesh discretization (a) 3D geometry, (b) Cross-sectional view

4.6.3 Analysis Plan

The analysis plan in this study was performed using a 3-D model to evaluate the response of the

stabilized unsealed pavements designed. Firstly, based on the laboratory results from this study, the

stabilized pavements were designed according to the empirical design charts for the unsealed road to

evaluate stresses, strains and deformations along the surface of each layer. Secondly, a parametric study
21
was performed to determine the minimum base thickness required over the unsealed pavements

stabilized with fly ash mixes enhanced with novel additives to satisfy the subgrade strain criteria.

4.7 Effect of Stabilizers on Unsealed Road Pavement Response

Results from the analysis are summarized in Table 9 for the elastic stresses, strains and deformations

along with surface pavement layers. Table 9 shows that subgrades of unsealed pavements stabilized

with combined fly ash and secondary additives require a smaller reservation of cover over the in-situ

subgrade compared to untreated and soil-fly ash stabilized pavements. This trend arises from the

improved bearing capacity of the stabilized subgrade soil. As a result, smaller stresses, strains and

deformation are induced in the pavement subgrade compared to the untreated pavements. For example,

Figure 21 shows the maximum strain distribution of 1640 µ and 1475 µ in the untreated and fly ash-

lime-enzyme stabilized pavement, respectively. Even though strain difference is mild (10%), the

implication on the base layer is significant, showing a 50% reduction in base layer thickness and hence

considerable savings in material. This is due to the stabilized subgrade having a subsequent CBR greater

12% for the given traffic and in-situ subgrade strength. Between the pavements with the same 150 mm

base material, stabilized subgrades with higher stiffness has lower stresses, strains and deformation on

the pavement. For example, soil-fly ash-lime stabilized subgrades have a 17% lower strain and 9%

lower deformation at the subgrade level compare to soil-fly ash-CSA cement. Moreover, the addition

of enzyme additive to soil-fly ash-lime stabilized subgrades further reduce the stains by 11.9%. Soil-fly

ash mixes are enhanced with novel additives, there improved bearing capacity serve to reduce the

required cover over the in-situ subgrade when using the empirical design charts. All recorded subgrade

strains, and deformations of stabilized pavements are comfortably within the allowable limit (2653.3

µε and 20 mm) for flexible pavements suggesting base layer thickness reduction should be re-evaluated

to yield an economical design.

Table 9: Critical elastic stress response due to traffic loading

Figure 21. Strain distribution (a), untreated pavement, (b) fly ash-lime-enzyme stabilized pavement

22
The results of the parametric study are summarised in Table 10. The study revealed that base thickness

determined from the empirical pavement design can be substantially reduced whilst fulfilling the

subgrade strain criteria. It can be seen that by stabilizing the subgrade with fly ash, the base material

can be reduced by about 57%. When fly ash stabilized soils are combined with secondary additives,

further reduction in base thickness up to 67% occurs owing to the improvement in bearing capacity of

the material.

Table 10. Minimum base thickness

5 Discussion

The performance of fly ash treated soils are controlled by both densification and chemical reactions.

Fly ash alters the soils compaction characteristic by filling the void space, while calcium hydroxide

reacts with silicates and aluminates in clay resulting in pozzolanic reactions. As a result, the strength

behaviour of fly ash stabilized subgrade soils has a direct implication on pavement performance.

5.1 Effect on bearing capacity

There is a negligible change in bearing capacity with the addition of only fly ash, suggesting that there

are no pozzolanic reaction contributing to strength from Class F fly ash. However, when secondary

additives (enzyme, lime and CSA cement) are combined with soil-fly ash mixes, there is considerable

improvements in bearing capacity. Addition of 3% CSA cement in the soil-fly ash mix increases CBR

by 20% in fly ash mixes. Similarly, the addition of 3% lime to the soil-fly ash mix yields a CBR of

25%. Hence, the addition of lime produces the most significant improvement in CBR, likely due to the

activation of Class F fly ash in the soil, which promotes pozzolanic reactions and cementitious products.

Contrary to this, polymers had an adverse effect on the bearing capacity.

As evidenced, by incorporating enzyme stabilizer to soil-fly ash mixes CBR increases by about 5%.

Moreover, where enzymes are used to complement secondary binders in soil-fly ash mixes there are

noticeable changes. Complementing the secondary binders in soil-fly ash mixes with enzyme produced

23
differing results. Additional 2 – 5% improvement in CBR is seen in lime and polymer treated soils with

enzymes, whereas a 2% decrease in CBR is observed with CSA cement with enzymes. Furthermore,

the results show that excessive amounts of fly ash (>7.5%) hinders the interaction between the soil

grains and secondary binders and hence does not yield any mechanical benefits. However, an excess

amount of fly ash does not impact adversely in lime-fly ash mixes due to the presence of sufficient

calcium hydroxide to activate the excess pozzolan. Overall, soil-fly ash-lime-enzyme produces the

highest improvement in terms of bearing capacity.

5.2 Effect on microstructural behaviour

The addition of fly ash to expansive subgrade soil serves to fill the void spaces in the mix. This change

in soil texture by flocculation, agglomeration, and the pozzolanic reaction from fly ash is observed in

the microstructural and imaging tests. Figure 4b (fly ash SEM) shows slight flocculation accompanied

by the integration of smooth particles within the voids of the soils caused by the pozzolan bonding to

clay. Microstructural tests also reveal the modification and breakage of expansive clay mineral

montmorillonite from fly ash. Despite these changes to soil texture, improvement in bearing capacity is

negligible as mentioned above.

The pozzolanic nature of Class F fly ash requires an activator to react and produce a matrix of

cementitious materials promoting further strength development and effectiveness of fly ash treatment.

Mixing lime to soil-fly ash mixes causes cationic exchange between the clay particles and the Ca+ ions

resulting in the slow development of pozzolanic and cementing products, which is responsible for

mechanical strength improvement. Similarly, the addition of CSA cement to soil-fly ash mixes induces

cementitious reactions which demonstrate an increase in bearing capacity. The microstructural tests

reveal the appearance of new peaks and bands corresponding to cementitious and pozzolanic

formations. In the case of soil-fly ash-lime mixes, the appearance of peaks and bands corresponding to

CSH and CAH formations develop. Soil-fly ash-CSA cement mixes include the presence of ettringite

formations. The SEM imaging shows a strong interconnected framework among clay particles

indicating the presence of cement products with ettringite needles, filling the pore spaces and binding

clay particles together in soil-fly ash-CSA cement samples.


24
On the other hand, microstructural tests on soil-fly ash-enzyme samples show that strength

improvement does not arise from pozzolanic or cementitious reactions, indicating enzyme soil

stabilizers are not an activator for fly ashes. The SEM imaging shows (Figure 4c) close packing of clay

particles attributing to densification from the addition of the enzyme stabilizer. Enzymes as a

complementary additive to secondary binders (lime and cement) in soil-fly ash do not change the

microstructural behaviour apart from lattice tightening (Rauch et al., 2003, Scholen, 1992). The

appearance of CSH, CAH and cementitious formations illustrate the benefit of supplementary additives

in enhancing fly ash stabilization due to pozzolanic and cementitious reactions.

5.3 Effect on compaction characteristics

The addition of fly ash marginally increases the MDD of subgrade soil by filling the void spaces

followed by the subsequent flocculation and agglomeration during mixing, which decreases the OMC.

The addition of secondary binders further alters the compaction characteristics. The incorporation of

lime to soil-fly ash samples decreases MDD and increases OMC, resulting in a significant enhancement

in bearing capacity. This behaviour is due to the cationic exchange between Ca2+ and clay soils.

There is a negligible change in the compaction characteristics of soil-fly ash-CSA cement, despite

significant improvements in bearing capacity indicating cementitious formations as the stabilization

mechanism. When polymers are added to soil-fly ash, there is an increase in OMC and decrease in

MDD without cementitious formations. Moreover, as reported in the literature, enzymes marginally

increase MDD and decrease OMC of expansive soil. When the enzyme is combined with soil-fly ash

mixes, there is a noticeable decrease in OMC and increase in MDD. Due to the improved density, there

is a further increase in bearing capacity when the enzyme is used as a supplementary additive in soil-

fly ash-lime and soil-fly ash-polymer mixes. The change in compaction characteristics is an important

consideration in the strength of soil-fly ash mixes where strength development is slow.

5.4 Implication in unsealed pavements

Performance modelling of pavements incorporating stabilized subgrade with fly ash directly reflects the

influence of bearing capacity. Fly ash stabilized subgrades reveals no change in stresses and, strains

25
below the base layer of the pavement due to the negligible change in CBR. However, with the addition

of secondary additives, unsealed pavement performance is largely enhanced. The improved bearing

capacity from fly ash-lime stabilization permits a 50% reduction in base thickness in pavement design

according to the empirical procedure and results in lower stresses, strains and deformations at the

subgrade level compared to the untreated pavement. The degradation is slightly larger (by about 13%)

for fly ash-CSA cement stabilized pavements compared to fly ash-lime stabilization. Moreover, the

improved density and bearing capacity from the addition of enzyme stabilization in fly ash stabilized

subgrades reduces by 100 mm compared to the 300 mm cover necessary for untreated pavements

without increasing performance indicators. Therefore, enzyme as a supplementary additive to soil-fly

ash-lime, marginally reduces subgrade stains and surface deformations by 3% and 2%, respectively

when compared to lime-fly ash soils. The addition of secondary binders to fly ash stabilized subgrades

reduces required base thickness on the unsealed pavement without compromising performance. The

parametric study indicates that empirical design method used for stabilized unsealed pavements yield

an overly conservative design. Due to the improved bearing capacity from stabilization, the economic

aspects of pavements can be further improved by further reducing the base thickness whilst meeting the

subgrade strain criteria.

The efficiency of fly ash in the stabilization of expansive clay soils are governed by both densification

and pozzolanic reaction. Fly ash with secondary binders that improve density and forms compounds

that possess cementitious properties enhance the effectiveness of fly ash stabilization, highlighting the

efficiency of combining fly ash stabilized soils with other additives for improvement in pavement

performance. The fly ash-enzyme-lime-stabilized clay soils result in the most significant improvement

in bearing capacity arising from densification and pozzolanic reaction. Its superior performance is

highlighted by the 50% reduction in base thickness using traditional pavement design with stress, strains

and deformations marginally lower than a comparable pavement with double the cover. Furthermore,

the substantial improvement from stabilization can permit the base layer thickness to be further reduced

by 100 mm when subgrade strain criteria are evaluated.

6 Conclusions
26
This paper presented the experimental and numerical modelling results of exploring the efficiency of

fly ash stabilization on weak subgrades as applicable to unsealed pavements. The study investigated the

effects of secondary additives in enhancing the performance of weak expansive soils treated with Class

F fly ash. The results were explained using mineralogical and micrographic techniques, and pavement

performance under tested combinations was evaluated using numerical simulation. In this study, lime,

enzyme, CSA cement, and polymer were used as secondary additives in conjunction with fly ash for

stabilizing expansive clay soil. Following conclusions can be made on the basis of results obtained from

the study.

- Fly ash addition alone, regardless of its content, produced negligible improvements in CBR.

However, its efficiency was further enhanced with supplementary additives.

- Soil stabilized with fly ash-lime yielded a CBR of 26% due to the formation of pozzolanic and

cementitious products. The improved CBR resulted in a considerable reduction in base

thickness without a large impact on pavement stresses, strains and deformations.

- Fly ash-CSA cement stabilized subgrade soils resulted in a CBR of 19% arising from the

development of pozzolanic and cementitious products (including ettringite). Pavement

degradation is slightly larger (by about 13%) compared to fly ash-lime stabilized pavements.

- Soil-fly ash-enzyme stabilization exhibited improved compaction characterises with greater

MDD accompanied by a reduction in OMC. Improved density amounted to an 80% increase in

strength and 100 mm reduction in base thickness.

- Polymer addition to soil-fly ash stabilization resulted in a decrease in compaction

characteristics, causing a negative impact on CBR.

- Enzyme as a supplementary additive to secondary binders facilitates the compaction process to

increase MDD and decrease OMC. Therefore, soil-fly ash-lime-enzyme stabilized soils had the

best combination in terms of CBR (greater than 25%) and pavement performance due to the

combined effect of pozzolanic and cementitious formations and density enhancement.

The study evaluated a number of secondary additives in improving the performance soil

stabilization of expansive soils using bituminous fly ashes. It showed that the efficacy of fly ash
27
stabilization could be enhanced using green and cost-effective additives, as revealed from the

results of the current study. Such improved efficiency of fly ash stabilization has the potential to

divert waste material into geotechnical applications for sustainable construction practice. The

current study focusses on the effects of sustainable additives on the strength improvement of fly

ash stabilized soils. The study needs to be extended to investigate the curing and durability

characteristics of the fly ash stabilized soils using secondary additives. Furthermore, investigating

how the optimised stabilized technique affects hydraulic properties of the mixture is also

important.

7 Data Availability Statement

All data, models, and code generated or used during the study appear in the submitted article.

8 References

ANAND J. PUPPALA, E. W., LAUREANO R. HOYOS 2003. Ranking of Four Chemical and Mechanical
Stabilization Methods to Treat Low-Volume Road Subgrades in Texas. Transportation
Research Board, 63.
ANTOUN. 2019. CSA rapid set cement – the best choice for underground construction [Online]. New
South Welse. Available: https://antoun.com.au/csa-rapid-set-cement-the-best-choice-for-
underground-construction/ [Accessed].
ARMSTRONG, C. P. & ZORNBERG, J. G. 2018. Monitoring of Moisture Fluctuations in a Roadway over
an Expansive Clay Subgrade. IFCEE 2018.
ARORA, S. & AYDILEK, A. H. 2005. Class F fly-ash-amended soils as highway base materials. Journal of
Materials in Civil Engineering, 17, 640-649.
ARVIND KUMAR, B. S. W., ASHEET BAJAJ 2007. Influence of Fly Ash, Lime, and Polyester Fibers on
Compaction and Strength Properties of Expansive Soil. MATERIALS IN CIVIL ENGINEERING,
19, 8.
ASTM-D2487 2017. Standard Practice for Classification of Soils for Engineering Purposes (Unified Soil
Classification System). 17e1. West Conshohocken, PA: ASTM International.
AUSTRALIANSTANDARD 2006. Methods of testing soils for engineering purposes - Soil compaction
and density tests - Compaction control test - Hilf density ratio and Hilf moisture variation
(rapid method). AS 1289.5.7.1-2006.
AUSTROADS 2009. Guide to Pavement Technology Part 6. Unsealed Pavements.
BELL, F. 1996. Lime stabilization of clay minerals and soils. Engineering geology, 42, 223-237.
BLUEY, T. 2019. Bluey CSA Cement product information. Bluey Tech Pty Ltd.
CAPP, J. 1978. Power plant fly ash utilization for land reclamation in the eastern United States.
CHEN, F. H. 1988. Foundations on Expansive Soils, Elsevier.
CHEN, Y. J., ZHU, Y., SHI, F.X. 1992. Air pollution prevention during fly ash disposal. Geotechnical
Engineering, 23, 10.

28
COKCA, E. 2001. Use of class C fly ash for the stabilization of an expansive soil. Geotechnical &
Geoenvironmental Engineering, 127, 6.
CORREIA, A. G., WINTER, M. & PUPPALA, A. 2016. A review of sustainable approaches in transport
infrastructure geotechnics. Transportation Geotechnics, 7, 21-28.
DACHTAR, J. 2004. Calcium Sulfoaluminate Cement as Binder for Structural Concrete. University of
Sheffield.
DALE P. BENTZ, A. S. H., JOHN M. GUYNN 2011. Optimization of cement and fly ash particle sizes to
produce sustainable concretes. Cement & Concrete Composites, 33, 8.
DAYAL, U., JAIN, S.K., SRIVASTAVA, A.K. 1988. Geotechnical investigation of ash properties for dyke
construction at Kobra superthermal power project. Nat Therm Pow Corp, New Delhi, India.
DHAR, S. & HUSSAIN, M. 2019. The strength and microstructural behavior of lime stabilized subgrade
soil in road construction. International Journal of Geotechnical Engineering, 1-13.
DU, Y.-J., JIANG, N.-J., LIU, S.-Y., HORPIBULSUK, S. & ARULRAJAH, A. 2016. Field evaluation of soft
highway subgrade soil stabilized with calcium carbide residue. Soils and Foundations, 56,
301-314.
EKOSOIL 2015. Eko Soil Manual, Melbourne, Australia.
ESKIOGLOU, P. & OIKONOMOU, N. 2008. Protection of environment by the use of fly ash in road
construction. Global NEST Journal, 10, 108-113.
FAIL, J. L. & WOCHOK, Z. S. 1977. Soybean growth on fly ashamended strip mine spoils. Plant and
soil, 48, 473-484.
FAUZI, A., RAHMAN, W. M. N. W. A. & JAUHARI, Z. 2013. Utilization waste material as stabilizer on
kuantan clayey soil stabilization. Procedia engineering, 53, 42-47.
HOLTZ, W. G. & GIBBS, H. J. 1956. Engineering properties of expansive clays. Transactions of the
American Society of Civil Engineers, 121, 641-663.
HOY, M., HORPIBULSUK, S. & ARULRAJAH, A. 2016. Strength development of Recycled Asphalt
Pavement–Fly ash geopolymer as a road construction material. Construction and Building
Materials, 117, 209-219.
J.P. PRASHANTHA, P. V. S., A. SRIDHARANBA 2015. Pozzolanic fly ash as a hydraulic barrier in land
fills. GEOENVIRONMENTAL ENGINEERING, 60 8.
JALA, S. & GOYAL, D. 2006. Fly ash as a soil ameliorant for improving crop production—a review.
Bioresource Technology, 97, 1136-1147.
JANZ & JOHANSSON, S.-E. The Function of Different Binding Agents in Deep Stabilization Mårten.
2002.
JHA, A. K. & SIVAPULLAIAH, P. 2015. Mechanism of improvement in the strength and volume change
behavior of lime stabilized soil. Engineering Geology, 198, 53-64.
JIANG, N.-J., DU, Y.-J., LIU, S.-Y., WEI, M.-L., HORPIBULSUK, S. & ARULRAJAH, A. 2015. Multi-scale
laboratory evaluation of the physical, mechanical, and microstructural properties of soft
highway subgrade soil stabilized with calcium carbide residue. Canadian Geotechnical
Journal, 53, 373-83.
KANIRAJ, S. R. & HAVANAGI, V. G. 1999. Compressive Strength of Cement Stabilized Fly Ash-Soil
Mixtures. Cement and Concrete Research, 29, 5.
KATE, J. 2005. Strength and volume change behavior of expansive soils treated with fly ash.
Innovations in grouting and soil improvement.
KHANNA, S. & JUSTO, C. 1991. Highway engineering, Nem Chand & Bros.
KOLAY, P. K., DHAKAL, B., KUMAR, S. & PURI, V. K. 2016. Effect of Liquid Acrylic Polymer on
Geotechnical Properties of Fine-Grained Soils. International Journal of Geosynthetics and
Ground Engineering, 2, 29.
KUMAR, A., WALIA, B. S. & BAJAJ, A. 2007. Influence of Fly Ash, Lime, and Polyester Fibers on
Compaction and Strength Properties of Expansive Soil. Journal of Materials in Civil
Engineering, 19, 242-248.

29
KUMAR, S. & PATIL, C. 2006. Estimation of resource savings due to fly ash utilization in road
construction. Resources, conservation and recycling, 48, 125-140.
LATHA, G., NAIR, A. & HEMALATHA, M. 2010. Performance of geosynthetics in unpaved roads.
International Journal of Geotechnical Engineering, 4, 337-349.
LATIFI, N., VAHEDIFARD, F., GHAZANFARI, E., HORPIBULSUK, S., MARTO, A. & WILLIAMS, J. 2018.
Sustainable improvement of clays using low-carbon nontraditional additive. International
Journal of Geomechanics, 18, 04017162.
LEEA, S. H., KIMA, H. J., SAKAIB, E. & DAIMONB, M. 2002. Effect of particle size distribution of fly
ash–cement system on the fluidity of cement pastes. Cement and Concrete Research, 33, 6.
MAHVASH, S., LÓPEZ-QUEROL, S. & BAHADORI-JAHROMI, A. 2017. Effect of class F fly ash on fine
sand compaction through soil stabilization. Heliyon, 3, e00274.
MARECOS, V., SOLLA, M., FONTUL, S. & ANTUNES, V. 2017. Assessing the pavement subgrade by
combining different non-destructive methods. Construction and Building Materials, 135, 76-
85.
MARTIN, J. P., COLLINS, R.A., BROWING, J.S., BIEHL, F.J. 1990. Properties and use of ̄y ashes for
embankments. Energy Engineering 116, 15.
MISHRA, B. & GUPTA, M. K. 2018. Use of randomly oriented polyethylene terephthalate (PET) fiber
in combination with fly ash in subgrade of flexible pavement. Construction and Building
Materials, 190, 95-107.
MURMU, A. L., JAIN, A. & PATEL, A. 2019. Mechanical Properties of Alkali Activated Fly Ash
Geopolymer Stabilized Expansive Clay. KSCE Journal of Civil Engineering, 23, 3875-3888.
PAGE, A., ELSEEWI, A. A. & STRAUGHAN, I. 1979. Physical and chemical properties of fly ash from
coal-fired power plants with reference to environmental impacts. Residue Reviews. Springer.
PEETHAMPARAN, S. & OLEK, J. 2008. Study of the effectiveness of cement kiln dusts in stabilizing Na-
Montmorillonite clay. Journal of Materials in Civil Engineering, 20, 137-146.
PÉRA, J. & AMBROISE, J. 2004. New applications of calcium sulfoaluminate cement. Cement and
Concrete Research, 34, 6.
PHANIKUMAR, R. S. S. 2007. Volume Change Behavior of Fly Ash-Stabilized Clays. MATERIALS IN
CIVIL ENGINEERING, 19, 8.
POONI, J., GIUSTOZZI, F., ROBERT, D., SETUNGE, S. & O'DONNELL, B. 2019. Durability of enzyme
stabilized expansive soil in road pavements subjected to moisture degradation.
Transportation Geotechnics, 21.
POONI, J., ROBERT, D., GIUSTOZZI, F., SETUNGE, S., XIE, Y. & XIA, J. 2020a. Performance evaluation of
calcium sulfoaluminate as an alternative stabilizer for treatment of weaker subgrades.
Transportation Geotechnics, 27, 100462.
POONI, J., ROBERT, D., GIUSTOZZI, F., SETUNGE, S., XIE, Y. M. & XIA, J. 2020b. Novel use of calcium
sulfoaluminate (CSA) cement for treating problematic soils. Construction and Building
Materials, 260, 120433.
PUPPALA, A. J. 2016. Advances in ground modification with chemical additives: From theory to
practice. Transportation Geotechnics, 9, 123-138.
RAUCH, A., KATZ, L. & LILJESTRAND, H. 2003. An analysis of the mechanisms and efficacy of three
liquid chemical soil stabilizers: Volume 1. Evaluation of Nontraditional Soil and Aggregate
Stabilizers. Austin, TX: Cntr Trans Rese.
RAYMOND, S. 1961. Pulverized fuel ash as embankment material. 19, 22.
RENJITH, R., ROBERT, D., FULLER, A., SETUNGE, S., O’DONNELL, B. & NUCIFORA, R. 2017. Enzyme
based soil stabilization for unpaved road construction. MATEC Web Conf.
RENJITH, R., ROBERT, D. J., GUNASEKARA, C., SETUNGE, S. & O’DONNELL, B. 2020. Optimization of
Enzyme-Based Soil Stabilization. Journal of Materials in Civil Engineering, 32, 04020091.
RINTU RENJITH, D. R., ANDREW FULLER, SUJEEVA SETUNGE, BRIAN O’DONNELL, AND ROBERT
NUCIFORA 2017. Enzyme based soil stabilization for unpaved road construction. MATEC Web
Conf, 138, 10.
30
RIOS, S., CRISTELO, N., VIANA DA FONSECA, A. & FERREIRA, C. 2017. Stiffness behavior of soil
stabilized with alkali-activated fly ash from small to large strains. International Journal of
Geomechanics, 17, 04016087.
SALAHUDEEN, A. B., EBEREMU, A. O. & OSINUBI, K. J. 2014. Assessment of Cement Kiln Dust-Treated
Expansive Soil for the Construction of Flexible Pavements. Geotechnical and Geological
Engineering, 32, 923-931.
SCHOLEN, D. E. 1992. Non-standard stabilizers. Washington, D.C.; Springfield, VA: Federal Highway
Administration.
SHARMA, N. K., SWAIN, S. & SAHOO, U. C. 2012. Stabilization of a clayey soil with fly ash and lime: a
micro level investigation. Geotechnical and geological engineering, 30, 1197-1205.
SHEN, S.-L., ATANGANA NJOCK, P. G., ZHOU, A. & LYU, H.-M. 2021. Dynamic prediction of jet grouted
column diameter in soft soil using Bi-LSTM deep learning. Acta Geotechnica, 16, 303-315.
SINGH, D. N. 1994. Engineering properties of compacted Panki fly ash. In Proceedings of Indian
Geotechnical Conference, 1, 5.
SINGH, M., TRIVEDI, A. & SHUKLA, S. K. 2019. Strength enhancement of the subgrade soil of unpaved
road with geosynthetic reinforcement layers. Transportation Geotechnics, 19, 54-60.
SINGH, S. R., PANDA, A.P. 1996. Utilization of fly ash in geotechnical construction. In Proceedings of
Indian Geotechnical Conference, 1, 4.
SIVAPULLAIAH, P. V., PRASHANTH, J.P, SRIDHARAN, A., NARAYANA, B.V. 1998. Reactive silica and
strength of fly ashes. Geotechnical andGeological Engineering, 18, 13.
STANDARDSAUSTRALIA 2001. Methods of testing soils for engineering purposes - Sampling and
preparation of soils - Preparation of disturbed soil samples for testing. AS 1289.1.1.
STANDARDSAUSTRALIA 2014. Methods of testing soils for engineering purposes - Soil strength and
consolidation tests - Determination of the California Bearing Ratio of a soil - Standard
laboratory method for a remoulded specimen. AS 1289.6.1.1.
STANDARDSAUSTRALIA 2017. Methods of testing soils for engineering purposes - Soil compaction
and density tests - Determination of the dry density/moisture content relation of a soil using
standard compactive effort. AS 1289.5.1.1.
THOMAS, M. 2007. Optimizing the Use of Fly Ash in Concrete. Concrete Thinking for a Sustainable
World.
THOMAS, M. D. A., SHEHATA, M., SHASHIPRAKASH, S.G. 1999. The Use of Fly Ash in Concrete:
Classification by Composition. Cement, Concrete and Aggregates, 12, 6.
TRZEBIATOWSKI, B. D., EDIL, T. B. & BENSON, C. H. 2005. Case study of subgrade stabilization using
fly ash: State Highway 32, Port Washington, Wisconsin. Recycled Materials in Geotechnics.
WACKERCHEMIE. 2019. Dispersible Polymer Powders [Online]. Germany: Wacker Chemie AG.
Available: https://www.wacker.com/h/en-us/dispersible-polymer-powders/vinnapas-5010-
n-ger/p/000010668 [Accessed].
WANG, D., TAWK, M., INDRARATNA, B., HEITOR, A. & RUJIKIATKAMJORN, C. 2019. A mixture of coal
wash and fly ash as a pavement substructure material. Transportation Geotechnics, 21,
100265.
YLMÉN, R., JÄGLID, U., STEENARI, B.-M. & PANAS, I. 2009. Early hydration and setting of Portland
cement monitored by IR, SEM and Vicat techniques. Cement and Concrete Research, 39, 433-
439.
ZHANG, Y., JOHNSON, A. E. & WHITE, D. J. 2019. Freeze-thaw performance of cement and fly ash
stabilized loess. Transportation Geotechnics, 21, 100279.
ZORNBERG, J. G., AZEVEDO, M., SIKKEMA, M. & ODGERS, B. 2017. Geosynthetics with enhanced
lateral drainage capabilities in roadway systems. Transportation Geotechnics, 12, 85-100.
ZUMRAWI, M. M. 2015. Stabilization of pavement subgrade by using fly ash activated by cement.
American journal of civil engineering and architecture, 3, 218-224.

31
32
Highlights:

 Demonstrated the ability of secondary additives to enhance fly ash based soil stabilization.

 Identified soil-fly ash-lime-enzyme as the best combination to improve bearing capacity.

 Demonstrated that pavement thickness can be reduced when secondary additives are used.

 Revealed that addition of polymer has a negative impact on bearing capacity.

33

You might also like