Sunteți pe pagina 1din 75

CHAPTER 1

INTRODUCTION

1.1 BACKGROUND
Geopolymer is an emerging alternative binder to Portland cement for making concrete. A number
of supplementary cementations materials such as ground granulated blast furnace slag, metakaolin
and fly ash are commonly used source materials in geopolymer for their availability and favorable
mechanical properties (Pacheco-Torgal et al, 2014). Geopolymer is an inorganic polymer which is
synthesized by activating aluminosilicate materials with alkaline solutions. The polymerization
process initiates a chemical reaction of aluminosilicate minerals with alkaline activators and results
in a three dimensional polymeric chain. The chemical composition of source materials and alkaline
activators govern the chemical and microstructural properties of the final products of
geopolymerisation.

Concrete is the most commonly used building material and Portland cement constitutes its major
component. The reasons for Portland cement being the most attractive construction material are
its easy availability and cheap price. The demand for Portland cement is on an increase and for the
year 2014, the consumption was about 4.2 billion tons and is expected to rise further
(https://en.wikipedia.org/wiki/List_of_countries_by_cement_production). Climatic have posed
global warming as the most concerned issue of this decade. The emission of greenhouse gases is
responsible for the greenhouse effect. One of the major greenhouse gas is carbon dioxide as it
contributes about 65 % to global warming (McCaffrey, 2002). The cement production industry
emits 6 % of total CO2 emissions. One ton of cement production emits approximately one ton of
CO2 (Daridovits, 1994 and McCaffrey, 2002). Waste materials and industrial by-products like coal
ash, blast furnace slag, rice husk ash, silica fume etc are creating environmental hazards. A lot of
research has been done to replace or minimize the use of Portland cement by incorporating these
supplementary cementing materials as these waste by products are rich sources of silica, alumina
and calcium oxide.

Low-calcium fly ash-based geopolymer binder cured in high temperature has been reported to
1

have good mechanical properties in both short and long term tests (Wallah et al, 2005). The
structural behavior of heat-cured fly ash geopolymer concrete was found to be similar or superior
to that of OPC concrete when tested for reinforced columns and beams, bonding and fracture
properties (Sarkar et al, 2006). Curing conditions have a great influence on the strength and
microstructural development of fly ash based geopolymer. Low calcium fly ash based geopolymer
pastes harden slowly at ambient temperature and show lower strength in the early ages as compared
to heat cured samples (Duxson et al, 2007). Hence, these systems are usually subjected to heat
curing at temperatures ranging from 50C to higher and wide ranges of relative humidity for a
curing time ranging from several hours to several days. The high calcium fly ash geopolymer
mixture can be cured at room temperature as a result of the reaction of calcium in the system.
However, the geopolymerization of high calcium fly ash without additives is still slow at ambient
condition (Somna et al, 2011), therefore, low strength is normally obtained. The addition of
calcium oxide (CaO) forms hydrated products such as calcium silicate hydrates (CSH) along with
the alumino-silicate geopolymer network. Slag based by products are also used for geopolymer
production (Nath et al. 2014) . Ca-rich slags such as ggbs (ground granulated blast furnace slag)
react at room temperature since their reaction, i.e. the hydration of Ca species and the creation of
a calciumaluminiumsilicatehydrate (CASH) gel, is different from low-calcium precursors.
Moreover, the reaction develops at a very rapid pace, often resulting in a very short initial setting
time.
The commercial repair materials with good mechanical property and bonding strength are widely
used for repair work in concrete. However, the costs of these products are rather expensive.
Alternative repair material with comparable properties but less expensive is, therefore, a subject
of interest. A number of researchers have tried to utilize geopolymer as a repair material by testing
slant shear, pull-out and direct shear. Hu et al. (2008) studied the bond strength between mortar
substrate and geopolymer in sandwich specimens and reported that geopolymer exhibited higher
bonding strength than that of comparable Portland cement mixture.

The present research studies the mechanical behavior of fly ash and ground granulated blast
furnace slag (GGBS) based geopolymer mortar and possibility of application of geopolymer
mortar for use as Portland cement concrete (PCC) repair material.

1.2 RESEARCH OBJECTIVE

The primary objective of this study is to investigate utilization of FA and GGBS based geopolymer
mortar containing PC for use as an alternative Portland cement concrete repair material. In light of
above, following properties of FA and GGBS based geopolymer mortars have been have been
investigated:
1. Mechanical properties:

Compressive strength of geopolymer

Slant Shear Strength

Bending Strength

Cylindrical split tensile strength

2. SEM studies of the interface between Geopolymer mortar (GPM)/ repair mortar (RM) and
Portland cement concrete (PCC).
3. X-ray diffraction analysis to determine the formation geopolymer end products.
The tests conducted in this study were according to relevant IS codes and ASTM codes.

1.3 LAYOUT OF THESIS


The present thesis is divided into 5 chapters.
Chapter 1 provides general introduction about additive minerals and geopolymerisation
technology.
Chapter 2 gives the literature review on geoploymer materials and its properties and the research
conducted on geopolymers in the past years.
Chapter 3 presents the details the experimental programe. It includes the material characteristics,
mix proportions and preparation of samples, test procedures and parameters, various equipments
used in research.
Chapter 4 deals with the presentation, analysis and discussion of test results.
Chapter 5 summarizes the main conclusion and future scope of the study.
References have been provided at the end of the thesis for the convenience of the reader.

CHAPTER 2
LITERATURE REVIEW
2.1 GENERAL
This chapter presents a brief review on the terminology and chemistry involved of
geopolymersation, and additional review on the properties of geopolymers from the available
published research is also included.

2.2 INTRODUCTION TO GEOPOLYMER TECHNOLOGY


2.2.1 TERMINOLOGY AND CHEMISTRY

The term geopolymer was first introduced by Davidovits in 1978 to describe a family of mineral
binders with chemical composition similar to zeolites but with an amorphous microstructure. He
also suggested the use of the term poly(sialate) for the chemical designation of geopolymers
based on silico-aluminate (Davidovits, 1988a, 1988b, 1991; van Jaarsveld et. al., 2002a). Siliate is
the term used for siliconoxo-aluminate.
Poly(sialates) are chain and ring polymers with Si4+ and Al3+ in IV-fold coordination with oxygen
and range from amorphous to semi-crystalline with the empirical formula:
Mn (-(SiO2) z AlO2) n . wH2O
where z is 1, 2 or 3 or higher up to 32; M is a monovalent cation such as potassium or sodium,
and n is a degree of polycondensation (Davidovits, 1984, 1988b, 1994b, 1999).
Polysialates are further of three different types Davidovits (1988b; 1991; 1994b; 1999):
1. Polysialate (-Si-O-Al-O)
2. Polysialate-siloxo (-Si-O-Al-O-Si-O)
3. Polysialate-disiloxo (-Si-O-Al-O-Si-O)
These structures are schematically shown in the Fig 2.1

Si:Al>3 Sialate link


Figure 2.1: Chemical structures of polysialates (Davidovits 2002)
Geopolymerisation process is the chemical reaction of the alumino-silicate oxides with
alkaline polysilicates resulting in the polymeric Si-O-Al bonds. These two materials are the
two source materials for this process. The source for alumino silicates can be naturally
occurring minerals like clays, kaolinite. Waste by products can also be used like fly ash,
GGBS, RHA silica fume etc. The source for alumino-silicates must contain a large
percentage of silica and alumina. The source of the polysilicates are sodium silicates or
potassium silicates which are commercially available. Potassium/sodium or any other
soluble alkali metal is used as a source for alkaline liquid. Commonly used alkaline liquids
are NaOH and KOH. Combinations of sodium/potassium silicates and NaOH/KOH can also
be used. The equation in Fig. 2.2 (Davidovits 1994) shows the polycondensation by an alkali
into polysialalte :

Figure 2.2: Schematic formation of geopolymer (Davidovits, 1994: van


Jaarsveld et al., 1997)

Figure 2.3: Conceptual model for geopolymerization (Duxson et al. 2007)


6

It is revealed in this equation that water is formed in this reaction. This released water during the
geopolymer reaction is ousted from the matrix in curing which creates pores in the matrix which
are beneficiary to the geopolymer performance. There is no role of water in the geopolymeric
reaction unlike the Portland cement where water is necessary for the chemical reactions. This
expelled water just provides some workability.

The complete chemistry of geopolymerisation (Fig 2.3) can be theoreticallyconsidered to be


divided in two phases :
1. Dissolution-Hydrolyses phase
2. Hydrolysis- Polycondensation phase
In the dissolution-hydrolyses phase, the solid aluminosilicates are dissolves in alkaline medium by
the process of hydrolyses which consumes some water and produces Al3+ and Si4+ ions in the
solution. This solution already comprise silicates. The rate of this dissolution depends upon the
alkalinity of the activator solution and is more in Case of higher pH values i.e greater the presence
of OH- ions in the solution, greater is the amount of water that will carry hydrolyses (Zuhua et
al.,2009) . This concentrated solution of aluminosilicates thus results in the gel formation. This
chemical process of gel formation releases water which earlier was used in hydrolyses Fig 2.3
After the gel formation, the system observes a rearrangement of the gel which consist of monomers
(sialiates) and there is the formation of chains or polymerization of monomers by interconnection
of the gel particles. The Si4+ and Al3+ aluminosilicate geopolymeric gels are tetrahedrally
coordinated and linked by oxygen bridges as shown in Fig 2.1.

We know that in Portland cement, there is the formation of C-S-H gel which is responsible for the
strength. There is no C-S-H gel formation in the geopolymers. They derive their strength from
different types of polysialates formed by the reaction of the alkali polysilicates with aluminisilicates. Hence, they are also known as alkali-activated binders. (Davidovits, 1994a; Palomo et.
al., 1999; Roy, 1999; van Jaarsveld et. al., 2002a). However they completely different from alkali
activates cements or fly ash. This term must also not confused with alkali aggregate reaction, which
is a property of concrete (Davidovits (1999; 2005).

2.2.2 SOURCE MATERIALS

There two major ingredients of geopolymers are alumino-silicate source materials and an Alkaline
activator liquid. The source material must be abundant in silicon (Si) and aluminium (Al).
Naturally occurring minerals like kaolin, clays, etc which contains large proportions of Silica,
Alumina, and oxygen can be used as source material (Davidovits, 1988c). Alternatively, industrial
by-product such as fly ash, GGBS, iron slag, silica fume, rice-husk ash, etc can be utilized as
source materials. The selection of source material for producing a geopolymer can also depend
upon factors like availability of material, cost of the by-product, application type, or it may be user
specific. Low calcium (ASTM Class F) fly ash is preferred as a source material than high calcium
(ASTM Class C) fly ash. The presence of calcium in high amount may interfere with the
polymerisation process and alter the microstructure (Gourley2003).Van Jaarsveld et. al., (1997;
1999) carried out a study for potential application of waste by products like fly ash, contaminated
soil waste to make geopolymer. He found out that it is possible to form geopolymers that can be
used as a building material. In addition to that it helps preventing pollution due to dumping of
these waste materials. On the nature of the source material, it was stated that the calcined source
materials, such as fly ash, slag, calcined kaolin, demonstrated a higher final compressive strength
when compared to those made using non-calcined materials, for instance kaolin clay, mine tailings,
and naturally occurring minerals. However, Xu and van Deventer (Xu and van Deventer 2000)
found that using a combination of calcined (e.g. fly ash) and non-calcined material (e.g.kaolinite
or kaolin clay and albite) resulted in significant improvement in compressive strength and
reduction in reaction time. Cheng and Chiu (2003) conducted a research to investigate the fire
resistant properties of GGBS source material geopolymer. He concluded that the K2O percentage
in the system effects the properties of the geopolymer material. Increasing K2O the content
increases the setting time and the compressive strength. It was therefore proved that GGBS had
great potential as a geopolymer source material. Metakaolin is preferred by the niche geopolymer
product developers due to its high rate of dissolution in the reactant solution, easier control on the
Si/Al ratio and the white colour (Gourley 2003). However, for making concrete in a mass
production state, metakaolin is very expensive. Gourley (2003) Concluded that low calcium fly as
can be used as a source material for geopolymerisation. ASTM class F fly ash blended with slag
was used in this experiment. It was also stated that presence of significant amount of calcium in
8

fly ash could interfere with the microstructure of the final product. Therefore, low calcium fly ash
(ASTM Class F) is more suitable source material as compared to High calcium fly ash (ASTM
Class C). Swanepoel and Strydom (2002), Van Jaarsveld (2002a; 2002b) and Bakharev (2005a;
2005b; 2005c) also presented their research on fly ash as the source material for makimg
geopolymers. Dutta et al. (2010) incorporates silica fume in fly ash based geopolymer in various
proportions and investigated its effects on various physical and mechanical properties of the
resultant geopolymer. Aguilar et al. (2014) prepared geopolymer mortar based on low grade
metakaoline and found out that low grade metakaolin are very effective to be used as green binders
as they have low demand of alkalis and high strength. From the test result it was found that
Calcined kaolinitic minerals with relatively low content of reactive phase (50% MK50% quartz)
can be used to produce geopolymers having compressive strengths above 85 MPa after 28 days
and thus can be successfully used as a source material.

2.2.3 ALKALINE ACTIVATOR SOLUTION

The alkaline solution is generally produced from soluble alkaline metals and generally these are
sodium/potassium based. Sometimes alkali activator solution is made by mixing the alkali solution
with sodium silicate. A sodium silicate solution with its modulus less than 2 is alkaline. Davidovits
(1988c; 1988d) produced geopolymer incorporating kaolinite as source material and used NaOH
and KOH to prepare alkaline solution. The process involved in the production is known as
SILIFACE process . Davidovits (1999) later used calcined kaolinite and found that it performed
better than naturally occurring clays. Xu and Van Deventer (2000) conducted a study on different
aluminio-silicates source materials to produce geopolymers. Their research consisted sixteen
naturally occurring alumino-silicate minerals. The alkaline solution used in the research was made
from both sodium and potassium hydroxide. It was concluded from the research that the many
naturally occurring minerals can be potentially used as source material for geopolymer production.
Also test results using KOH as alkaline solution were better than NaOH. De Vargas et al. (2011)
did a research on the effects of the activator on different properties of the geopolymers. Three
variables were investigated and the Na2O/SiO2 molar ratio of the activator solution (N/S 0.20, N/S
0.30 and N/S 0.40) was one of them. The results depicted that the N/S molar ratio has a definitive
impact on the mechanical and morphological properties of geopolymers N/S ratio of 0.4 was
9

observed to give the best results and gave greatest strengths as well as denser matrix. As the N/S
molar ratio increased i.e more concentration of NaOH in the activator solution, extra water required
to make sure that the pastes and mortars acquire normal consistency decreased. This was associated
with greater fly ash particle solubilization caused by the alkaline activator, which enabled the
production of a greater amount of aluminosilicate gel. This gel was responsible for the greater
workability of fresh samples and for mechanical strength when the material hardened. Gorhan et
al (2014) studied the effect of different NaOH concentrations in the solution on the final
geopolymer product. These changes on the properties of the final geopolymer matrix were
investigated by applying different curing temperatures on different NaOH concentrations. The
objective of this process was to frame a relationship between alkaline solution concentration,
curing temperature and curing time. In order to determine the effect of NaOH concentration on
geopolymer mortars, three different molarities of NaOH concentrations (3 M, 6 M and 9 M) were
used together with sodium silicate (water glass) solution. Mijarsh et al (2015c) studied the effect
if Na2SiO3 concentration on the geopolymer mortar. NaOH and Na2SiO3 are the two components
of the alkaline activator. It was found that mixtures with higher Na2SiO3 concentration showed
positive results. Relation between water content and Na2SiO3 concentration was established and it
was observed that it only had effects on early strengths. Mijarsh et al (2015b) studied the possible
use of palm oil treated fly ash (TPOFA) and its effects on the strength of the geopolymer produced.
The geopolymer mortar was prepared using mixtures of TPOFA and calcium hydroxide, aluminum
hydroxide, and silica fume as mineral additives. The alkaline activator was prepared from NaOH
and Na2SiO3. The samples were also tested for different molar ratios of the activator and their
effects on the final geopolymer matrix formed.The samples were prepared using different total
oxide molar ratios (Al2O3:SiO2, CaO:SiO2, Al2O3:Na2O, and SiO2:H2O). The impact of
unreacted mineral additives in the final geopolymer matrix was also evaluated. It was concluded
in this study that TPOFA can successfully be used as a source mineral. Best results were given by
samples using TPOFA and calcium hydroxide. Tchakout et al (2016) prepared used rice husk
ash and waste glass as an alternative source to prepare water glass. This water glass was then mixed
with sodium hydroxide to prepare sodium water glass. The molar ratios SiO2/Na2O and
H2O/Na2O equal to 1.5 and 10. These alternative alkaline solutions were then used to prepare
metakaolin based geopolymer. It was found that the presence of calcium in the sodium waterglass
from waste glass increases the depolymerization of metakaolin particles and assist the formation
10

of geopolymers with long polysialate chains with high cross linking geopolymer framework. This
research has proven that sodium waterglass fromwaste glass and rice husk ash are suitable
alternative alkaline solutions for the production of metakaolin based geopolymer binders and could
replace sodium silicate solution obtained from mineral sources.

Every source material has advantages and disadvantages. For example, metokaolin as a source
material has high dissolvability in the reactant solution, produces a controlled Si/Al ratio in the
geopolymer, and is white in colour (Gourley, 2003). However, metakaolin is expensive to produce
in large volumes because it has to be calcined at temperatures around 500C 700C for few
hours. In this respect using waste materials such as fly ash is economically advantageous.

2.3 PROPERTIES OF GEOPOLYMERS

Duxson et al. (2007) studied the influence of the alkaline solution and Si/Al ratio on the
mechanical properties of metakaolin based geopolymers . The sodium silicate solution used in this
study had the modulus ratio (SiO2/ M2O)= Mu = 0,0.25,0.5,0.75,1 and the ratio of water to sodium
oxide (H2O/ Na2O) = 11. The solution was prepared by dissolution of silica in NaOH/KOH
solution. Cylindrical Specimens with 25 mm diameter and 50 mm length ( aspect ratio = 2 ) were
used. The results show that with increase from 1.15 to 1.9 in Si/Al ratios the compressive strength
increases but beyond 1.9 it shows a small decrease. At Si/Al of 1.15 with decrease in sodium
content, the compressive strength decreased. When Si/Al ratio is between 1.4 and 1.9 compressive
strength was found to be greater than both Na or K specimens. At Si/Al ratio of 1.15 and 2.15 Na
specimens were stronger and in intermediate compositions K specimens exhibited higher strength
values . It was further found that alkalies had very less effect on mechanical properties after 7 days
but after 28 days with Si/Al 1.9 in case of mixed alkali samples increased upto 30% but decreased
in comparison to Na and K specimens. It was concluded that mixed-alkali specimens with high
Si/Al ratio exhibited significant increases in strength, while pure alkali specimens displayed
decreased strength. The development of Youngs modulus of geopolymers between 7 and 28 days
was observed to be dependent on alkali, with the Youngs moduli of Na-specimens decreasing at
low Si/Al ratio, but increasing at high Si/Al ratio, while K-specimens exhibited the opposite effect.
Mixed-alkali specimens all exhibited nominal change in Youngs moduli, without any significant
11

effect of Si/Al ratio being observed.

Criado et al. (2007) discussed the influence of modulus ratio (SiO2 /Na2O) on alkali activated fly
ash . In order to activate fly ash four different alkaline solutions along with varying soluble silica
content were used with constant sodium oxide ( 8 % ). Only the silica content was varied. Prismatic
specimens of size 1 x 1 x 6 cm were used and cured for 8 hours, 7 , 28 , 60 , 90 and 180 days. The
results show that high strength could be achieved with increase in silica content at 8 h, whereas at
20 h increase in lower silica content specimens was substantial and equal to specimens having high
silica content. Final compressive strength of 70 MPa was observed in all samples which was
attributed to longer curing periods. In the secondary reaction zeolites are formed and percentage
of zeolites increased with curing in the presence of greater silica content. The formation of zeolites
is retarded which is attributed to high degree of polymerization of silica , therefore the ones with
high silica content had higher gel than the ones with lower silica content . The amount of zeolites
slightly declines at longer reaction times. These phases are partially attacked under the aggressive
conditions prevailing in such reactions.

Maharani et al. (2010) studied the effects of addition of alumino-silicate on the alkalisilica
reaction and the properties of final geopolymer product .Results depicted highest strength with
Si/Al ratio of 5 . At 7 days curing with Si/Al =5 compressive strength > 50 MPa was achieved and
at 14 days curing substantial increase was observed . Si/Al ratio was varied from 1.5 to 6 and
geopolymerization process was studied . It was found that with addition of corundum no significant
change in strength was observed , though SiO2 addition doubled the strength when compared to
corundum. It was proved that the reactivity of starting material affect the mechanical properties
where highly reactive starting material produces high mechanical strength. The lower reactivity of
the quartz, the interaction between the source material and the alkaline solution and the reinforcing
effect caused by the unreacted quartz particles give satisfactory mechanical strength of the formed
geopolymers. This study also indicates that not only amorphous phases but also crystalline phases
involve in the geopolymerization process.

Temuujin et al. (2010) studied the preparation and characters of geopolymer mortar based on fly
ash . For this varying levels of sand aggregate were used and binder to aggregate ratio was varied
12

from 9 to 1. The value of compressive strength was found to be 60 MPa which did not change upto
50 % wt sand aggregate addition. Increasing the proportion of aggregate in the mortar reduced the
amount of geopolymerisation but did not significantly impact on compressive strength. The
compressive strength of the geopolymer mortar depends on the strength of the geopolymeric gel,
the interfacial bonding between the geopolymeric gel and aggregate and to some extent the
aggregate itself. The results suggest that the interfacial bonding between aggregate and
geopolymer was comparable in strength to the geopolymer and/or the sand aggregate by
themselves. It is proposed that compressive strength of geopolymer mortars with high levels of
aggregate can be increased by optimising the amount of alkali.

Thokchom et al. (2011) carried out a research to study the effects of silica incorporated in fly ash
based geopolymer on physical, mechanical and durability properties. The percentage replacement
of silica was in between 2.5 to 5 %. Activator used for geopolymer preparation consisted of
solution of NaOH and water glass having modulus of 3.3. The specimens were prepared in 50 mm
3

cubes and were cured at 850C temperature for 48 hours. The strength of geopolymer mortar

incorporating silica fume (5%) was more as compared with 100 % FA specimen. However the
addition of silica fume increased porosity and water absorption.

Gorhan et al. (2014) carried out a study to find the influence of NaOH solution on various
properties of geopolymer made from fly ash. The variations in geopolymers mortar properties
induced by different curing temperatures and in different molarities of NaOH solutions. Effects on
various physical and mechanical properties were observed due to their variations. The activator
was produced by mixing NaOH and Na2SiO3 with modulus 3. Sand with fineness modulus 1.25
was used. It was observed that for low molarity solution the final strength produced was by even
for elevated temperatures due to weak chemical reactions. The compressive strength was expected
to improve for higher molarities of

NaOH solutions as the dissolution accelerates as the

concentration of OH- ions is is high for high molarities. The strength for 6 M solutions in this study
was highest. The strength was reduced slightly for 9 M solution as high alkalinity resulted in
coagulation of silica. A difficult pattern was followed by flexural strength values. There was a
general increase in flexural strength with increase in temperature. It was thus concluded that the
13

samples of the seven day geopolymer mortars which were activated by 6 M NaOH were observed
to have the optimum condition and the highest compressive strength values. When strength values
were considered the samples of the seven day geopolymer mortars which were activated by 6 M
NaOH were observed to have the optimum condition and the highest compressive strength values
when cured at 85C. It was determined that an increase in the curing temperature increased the
compressive strength while it did not have a significant effect on the physical properties.

Phoo-ngernkham et al. (2015) investigated the effects of NaOH and Na2SiO3 activator solution
on shear bond and compressive strength of fly ash and blast furnace slag geopolymer . The types
of specimen prepared were FA, FA+ GGBS, GGBS. They were activated with three different types
of activators namely NaOH , Na2SiO3 and solution of sodium hydroxide and sodium silicate
combined. 10 M NaOH solution and Sodium silicate solution with 11.67 % Na2O, 28.66 % SiO2
and 59.67 % H2O were used for producing the activator. The FA paste contains amorphous NASH
gel and some crystalline phases of the remain of fly ash. The increase in GBFS content enhanced
the compressive strength and microstructure of geopolymer pastes due to the formation of
additional CSH. The use of NaOH solution and sodium silicate solutions resulted in crystalline
CSH and amorphous gel, whereas the use of only sodium silicate solution resulted in mainly the
amorphous products. For the FA and FA + GBFS pastes, the use of NaOH solution or sodium
silicate solution alone gave low strengths when cured at ambient temperature. Better strength
development was obtained with the use of NaOH solution plus sodium silicate solution. For the
GBFS paste, the presence of silicate enhanced the strength development and thus pastes containing
sodium silicate solution performed better. Relatively high 28-day compressive strengths of 171.7
and 173.0 MPa were obtained for GBFS pastes with NaOH plus sodium silicate and only sodium
silicate solutions, respectively. The shear bond strength (slant angle of 45) between concrete
substrate and geopolymer paste was increased with the increase in compressive strength and
amount of NASH gel of geopolymer paste. The highest 28 day shear bond strength of 31.0 MPa
was obtained with FA + GBFS paste with NaOH plus sodium silicate solution. This indicates that
it may be possible to use FAGBFS geopolymer pastes as a repair material.

Mijarsh et al. (2015b) conducted a study to check the compressive strength of palm oil fuel based
geopolymer having calcium hydroxide, aluminium hydroxide and silica fume. NaOH and Na2SiO3
14

were used as alkaline activators . The compressive strength was tested for different total oxide
molar ratios. The effect of total unreacted minerals in final matrix were also studied. The material
used for this research were finally divided palm oil fuel ash. Ca(OH)2 , Al(OH )3 and silica fume
were used as mineral additives . The alkaline activator was prepared by mixing NaOH solution of
varying molarity with Na2SiO3 solution with modulus of 2. A total of six mixes were prepared. It
was observed that when Ca(OH)2 was used to substitute TPOFA there was an increase in the
compressive strength in initial 3 days but reduced at 28 days in comparison with control mixture
of TPOFA which showed an increase in strength throughout . However this trend was observed
only for replacement of Ca(OH)2 greater than 10%. This was due to the fact that formation of CS-H gel and N-A-S-H geoplymeric reaction compete for soluble silicates resulting in formation
of poor matrix. It was proved by XRD analysis. However the compressive strength

Figure 2.4 : XRD diffractograms (Mijarsh et al, 2015b)

was maximum when Al(OH)3, Ca(OH)2 and SF were substituted due to better interaction of C-SH and N-A-S-H gels formed. Also at longer curing temperature the strength was observed to
increase due to transformation of N-A-S-H gel and C-S-H gel into C-A-S-H gel which is a much
more stable product. These changes were in agreement with the XRD analysis shown in fig. 2.4.

Phoo-ngeinkham et al. (2015) did a research on possible use of high calcium fly ash incorporated
15

in OPC geopolymer s an alternative repair material . Bending Strength test and shear bond of
repair binder were compared with commercially available pre-mixed mortars. The geopolymer
repair mortar was prepared using high calcium fly ash. The alkaline activator was prepared with
NaOH and sodium silicate solution (13.89% Na2O, 32.15 % SiO2 and 46.04 % H2O). The
percentage replacement with OPC 0%, 5%, 10%, 15 % by weight with 3 molarities (6M, 10M,
14M). The GPM with high NaOH concentration containing PC as additive material gave good
performances in the shear bond strength prism test and bending stress of PCC notched beam test.
The highest shear bond strength of 24.2 MPa was obtained with 14 M NaOH geopolymer with
10% Portland cement (14M10PC mix). The bending stresses of PCC notched beams with filled
GPM were enhanced as expected. The GPM mix with 14 M NaOH and 10% PC gave excellent
bending stress of 3.1 MPa. However, with high NaOH concentration (14 M) and high PC (15%),
slight decreases in shear bond strength and bending stress were observed. The performance of
GPM was found to be comparable to those of the commercial repair materials. The average shear
bond strength of RM was 17.9 MPa, while that of GPM was slightly higher at 20.0 MPa.Test
results indicated that the use of GPM gave sufficiently high shear bond and bending strengths
compared with the use of RM suggesting that it could be used as an alternative product for concrete
repair works. In addition, the results from scanning electron microscopy of fracture surfaces
indicated that the interface zone of concrete and GPM was more homogeneous and denser than
that of concrete and RM. The GPM with 14 M NaOH solution and 10% PC was the optimum
mixture for improving the shear bond and bending strengths. The results indicated that the
performances of GPM containing PC with high NaOH concentrations as a repair material was at
least equal to those of commercially available ones according to the slant shear bond test and beam
notch filled bending test.

Alannazi et al. (2016) investigated the bond strength of metakaolin geopolymer used pavement
repair mortar. In order to characterize the bond strength of the repair material with the cement
mortar slant shear test and splitting tests were performed (Fig 2.5a). The alkaline activator was
prepared 24 hours prior to the experiment using NaOH and sodium silicate having modulus ratio
= 3.21. The specimens used for casting were 50 x 100 mm for testing of bond strength.

16

(a)

(b)

Figure 2.5: (a) Cylindrical split tensile test ; (b) Slant shear test with line interface of 30 and 45o
Alannazi et al. ( 2016 )

The specimen is half filled with mortar and half with geopolymer. The test was conducted with
two interfacial angles of 30 and 45o Fig 2.5(b). The prepared samples were cured at room
temperature. The SiO2 / NaO2 ratio was kept at 1.4. At ratio SiO2 / NaO2 = 1 greatest compressive
strength was observed. It is observed from the experiment that 80 % of the 28 days strength is
achieved within 3 days however the 28 days strength was less due to ambient temperature curing.
The slant shear strength obtained was more for 45o specimen. Through the splitting test conducted,
the experimental results indicated that the bond strength decreases as the cement mortar
deteriorates. The majority of the failure modes in the samples tested by splitting tests were always
an adhesion failure at the interface. The comparison between the suggested geopolymer mortars
with other pavement repair materials in the market shows the superior bond strength of the
suggested metakaolin based geopolymer to mortar substrate. The bond strength between the
cement mortar and the geopolymer mortar through the Slant shear test was also conducted. It is
found that the failure started at the mortar substrate and most of failure modes occurred through
mortar substrate, which indicates an excellent bond interface formed by geopolymer mortar with
cement substrate. However, this study has a limitation on the process of creating alkaline sodium
silicates solution. The increase of NaOH in sodium silicates solution may alter the
geopolymerization process, and therefore further study should be performed in the future.
Additionally, the process of mixing sodium hydroxide and sodium silicates should be further
studies to investigate its impact on NaOH action of forming silica and alumina and sodium silicates
17

action of polymerization.

Soutsos et al. (2016b) investigated factors that influence the compressive strength of geopolymers
based on fly ash. Samples were made of GGBS and fly ash from 13 sources along with NaOH and
sodium silicate having composition of 12.8 % Na2O, 25.5 % SiO2 and 61.7 % H2O were used. The
alkali modulus of 1.25 was fixed with varying percentage of activating solution. The solution was
made 24 h prior to experiment. 4 set of curing temperature were used and tested 1, 3 and 7 days of
curing. A total of 24 combinations of activator solution and alkali modulus were made for each
curing temperature. The effects of activator dosage, curing temperature and stand time on
compressive strength were studied. The samples were made by varying proportions of FA/GGBS.
Samples cured at elevated temperatures were stronger ans a stand time of 1 h was found to be
sufficient. The alkali modulus of 1.25 gave highest compressive strength amongst different ratios,
also early strength development depends on dosage of activator. Replacement of FA particularly
with GGBS increased the compressive strength because of dense microstructure as compared with
non-replaced mix.

Figure 2.6 : XRD diffractograms of raw materials and reacted pastes. (A) 100% FA (B)100% ggbs
Soutsos et al . (2016b)

This was due to production of both C-A-S-H and N-A-S-H. The N-A-S-H gel of GGBS was found
to be denser than FA made from NaOH. With 50/50 FA/GGBS homogeneous dense gel was
observed with less micro cracking an better bonding to unreacted GGBS and FA particles whereas
18

the ones with 100 % GGBS and 10 % FA resulted in coarser granularity and desiccation cracking
respectively. Partial FA replacement with ggbs leads to increases in the compressive strength. The
other benefit from such blends is that curing at room temperature only is sufficient and no elevated
curing temperatures are needed. A calcium alumina silicate hydrate gel with inclusions of Na+
cations in the structure was found in the samples containing ggbs. This was denser than the sodium
alumina silicate hydrate gel (NASH) found for 100% FA samples and this could explain the
improved compressive strengths. This is explained in the XRD result as shown in Fig 2.6.

Wee Zhou et al. (2016) investigated the influence of high and low Al2O3 content fly ash based
geoploymer at different curing temperature. The objective of this study was to check whether fly
ash with low Al2O3 content could be used as alternative material source for geopolymerization by
balancing Si/Al ratio from additional sources. The variables in this study were Si/Al ratio, variation
in modulus of alkaline activator, water/solid ratio, curing temperature and their effects on
compressive strength were investigated. For microstructure analysis various tests such as SEMEDS , XRD , TG-DSC and FT-IR were carried out . In this experiment two different kinds of low
calcium fly ash were used. The alkali activator used consisted of liquid sodium silicate having 26
% SiO2, 8.2 % Na2O and 5.6 % H2O with modulus 3.2 and sodium hydroxide in granular form. 24
hours after the preparation of geopolymer moulds were kept for incubation for 7 days at a fixed
temperature. The compression testing was done on 20 mm3 specimens. It was indicated from the
results that 7 days that the strength of samples was 70-95 % of the 28 days strength. The results
show that geopolymerization of fly ash was activated when low Al2O3 fly ash was used with
mixture of NaOH and sodium silicate and compressive strength of 27 MPa was attained at 7 days.
Geopolymers modified with high Al2O3 fly ash possess superior performance than that of low
Al2O3 fly ash modified samples, in terms of compressive strength and microstructure. This study
proves that the application of low Al2O3 fly ash can be used as a substitute material for geopolymer
cement manufacture. Geopolymerization results were well validated with SEM, XRD which shows
dense structures in geopolymer products. SEM and FTIR combined together concluded that low
modulus of activator fly ash reacted partially but with more than 1.8 modulus of alkali activator
microstructure of geopolymer developed cracks. High Al2O3 fly ash geopolymera as compared to
low Al2O3 had superior properties which were attributed to low mass loss and higher strength due
to high temperature curing.
19

2.4 FIELDS OF APPLICATIONS


Geopolymeric materials have a wide range of applications in the field of industries such as in the
automobile and aerospace, non ferrous foundries and metallurgy, civil engineering and plastic
industries (Davidovits 1988b). The application of geopolymer materials is determined by its
chemical structure in terms of the atomic ratio Si:Al in the polysialate. Davidovits (1999) classified
the type of application according to the Si:Al ratio as presented in Fig 2.7. A low ratio of Si:Al of
1, 2, or 3 initiates a 3D-Network that is very rigid, while Si:Al ratio higher than 15 provides a
polymeric character to the geopolymeric material. It can be seen from Fig 2.7 that for many
applications in the civil engineering field a low Si:Al ratio is suitable.

Figure 2.7: Applications of Geopolymeric Materials Based on Si:Al ratio.(Davidovits 2002)

20

Toxic Waste Management


One of the potential fields of application of geopolymeric materials is in toxic waste management
because geopolymers behave similar to zeolitic materials that have been known for their ability to
absorb the toxic chemical wastes (Davidovits, 1988b).
Geopolymeric Cement
Geopolymer cements are a class of material that combine an aluminum silicate with a chemical
activator such waterglass. Aluminum silicates include a variety of naturally occurring clays as well
as industrial byproducts such as blast furnace slag and fly ash from coal combustion. The new
terminology was the key to the successful development of new materials. For the user,
geopolymers are polymers and, therefore, by analogy with the organic polymers derived from oil,
they are transformed, undergo polycondensation, and set rapidly at low temperature, within few
minutes. But they are, in addition, GEO-polymers, i.e. inorganic, hard, stable at temperature up to
1250C and non-inflammable. GEOPOLYMITE, TROLIT and WILLIT binders are some of
the first patented geopolymer binders.

Geopolymer Brick, low cost construction material


This brick uses a very cheap material available in great quantity: lateritic clay earth. This special
and abundant earth, mixed with a simple geopolymer binder is compressed to give the shape of a
brick then heated in a furnace. Heated at 85C, this brick is water stable and has enough
compressive strength to build a wall. Heated at 250C, it resists to freezing. At 450C, its strength
increases more, so that it is possible to manufacture structural elements like beams for doors and
windows. Compared to a traditional brick fired at 1000C in a kiln, the geopolymer brick needs
about eight times less energy for an equivalent strength. Contrary to a traditional brickyard,
it requires less equipment and is less expensive to produce. A traditional brickyard must have a
certain size before being profitable, whereas geopolymer brick can be produced by small
brickyards in a village or a small city with less equipment and finance.

21

Retrofitting of Existing Structures


Another application of geopolymer is in the strengthening of concrete structural elements.
Balaguru et. al. (1997) reported the results of the investigation on using geopolymers, instead of
organic polymers, for fastening carbon fabrics to surfaces of reinforced concrete beams. It was
found that geopolymer provided excellent adhesion to both concrete surface and in the interlaminar
of fabrics. In addition, the researchers observed that geopolymer was fire resistant, did not degrade
under UV light, and was chemically compatible with concrete.
Aviation applications
Some aircraft cabins are made from geopolymeric materials (GOPOLYMRE). This include
cargo liners, ceiling,floor panels, partitions and sidewalls, stowage bins, wire insulation, etc. There
is an increase demand for fire-resistant containers in this industry.

22

CHAPTER 3
EXPERIMENTAL WORK AND METHODOLOGY
3.1. INTRODUCTION
The details of the experiment procedure planned for the research work has been presented in this
chapter. The experiment work has been planned in two phases. The first phase includes the
characterization of materials used in the investigation. The second phase includes tests on the
mortar specimens.

3.1.1 OUTLINE OF THE RESEARCH PROGRAM


To achieve the objectives of the present investigation, following phases of the planned program
were executed:1. Determination of the physical and chemical properties of the FA,GGBS,OPC and fine
aggregates.
2. Preparation of the alkaline solution of three different molarities (8M, 10M, 12M) and then
preparing the activator solution by mixing these alkaline solutions with sodium silicate.
3. Preparation of mix combinations for the two source materials used i.e. FA ang GGBS.
4. Casting of geopolymer specimens for compression testing in 50x50x50 mm cubes.
5. Casting of prism samples of size 50x50x125mm for slant shear testing.
6. Casting of beam specimens of size 75x75x300 mm with a notch of size 12x30 mm in the
middle of the beam to obtain bending strength.
7. Preparation of split tensile strength specimens in cylinders of size 100x200 mm.
8. Testing of various samples as per relevant IS/ASTM codes and comparing the results of
the fly ash samples and GGBS samples with two commercial repair mortars.
9. Analysis of failed mortar samples using SEM, and XRD results.

3.1.2 MATERIALS
The materials used for making geopolymer mortar specimens are low-calcium dry fly ash and
ground granulated blast furnace slag as the source materials, fine aggregates, alkaline activator
made from sodium hydroxide and sodium silicate, distilled water and OPC 43 grade.
23

3.2 SOURCE MATERIALS


There are two main constituents of geopolymers, namely the source materials and the alkaline
liquids. The source materials for geopolymers based on alumino-silicate should be rich in silicon
(Si) and aluminium (Al). The source materials used in this research are low calcium fly ash (FA)
(class F) and ground granulated blast furnace slag(GGBS).

3.2.1 FLY ASH


Fly ash used in this study was low-calcium (ASTM Class F) dry fly ash from Hopper number 7 of
Rajpura Thermal Power Plant, Punjab. The fly ash used was obtained from a single batch from
one source only. The chemical composition of the three batches of the fly ash, given in Table 3.1,
was determined by EDX analysis. As can be seen from Table 3.1 that the silicon and aluminium
constitute about 80% of the total mass. Enlarged particles of it are shown in an SEM image
(Fig.3.1a). It shows that the shape of particles is almost spherical. This fly ash contains high levels
of amorphous silica and alumina.

(a) FA

(b) GGBS

Figure 3.1 : SEM image of sample of source materials at 10m scale.

3.2.2 GROUND GRANULATED BLAST FURNACE SLAG


In the present study, a product Alcofine of Counto Microfine Products Pvt. Ltd., has been used as
is shown in Fig 3.1b.The SEM images of GGBS indicate the angular shape of its particles. The
chemical composition obtained from the EDX analysis is given in table 3.2. The GGBS used in
24

this research was in accordance with the IS: 12089 -1987. It was observed from the chemical
analysis that this GGBS is rich in calcium.

Table 3.1: Chemical composition of the Fly Ash used in this study

Constituent

Compound (%)

IS 3812-2003 :
Requirements for Class
F Flyash

SiO2

55.49

35% minimum

Al2O3
Fe2O3
SiO2+ Al2O3+
Fe2O3
CaO
Na2O
MgO
CuO
K2O

36.41
1.6

87.64

70% minimum

0.01
0.32
0.73
1.87
3.32

1.5% maximum
5% Maximum
-

Table 3.2: Chemical composition of GGBS used in this study

Constituent

Compound (%)

IS : 12089-1987
Specifications for
GGBS

SiO2
Al2O3
Fe2O3
CaO
Na2O

38.13
20.78
1.6
28.82
0.57

MnO
MgO
SO3

0.68
7.51
0.18

5.5% maximum
17% maximum
2% maximum

3.2.3 ORDINARY PORTLAND CEMENT


Ordinary Portland Cement (OPC) of grade 43 manufactured by JK Lakshmi Cement Company
satisfying IS 8112:1989 was used in this research. The cement is of uniform colour and was free
25

from hard lumps. Various physical properties of cement are shown in Table. It can be observed
from the table that the physical properties of the cement complies with the specifications of IS
8112: 1989. The chemical composition of cement used is presented in Table 3.5.

Table 3.3: Chemical composition of cement observed present study.


Constitutuent

Composition

IS 8112: 1989 Specifications

CaO

63.49%

SiO2

21.24%

Al2O3

4.75%

Fe2O3

4.30%

SO3

2.92%

Max. 3.5 %

MgO

1.02%

Max 6 %

K 2O

0.78%

TiO2

0.36%

BaO

0.32%

Na2O

0.30%

P2O5

0.21%

Cl

0.09%

Max 0.1 %

MnO

0.08%

SrO

0.04%

Table 3.4: Physical properties of Cement used in the study


Property

Value

IS 8112:1989 Specifications

Grade

OPC-43

OPC-43

Specific Gravity

3.15

3.10-3.25

Initial Setting Time

134 minuts

Minimum 30 minuts

Final Setting time

260 minuts

Maximum 600 minuts

28 days Compressive Strength

46 Mpa

43-58 Mpa

26

3.3 FINE AGGREGATES

The fine aggregates used for the experimental work is locally procured and conformed to grading
zone IV .Sieve analysis of the fine aggregate is carried out in the laboratory as per IS 383-1870.The
fineness modulus of the fine aggregates was 1.38. The grading of the aggregates is as presented in
Table 3.6 and the plot is presented in fig 3.3. Sieves of the sizes given in Table 3.6, conforming to
IS : 460-1962 specification for Test Sieves were used. The weight of sample available shall be not
less than the weight given in 200gms. The sample for sieving taken for this test was 500gms. The
sample were brought to an air-dry condition before weighing and sieving. This was achieved by
heating the sample at a temperature of 100C. The air-dried sample were weighed and sieved on
the appropriate sieves starting with the largest. The sample was shaken on sieve shaker for 10
minutes and the weight retained on each sieve was measured. The results were obtained and are
shown in table and fineness modulus was calculated. Other physical properties of fine aggregates
are presented in Table 3.5.

Figure 3.2: Sieve Analysis setup


27

Table 3.5: Physical properties of fine aggregates


S.No

Property

Result

1
2
3
4

Specific gravity
Fineness modulus
Bulk density
water absorption

2.51
1.38
1.35g/cc
0.85

Table 3.6: Sieve analysis of fine aggregates


Sieve Size

Mass of each
retained

4.75mm
0
2.36mm
0
1.18mm
10
600
10
300
140
150
340
Pan
0
Total weight of sample taken: 500gm

Percentage
of mass
retained
0
0
2
2
28
68
0

Cumulative
percentage

Percentage Finer

0
0
2
4
32
100
100

100
100
98
96
68
0
0

Figure 3.3 : Sieve analysis graph


28

Fineness modulus :

2+4+32+100
100

= 1.38

3.4 REPAIR MORTAR


Two commercially available repair mortars were used in this study. These were Sika TOP 112
HS (Fig. 3.4) and BASF MasterFlow. Sika TOP 112 HS is general purpose non-shrink grout
mortar and BASF MasterFlow is fiber-reinforced non-shrink mortar. Both of these repair mortars
are high strength quick setting repair mortars which gain about 70% of their characteristic strength
within 24-48 hours.

Figure 3.4 : Sika TOP 112 HS repair mortar

3.5 ALKALINE ACTIVATOR SOLUTION


The alkaline liquid used was a combination of sodium silicate solution and sodium hydroxide
solution.Sodium based activator was selected as it is cheaper than Potassium based activator.
Usually in industry, a sodium silicate solution is characterized by its SiO2/Na2O weight ratio in a
range of between 2 and 3.75. The SiO2/Na2O ratio greater than 2.85 will classify solution as neutral.
A solution with SiO2/Na2O ratio less than 2.85 is alkaline. Popular solutions produced in industry
are with a range of SiO2/Na2O ratio from 1.6 to 3.3. The sodium silicate solution (Na2O= 19.7%,
SiO2=59.4%, and water =20.9% by mass) was purchased from a local supplier. The sodium
hydroxide (NaOH) used in this research was in flakes from having 97%-98% purity. The NaOH
29

solids were dissolved in water to make the solution and was mixed with sodium silicate to produce
activator solution.

3.6 MIX PROPORTIONING

3.6.1 DRY MIX


Twenty four mix combinations categorized under two different series of FA and GGBS were
prepared to meet the objectives of the study. Each series consist of 12 mix combinations of 3
molarities (8M, 10M, 12M) and 3 replacements of 10%, 15% and 20% by weight with OPC. A
fixed ratio of sand to binder = 1 was used. The ratio of sodium silicate solution to alkaline solution
(NaOH) of 2 was fixed for every molarity. Constant liquid alkaline activator to binder ratio was
0.60. No variations in these two ratios was made throughout the research. This dry mix
proportioning was done on the bases of previously published literature ( Hardjito et. al., 2002a;
Hardjito et. al., 2003,; Rangan et. al., 2005; Duxson et al. (2007); Gorhan et al. (2014); ngernkham
et al. (2015); Krkl et al. (2016). These mixes are presented in table 3.7 and table 3.8.
Table 3.7: Mortar Mixture Proportion (kg/m3)containing Fly Ash
MIX

FA

OPC

Sand

F8M0PC
F8M10PC
F8M15PC
F8M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

F10M0PC
F10M10PC
F10M15PC
F10M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

480.64
480.64
480.64
480.64

961.29
961.29
961.29
961.29

F12M0PC
F12M10PC
F12M15PC
F12M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

480.64
480.64
480.64
480.64

961.29
961.29
961.29
961.29

30

NaOH(solution)
8M
10 M
12 M
480.64

480.64

480.64

480.64

Na2SiO3aq
961.29
961.29
961.29
961.29

Table 3.8: Mortar Mixture Proportion(kg/m3) containing GGBS


MIX

GGBS

OPC

Sand

NaOH(solution)

Na2SiO3aq

G8M0PC
G8M10PC
G8M15PC
G8M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

8M
480.64
480.64
480.64
480.64

10 M

12 M

961.29
961.29
961.29
961.29

G10M0PC
G10M10PC
G10M15PC
G10M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

480.64
480.64
480.64
480.64

961.29
961.29
961.29
961.29

G12M0PC
G12M10PC
G12M15PC
G12M20PC

2403.23
2160.01
2040
1920

240
360
480

2403.23
2400.01
2400
2400

480.64
480.64
480.64
480.64

961.29
961.29
961.29
961.29

3.6.2. PREPARATION OF ALKALINE ACTIVATOR

The sodium hydroxide (NaOH) solids were dissolved in water to make the solution. The mass of
NaOH solids in a solution varied depending on the concentration of the solution expressed in terms
of molar, M. To prepare NaOH solution with a concentration of 8M consisted of 8x40 = 320 grams
of NaOH solids (in flake form) per litre of the solution, where 40 is the molecular weight of NaOH.
The mass of NaOH solids was measured as 262 grams per kg of NaOH solution of 8M
concentration. Similarly, the mass of NaOH solids per kg of the solution for 10M and 12M
concentration was measured as 314 grams and 361 grams respectively (Table 3.9) . Note that the
mass of NaOH solids was only a fraction of the mass of the NaOH solution, and water was the
major component. The sodium silicate solution and the sodium hydroxide solution were mixed
together 24 hours prior to use to prepare the activator solution as presented in Fig 3.5.

31

Table 3.9:Percentage of NaOH Flakes in Various Molarity


NaOH
solution

Percentage (%)

8M

26.23

10M

31.37

12M

36.09

(a)

(b)

(c)

(d)

Figure 3.5: (a) NaOH flakes, (b) alkaline solution, (c) Sodium Silicate solution,
(d) Alkaline Activator solution

32

3.7 MANUFACTURE OF MORTAR AND CASTING

Figure 3.6 : Flow chart representing manufacture of Mortar and Casting

Fly ash/GGBS, Portland cement and the fine sand were first mixed together for about 3 minutes.
Davidovits (2002), suggested that it is preferable to mix the sodium silicate solution and sodium
hydroxide solution before adding it to the solid constituents. It was also suggested in his findings
that the sodium silicate liquid obtained from the market usually is in the form of a dimer or a
trimer, instead of a monomer, and mixing it together with the sodium hydroxide solution assists
33

the polymerization process. Hence, following this mixing procedure, the liquid component of the
mixture was then added to the dry materials and the mixing continued for further about 4 minutes
to manufacture the fresh mortar. The geopolymer mortars were mixed manually in a laboratory
pan to obtain a uniform mixture. The combined alkaline solution and source minerals (FA/GGBS)
produces a sticky mixture due to high viscosity of sodium silicate. In the absence of substantial
amount of additional water or super plasticizer, this solution usually forms a thick and cohesive
paste with fly ash and GGBS but the addition of sodium silicate increase the workability and setting
time. Thus, the mixture of aggregates and geopolymer paste becomes highly cohesive. The mixture
can be quite stiff when the liquid content is relatively low. The fresh mortar (Fig 3.7) was then cast
into the moulds immediately after mixing and compacted by vibrating the moulds for 30 seconds
on a vibrating table. It was observed during the experimentation that inclusion of OPC decreased
the setting time drastically. The early improvement of strength due to OPC inclusion can be
attributed to the increase of dissolute binder which produced reaction product from both alkali
activated fly ash and OPC. This observation was in line with the research findings of Nath et al.
(2015) that inclusion of OPC greatly reduced workability and setting time . The mortar paste
became stiffer for every increase in percentage of OPC replacement in fly ash based geopolymer.
However for GGBS based geopolymer , both the workability and setting time were observed to be
very low. This was due to the fact that GGBS is easily activated by alkaline solution due to its
relatively flat particles. The reaction occurs immediately which greatly reduces the setting time as
well as workability even at room temperatures.

Figure 3.7: Freshly prepared Geopolymer mortar


34

(a)

(b)

Figure 3.8 : (a) Prepared mortar samples wrapped in polythene wraps for curing ; (b)
Mortar sample after 28 days of curing

(a)

(b)

Figure 3.9: Fresh prepared (a) Notched Beam specimen (b) Prismatic sample

35

3.8 TESTING OF SPECIMENS


3.8.1 COMPRESSIVE STRENGTH TEST

For compression testing of mortar cubes, a set of cubes were made (Fig 3.8b). The size of cubes
was 50x50x50 mm (IS: 2250 1981). The prepared mortar cubes were wrapped in polythene
sheets and were kept were allowed to cure at room temperature as shown in fig 3.8a. The average
room temperature was around 25-30C at the time of casting.

The cubes were tested in

compression in accordance with the test procedures given in the IS: 2250 1981. The requirements
of the machine used for testing of concrete and mortar test specimens in compression were
according to IS 14858:2000. The sample was placed as shown in fig3.10 and a typical failure
pattern of cubes is presented in fig3.11. Load at the failure divided by area of specimen gave the
compressive strength of concrete. The loading was applied at a constant rate of 140kgf/cm2 per
minute.
=P/A
Where = Compressive Strength (N/mm2)
P = Maximum load (N)
A = Cross section area of cube (mm2)

Figure 3.10: Compression testing of cube


36

Figure 3.11: Failed Compression testing specimen

3.8.2 SLANT SHEAR TEST


The shear bond strength was evaluated using the slant shear test of PCC substrate and Geopolymer
mortar or repair mortar as described in ASTM C882 with stiffer slant shear angle of 45. The slant
shear angle of 45 is officially used for standard evaluation of epoxy bond with concrete and was
also used successfully in testing shear bond of concrete and geopolymer (Phoo-ngernkham et al.
2015). Figure 3.12 represents specimen dimensions and point of load application on the
specimen.For casting of the specimens, the GPM was placed in two equal layers into a 50x50x125
mm prism mold, half-filled with slant PCC substrate. The slant PCC substrate was prepared 28
days prior to the filling. Each layer was tamped 25 times and vibrated for 45 seconds. The samples
were covered with polythene wraps and kept for curing at room temperature until test date.The
specimens (as shown in Fig 3.13) were tested in a constant loading rate of 180 kgf/cm2 per minute
(~0.30MPa/sec). The shear bond strength was the ratio of maximum load at failure and the bond
area.

37

Figure 3.12: Test setup for slant shear testing at 45 angle (Ngernkham et. al., 2015)

Figure 3.13: Slant Shear testing of Prismatic specimen

3.8.3 BENDING STRESS TEST


This test used the same type of specimens as in the test of fracture characteristics as shown in Fig
3.14. For casting of specimens, the GPM or RM was mixed and filled in the notch to act as repair
38

materials. For the preparation of specimens for bending stress, the PCC was cast in 75x75x300
mm long beams.

Figure 3.14: Test set up of bending stress of PCC notched beam with filled GPM or RM
specimens. (Ngernkham et. al., 2015)

They were cured similarly to the PCC substrate in the slant shear test specimens. A notch with
height to beam depth (ao/d) ratio of 0.4 and notch width to notch height (wo/ao) ratio of 0.4 was cut
in the middle of beam. The dimensions of the notch were ao = 30mm and wo = 12mm.The notch
was then filled separately with geopolymer and repair mortars. The samples were covered with
polythene wraps to retain the moisture and kept for curing at room temperature until test date. The
specimens were tested by three point bending with deflection control using loading rate of 0.05
mm/min (Bharatkumar et al. 2005)

(a)
39

(b)
Figure 3.15: (a) Bending stress beam setup (b)Failed Bending Stress Specimen

Fig 3.15a represent load being applied on a notched beam. Typical failure pattern of geopolymer
mortar is presented in fig3.15b.
3.8.4 CYLINDRICAL SPLIT TENSILE TEST
The splitting strength of the composite cylinder was used to measure the bond strength between
the mortar or PCC substrate and the repair material (Geissert et al. 1990). Split tensile test (ASTM
C496/C496M-11) is a standard test method of split tensile strength of homogeneous cylindrical
specimen. This test can also be used for composite cylinders, made of half mortar/PCC substrate
and half repair material. The specimen of splitting tensile test was prepared by casting a 100x200
mm. cylinder for bond strength testing. The cylinder was longitudinally divided in half. One half
contained PCC and cured for 28 days at room temperature. The PCC has the same mix design for
all specimens. The cured cement mortar resulted in a compressive strength of 35.0 MPa. After the
halves of cement mortar specimens cured for 28 days, they were put back in the molds and the
other empty halves of the mold were filled with geopolymer mortar after 28 days of cure period of
PCC substrates. The sample were tested after 28 days under compression testing machine as shown
in figure. The bond strength was calculated by the following:
= 2P/A
P = maximum load in Newtons
A = L x d is the area of bond interface (mm2)
40

L = length of the specimen (in mm), and d = cross sectional dimension of the specimen
The test specimen and apparatus are shown in fig 3.16. Typical failure pattern is shown in figure
3.17.

Figure 3.16: Cylindrical Split Tensile Test

Figure 3.17: Failed cylinder split-tensile Strength Specimen

41

CHAPTER 4
EXPERIMENTAL RESULTS AND DISCUSSION
4.1. GENERAL
In this Chapter, the mechanical properties and microstructural characteristics of FA and GGBS
based geopolymer mortar are presented and discussed. The test results cover the results obtained
from compression testing, slant shear test, bending stress test and cylinder split tensile strength of
fly ash and GGBS based geopolymer mortars and repair mortars. Table 4.1 to 4.11 presents the
mean values of results obtained from at least three specimens.

4.2. COMPRESSIVE STRENGTH OF GEOPOLYMER SAMPLES


For each batch of geopolymer mortar made in this study, 50x50x50 mm cube specimens were
prepared. At least three of these cubes were tested for compressive strength at an age of seven days
after casting.

4.2.1 COMPRESSIVE STRENGTH OF FLY ASH BASED GEOPOLYMER SAMPLES


60

50

47.87

44.19

Strength (MPa)

41.42

41.47

40

39.11

36.44
31.19

30

32.91
31.42

27.79
25.1

20

11.97
10

14.4

8.64
4.11
0

0
0

10

20

0%OPC

10%OPC

15%OPC

20%OPC

30

40

50

Curing days

Figure 4.1 : Compressive strength of 8M FA samples


42

60

60

50

47.14

49.26
45.93
45.15

44.89

42.48

Strength (MPa)

40

38.54

37.42

36.56

32.51
30

24.54
20

15.82
14.32
11.24
7.37

10

0%OPC

10%OPC

15%OPC

20%OPC

0
0

10

20

30

40

50

60

Curing days

Figure 4.2 : Compressive strength of 10M FA samples

60

53.54
50.44

50.63

50

47.74

48.24
43.35

Strength (MPa)

44.59
40

38.91
31.22
28.63

30

20

40.47

20.07
16.51
11.68

18.26

10

0%OPC

10%OPC

15%OPC

20%OPC

3.31
0
0

10

20

30

40

50

Curing days

Figure 4.3 : Compressive strength of 12M FA samples


43

60

Table 4.1: Percentage increase in Compressive strength of FA based geopolymer

8M

10M

12M

Age(days)

10%OPC

15%OPC

20%OPC

242.5

620

897.5

74.30

92.98

116.5

28

15.9

31.82

40.64

56

18.8

26.01

45.45

Age(days)

10%OPC

15%OPC

20%OPC

235

410.90

619.09

71.36

127.02

161.31

28

16.19

22.78

28.93

56

17.15

19.17

27.81

Age(days)

10%OPC

15%OPC

20%OPC

252.8

398.79

506.34

56.79

70.97

113.08

28

10.18

17.96

25.10

56

11.28

16.35

23.50

The percentage increase in compressive strength development of FA based geopolymer mortar


samples (FA-GPM) are shown in table 4.1. The FA based GPM samples without OPC are found
to be very weak and failed to produce a reasonable strength even after three days of casting when
cured in ambient temperature condition (2530C). Fig 4.1 shows negligible 3 days strength of the
FA-GPM. When OPC is incorporated at 10%, 15% and 20% FA replacements keeping other
parameters constant, the strength increased significantly from the early age of 3 days. This trend
is true at given 3 molarities. A percentage increase of 897% is observed in 8M FA sample with
20% OPC replacement. This percentage increase in 3 days strength is reduced when molarity of
the activator was increased. This is due to the fact that activation of FA increases at higher
molarities which provides higher 3 days strength for higher molarities. From the Table 4.1, there
is a significant increase in the 7 days strength which is ~60 to 120% of 3 days strength which
increases with the increase of OPC replacement. The 7 days strength is about 50-70% of the 28
days strength. Rate of strength development is retarded after 28 days and the percentage increase
is lowest for 12M mortar. It is also observed from the table 4.1 that 8M mix keeps on gaining
strength at 56 days but 12M (higher alkaline) mix gained strength earlier than 56 days and the rate
44

of gain retards at 56 days. Fig 4.1-4.3 show the variation of compressive strengths of FA-GPM
mixes at the age of 3, 7 days and 28 days due to the variation of OPC content as well as increasing
molarity of the activator. The 12M specimen shows greatest compressive strength. Low early
strength of 8M0OPC is due to the fact that the geopolymer gels produced by the dissolution of low
calcium fly ash generally have negligible traces of calcium in the reaction products, which is
primarily sodium alumino-silicate hydrate (N-A-S-H) (Garcia et al, 2010). Inclusion of OPC
supplies additional calcium and thus contributes to formation of the binding product containing
calcium ion (C-A-S-H/C-S-H) depending upon the pH of the environment. The N-A-S-H gel
remains dominantly in this system but in the presence of Ca, it is only stable at low pH. At high
pH (12), the presence of Ca will degrade N-A-S-H in favour of C-A-S-H formation until the
exhaustion of free Ca2+ ions (Garcia-Lodeiro et al, 2011). Due to this increase in the reaction
products, the strength for higher molarities is higher.

4.2.2 COMPRESSIVE STRENGTH OF GGBS BASED GEOPOLYMER SAMPLES

60

50.38
49.25

48.21 47.71

50

42.23

43.25
40.12

Strength (MPa)

40

38.52
32.17 31.72
30.12
29.4

30

20

20.23
18.1119.64
14.4
0%OPC

10

10%OPC

15%OPC

20%OPC

0
0

10

20

30

40

Curing days

Figure 4.4: Compressive strength of 8M GGBS samples


45

50

60

60

55.08
53.78
51.77

49.22

50

44.43
40

Strength (MPa)

46.35
44.35

41.2
40.21

36.55
36.09
31.22

30

23.41
20.3421.29
20
17.19
10

0%OPC

10%OPC

15%OPC

20%OPC

10

20

30

40

50

60

Curing days

Figure 4.5: Compressive strength of 10M GGBS samples


70

62.07
60

58.77

56.32 55.78

56.35

51.89

50

49.35

Strength (MPa)

46.47
41.19
40

43.71

38.23

30.69

30 26.68

24.1726.33
20.12

20

10

0%OPC

10%OPC

15%OPC

20%OPC

0
0

10

20

30

40

50

Curing days

Figure 4.6: Compressive strength of 12M GGBS samples


46

60

Table 4.2: Percentage increase in Compressive strength of GGBS based geopolymer

8M

10M

12M

Age(days)

10%OPC

15%OPC

20%OPC

25.76

36.3

40.48

7.89

2.44

9.42

28

9.6

23.85

25.1

56

7.80

22.75

25.57

Age(days)

10%OPC

15%OPC

20%OPC

18.32

23.85

36.18

15.59

17.07

31.96

28

10.49

22.40

33.74

56

4.50

16.73

24.19

Age(days)

10%OPC

15%OPC

20%OPC

20.12

30.81

32.60

24.56

34.21

51.4

28

18.71

27.61

28.84

56

14.18

19.08

25.77

The results of compressive strengths of geopolymer pastes are shown from Fig 4.4-4.6. The
compressive strength of pastes increased with increasing alkalinity of the activator. This is similar
for both FA and GGBS based geopolymers. The 28-days compressive strengths at for 8M, 10M
and 12M is 50.38, 55.08, and 62.7 MPa, respectively for 20% OPC replacement. It is observed
from the compressive strength of FA and GGBS based GPM is that the GGBS based geopolymers
shows higher strengths for every molarity and OPC replacement. From table 4.2, it is evident that
the gain of strength was similar for 8M, 10M and 12M molarities. Although the gain of strength
at 7 days accelerated for 12M activator. The readily available free calcium ions react with silica
and alumina and forms C(A)SH gel which coexist with the geopolymer gels. The addition of
calcium oxide (CaO) forms hydrated products such as calcium silicate hydrates (CSH) along with
the alumino-silicate geopolymer network. The amount of CaO content of the precursor materials
have considerable effect on the resulting hardened geopolymer.. The main reaction products
formed as a result of alkali activation of GBFS are CSH and/or CASH gels similar to those of PC
(Ismail et al, 2014), whereas the main product of alkali activation of FA is NASH gel as GGBS is
47

a glassy phase material, it is therefore easier to activate than FA. In addition, the reaction of GGBS
and alkali solutions is an exothermal process and generate heat which promote the
geopolymerization process. Also, the pozzolanic reactions are exothermic which further improves
accelerate the geopolymerisation process. The final strength obtained was 60.07MPa at 56 days
for 12M20OPC specimen.

4.2.3 COMPRESSIVE STRENGTH OF REPAIR MORTAR SAMPLES


Two available commercial repair material products (RM) were also tested. This was done so that
the performance of GPM could be compared with those of commercial products. RM1 (Sika TOP
112 HS) is general purpose non-shrink grout mortar; RM2 (BASF MasterFlow) is fiberreinforced non-shrink mortar. The amount of water in these products directly influences the
compressive and flexural strengths including the bond adhesion; therefore, the recommended
water/binder (W/B) ratios were strictly used. These repair mortars achieve 70% of their total
compressive strength within 24 to 48 hours (Fig 4.7). The obtained compressive strengths of these
repair materials are summarized in Table 4.3.

60

Compressive strength (MPa)

50
40
30

RM1(Sika)
RM2(BASF)

20
10
0
0

10

20

30

40

50

60

Age (days)

Figure 4.7: 28 days compressive strength of GGBS Repair mortars

48

Table 4.3: Water to binder ratio of repair mortars used in this study

Mortar

Water to binder
ratio

28 Days
Compressive
Strength (MPa)

56 Days
Compressive
Strength (MPa)

RM1

0.16

42.8

44.57

RM2

0.14

53.16

55.23

Repair

4.3 SHEAR BOND STRENGTH OF GPM AND RM

In this section, the bond strength between PCC substrate and proposed repair materials have been
analysed.

Table 4.4 : Percentage increase 28 days Shear bond strength of FA based geopolymer

Morality

10

12

% Replacement of FA with
OPC

% Increase

10

23.35

15

39.7

20

49.82

10

10.91

15

22.06

20

35.29

10

11.97

15

22.48

20

17.83

49

25

23.42

Shear Bond Strength (MPa)

21.41

15

17.31

19.74

19.2

19.12

20

21.13

23.42
21.17

22.53

17.43

14.13

10

0
0

10
15
% replacement by OPC
8M

10 M

20

12 M

Figure 4.8 : 28 days Slant shear strength of GGBS based geopolymer

Table 4.5: Percentage increase 28 days Shear bond strength of GGBS based geopolymer

Morality

10

12

% Replacement of GGBS
with OPC

% Increase

10

23.91

15

30.87

20

34.47

10

9.69

15

20.48

20

33.72

10

10.88

15

20.91

20

27.46

50

30

27.41
24.72

25

20

31.51

29.89

21.35

21.14

28.27

25.47

23.19

23.17

22.55

17.23

15
10
5
0
0

10

15

20

% replacement by OPC
8M

10 M

12 M

Figure 4.9: 28 days Slant shear strength of GGBS based geopolymer

Table 4.6: 28 days Shear bond strength of Repair Mortars

Repair Mortar

SHEAR BOND
STRENGTH (MPa)

23.43

29.42

35

Shear Bond Strength

Shear Bond Strength (MPa)

35

30
25

29.42
24.43

20
15
10
5
0
Repair Mortar
Sika

BASF

Figure 4.10: 28 days Slant shear strength of Repair Mortars


51

The results of 45 slant shear load carrying capacity of PCC substrate and GPM and RM are shown
from Fig.4.8 to 4.10. The shear bond strengths increased with the increased OPC content and
NaOH concentration in the alkaline activator (Table 4.4,4.5). The increase in shear bond strength
is directly related to the increase in compressive strength of geopolymer paste. The percentage
increase in the shear bond strength is in agreement with the increase in the compressive strengths
obtained earlier with varying molarities and replacements. The percentage increase in shear bond
strength of the FA mix reduces as the molarity as well as the OPC replacement increases. In case
of GGBS based geopolymer, the change in the percentage increase is less than 10% for every
molarity and replacement. This is due to the fact that the activation of the material increases with
increase in the molarity of the activator but GGBS gets easily activated than FA. The specimens
of G12M20OPC mix gave the maximum slant shear strength. The noticeable increase in shear
bond strength was due to the increase in the reaction products. This is in line with the previous
results (Dombrowski et al. 2007) on the improved strength of fly ash based geopolymers with
increased calcium content which the additional CSH and CASH gel co-existed with NA
SH gel of GPM. The SEM images (Fig 4.16) discussed later in this chapter shows the increase in
reaction products at the interface transition zone between PCC substrate and GPM enhanced the
strength at contact zone. The GGBS samples showed greater slant shear strength than FA-GPM
for every molarity of the alkaline activator. This is due to the fact that the GGBS contain high
amount of CaO. The high CaO content from GGBS promotes the formation of additional CSH
and/or CASH gel, therefore, the bonding at the interface transition zone between old concrete and
geopolymer is enhanced (Phoo-ngernkham et al,2015). Addition of OPC shows increase in
strength but did not show much increase after further replacement with GGBS mortars as the
mortar paste already contains large percentage of CaO in it. Also, the reason for greater bond
strength is that geopolymer is rich in Si4+ and Al3+ ions, therefore, it can react with Ca(OH)2 at the
surface of PCC substrate leading to increased strength development at the contact zone. From to
Fig 4.10, the results of 45 slant shear load carrying capacity of PCC substrate and RM were 23.43
MPa (RM1) and 29.42 MPa (RM2). The average strength being 26.42 MPa of the two repair
mortars. The mixes with 10M and 12M of activator exhibited comparable and in case of samples
GGBS based geopolymer, 3 samples showed results higher than the average strength of the repair
mortars. This indicated that the GPM containing OPC as additive can be used as an alternative
repair material. Most failure modes for the slant shear test with line interface at 45 follow the
52

interface surface with small particle disintegration as shown in Fig 4.11 due to decreased
compressive stress along the interface. Different failure patterns of the slant shear prisms are
shown in Fig 4.11. Two failure patterns could be identified. The first was the failure in the GPM
from which the cracks were formed in the GPM and the interface while the PCC substrate remained
relatively intact. This occurred with the FA and GGBS based geopolymer samples having low
alkaline activator concentration and without OPC and low OPC content (F8M0OPC and
F8M10OPC mixes) geopolymer mortars. For other mixes with relatively high strengths, high
NaOH content i.e. high alkaline activator and high OPC such as G10M10OPC and F12M10OPC
mixes, the slant shear bond prisms failed in the monolithic mode where cracks were formed in
both sections of GPM and PCC substrate. It was also observed that for every GGBS based GPM
specimens the failure mode was monolithic. This indicated the relatively high resistance to
cracking of GPM and the high bonding between the two surfaces. Thus cracks went through the
slant plane resulted in monolithic type of failure as in Fig 4.11(b), (e).For the prisms with RM2
the failures occurred in the monolithic mode as shown in Fig 4.11(f). This again indicated the
relatively high resistance to cracking of RM2 and the high bonding between the two surfaces. All
the mixes having high alkalinity provided comparable results with average slant shear strength of
the two repair mortars. The GGBS mixes provided strength much better than the average strength
of the RM, G12M20OPC provided strength 31.51MPa which was higher than the two repair
mortars.

GPM

PCC
(a)F10M10OPC

GPM

GPM

PCC

PCC

(b)F12M20OPC
53

(c) G8M10OPC

GPM

GPM

PCC

RM2

PCC

PCC

(d) G10M10OPC

(e) G12M20OP

(f) RM2

Figure 4.11: Different Patterns of failed Slant Shear Specimens

4.4 BENDING STRESS OF PCC NOTCHED BEAM WITH FILLED GPM AND RM
This test used the same type of specimens as in the test of fracture characteristics of the geopolymer
mortar and repair materials.
3.5
2.96

Bending Stress ( MPa )

2.64

2.53
2.5

2
2

2.1

2.26

2.73
2.46

2.24

1.93

1.57
1.5
1
0.5
0
0

10

15
% Replacement by OPC

8M

10 M

12 M

Figure 4.12: 28 days bending strength of FA geopolymer mortar


54

20

3.12

4.5

Bending Stress ( MPa )

4
3.35

3.5

3.01

3
2.5

3.16

3.07

2.83

3.76

3.73

3.62

3.94

3.34

3.25

2.47

2
1.5
1
0.5
0
0

10

15

20

% Replacement by OPC
8M

10 M

12 M

Figure 4.13: 28 days bending strength GGBS geopolymer mortars

Table 4.7 : Percentage change in 28 days Bending strength of FA geopolymer mortar


Morality

10

12

% Replacement of FA with
OPC

% Change in Bending
Strength (MPa)

10

22.92

15

42.67

20

56.68

10

13

15

32

20

48

10

20.47

15

30

20

48.57

55

Table 4.8: Percentage change in 28 days Bending strength of GGBS geopolymer mortar

Morality

10

12

% Replacement of GGBS
with OPC

% Change in Bending
Strength (MPa)

10

14.57

15

24.29

20

27.12

10

4.98

15

7.97

20

24.91

10

8.05

15

11.34

20

17.61

3.7

3.5

Flexural Strength ( MPa)

3.04
3
2.5
Sika

BASF

1.5
1
0.5
0
Repair Mortar

Figure 4.14: 28 days Bending strength of Repair mortars

56

Table 4.9: 28 days Bending strength of Repair mortars

Repair Mortar

Flexural Strength
(MPa)

3.04

3.7

The bending strength of PCC notched beams filled with GPM and RM are shown from Fig 4.12
to 4.14. The percentage change in the bending strength due to change in molarity and OPC content
are presented in table 4.7 and 4.8. The percentage increase is observed to be more in case of FA
geopolymer mixes as compared to GGBS based mixes. The maximum percentage increase was
observed in F8M20OPC mix. The percentage increase is less at higher molarity activator in FA
mixes, as activation accelerates at higher molarities. The repair mortars bending stresses are 3.04
and 3.7 MPa for RM1 and RM2 (Table 4.9) respectively giving an average value of 3.37 MPa.
These values are much more than the value for PCC beam without any repair material in the notch
(base line strength 1.8MPa). For the FA-GPM, the bending stresses is lowest for F8M0OPC sample
and the value obtained is 1.57 MPa which is lower than the base line beam. With the exception of
mixes with low alkaline activator concentration (8M) and OPC content of 0% and 10%, all other
mixes produce sufficiently high bending stresses especially mixes with high NaOH concentrations
and high OPC contents. As explained previously, this increase in bending stress is due to the
increase in reaction products with associated improvements of strength and bonding capacity of
GPM which leads to overall improvement in the bending stresses of PCC notched beams. From
fig.4.12, the notched beam filled with GPM at 20% OPC and 12 M NaOH gave excellent bending
stress of 3.12 MPa representing around 85% improvement from the base line. This is close to the
average value of the bending strength provided by the repair mortars. The good performance with
high NaOH concentration mixes can be attributed to the interaction between the NaOH and PCC
substrate at the transition zone. On the other hand, the GGBS based geopolymer mortars gave high
value of bending strength as seen earlier in slant shear test. GGBS geopolymer sample
(G12M20OPC) made from 12M alkaline activator provided excellent bending strength of 3.94
MPa and was much more than average repair mortar bending strength. The reason for increase of
bending strength as explained earlier. The lowest bending strength was given by F8M0OPC mix,
57

but results of bending stress test obtained from higher molarity mixes of FA and GGBS mortar
samples confirmed the suitability of GPM containing PC as additive as an alternative repair
material, GGBS mortar being more archaic. Fig 4.16 shows that the beam and geopolymer mortar
fails monolithically. Geopolymer mortar like commercial repair mortar did not failed earlier or fell
out of the notch on application of loading

(a) F10M15OPC sample

(b) G10M15OPC sample


Figure 4.15 (a) and (b): Typical monolithic failure pattern of notched beams filled with
geopolymer mortars

4.5 BOND STRENGTH BY CYLINDER SPLITTING TEST


The purpose of this test in this study is to check the bond quality of the specimens with maximum
bond strength and compare it with the standard results provided by Ozyildirim et al. (2000) as
shown in Table 4.10. Specimens from mix combination showing improved compressive and slant
shear strength. Mixes in table 4.11 F12M15OPC, F12M20OPC and G12M15OPC and
G12M20OPC were tested to determine the bond strength quality.
58

Table 4.10 :Bond strength quality (Ozyildirim et al. 2000)

Bond Strength(MPa)

Quality

2.1

Excellent

1.7 to 2.1

Very good

1.4 to 1.7

Good

0.7 to 1.4

Fair

0 to 0.7

Poor

Table 4.1 :Bond strength by cylinder split tensile strength

Specimen

Bond Strength (MPa)

Quality

F12M15OPC
F12M20OPC

1.13
1.42

Fair
Good

G12M15OPC
G12M20OPC

2.88
3.14

Excellent
Excellent

RM1 (Sika)
RM2 (BASF)

2.15
3.47

Excellent
Excellent

The average bond strength of the two repair mortars obtained is 2.81 MPa which is categorized as
an excellent bond strength. GGBS based geopolymers gave bond strength more than the average
of the repair mortars and were also excellent. The reason for good bond properties of GGBS has
been explained earlier. The bond strength by FA geopolymer range from fair to good which is
satisfactory.

59

4.6 SEM OF INTERFACE ZONE BETWEEN PCC SUBSTRATE AND GPM

Images for microstructural analysis were acquired with either a JOEL JSM-6510LV scanning
electron microscope equipped with an Oxford Instruments INCAx-act energy dispersive X-ray
spectroscopy (EDS) system. The EDS systems allow the collection of chemical information for
spots and areas in the samples. The samples comprised broken pieces (~5.0x5.0x2.5 mm) of
hardened paste mounted onto aluminum stubs using epoxy adhesive, orienting the fractured
surface in a convenient position towards the SEM beam also bearing in mind the location of the
EDS detector. Samples were coated with a thin goldpalladium layer to provide a conductive
pathway to prevent surface electrical charging. The SEM and XRD analysis were performed in
SAI lab, Thapar University.

(a) G12M20OPC

(b) G10M15OPC

(c) G8M0OPC

(d) F12M20OPC
60

(e) F10M15OPC

(f)F8M0OPC

Figure 4.16: SEM of interface zone between PCC substrate and GPM

The SEM results of some of the fracture interfaces between PCC substrate and GPM are shown in
Fig.4.16. The fly ash and GGBS geopolymer mixes having low molarities [Fig 4.16 (c),(f)]
showed loose matrix in comparison with mortars made from activator of higher molarity. The
GPM with higher OPC and higher alkaline activator solution (G12M20OPC,F12M20OPC)
appears denser than the low OPC and less alkaline activator (8M0OPC) based GPM. This is in line
with the previous research which indicated that the use of activator solution made from NaOH and
Sodium Silicate could accelerate the geopolymerization process. The GGBS mortar matrix
appeared denser than those of FA based GPM. Also, the reaction of GGBS with alkali solution is
an exothermal process and liberated heat and thus formation of additional CSH and CASH led to
the overall strength development. In general, the fracture surface of mortar with low NaOH
concentration and without OPC (F8M0OPC) as shown in Fig.4.16f showed the relatively plane
fracture surface between the PCC substrate and GPM indicating the clean separation of the two
surfaces and the low bonding. This result corresponded to a low bending stress and low shear bond
strength of the mix with low NaOH and without OPC mortar (F8M0OPC). The fracture interfacial
zone of the mix with high NaOH and high OPC mortar (G12M20OPC and F12M20OPC) as shown
in Fig.4.16 (a),(d) indicated that the bonding surface was still intact. The cracks passed through
the PCC substrate and GPM interface, and there was no significant gap between the two bonding
surfaces. For the F12M20OPC mix as shown in Fig.4.16a,d, the SEM showed a highly irregular
or rugged fracture surface with no visible plane fracture surface indicating the good bonding
61

between the two surfaces. The results confirmed the increase in shear bond strength and bending
stress with more alkaline activator and high OPC mortar and the associated monolithic failure
mode as shown in Fig.4.11(b) and (e). The increase in bonding has been reported by PachecoTorgal et al. (2008). Also, Shi et al. (2012) report indicated the better bonding of alkali activated
binder than that of conventional Portland cement. Furthermore, it has been reported that when
recycled concrete is used, the cementing property is activated both by the alkali activation and by
the calcium hydroxide presented in the residual paste (Achtemichuk et al, 2009) Generally,
geopolymer is rich in Si4+ and Al3+ ions, therefore, it can react with Ca(OH)2 at the surface of PCC
substrate leading to increased strength development at the contact zone. Moreover, the increase in
Ca2+ ion balanced the negative charge of [Al(OH)4]- ions, which resulted in the increase in reaction
products at the interface transition zone between PCC substrate and geopolymer matrix leading to
a dense interface zone and high strength geopolymer

4.7 XRD ANALYSIS


X-ray diffraction on raw materials and reacted pastes was carried out with Panalytical XPert Pro
MPD Diffractometer with XCelerator detector scanned the range of 1060 2, using Cu K
radiation. Zincite was used as internal standard. HighScore Plus software was used to obtain semiquantitative mineral weight percentages. the XRD patterns of the geopolymer pastes at 28 days of
age. For all specimens, diffraction of X-rays apparently resulted in a broad diffuse halo in the value
of 2 ranging from 20 to 60 (Fig 1.7a).This indicates the presence of large quantity of amorphous
or non -crystalline gels in the FA geopolymer peaks due to the crystalline components of Quartz,
Hydrosodalite, Mullite and Hematite from the fly ash were evident in all the pastes. Mixtures
having OPC revealed peaks of Alite (tri-calcium silicate) and CSH (calcium silicate hydrate) in
addition to other common fly ash geopolymer phases. The identification of these compounds is
done on the bases of peaks formed in this study and its comparison with typical pattern of peaks
obtained by different phases at different degrees on X-axis. The peaks obtained were conforming
to previous studies on this material done by Nath et al. (2015).

62

(a) 1: F8M15OPC 2:F10M15OPC 3:F12M15OPC

(b) (a) 1: G8M15OPC 2:G10M15OPC 3:G12M15OPC


Figure 4.17: XRD diffractograms of (a)FA based geopolymer; (b)GGBS based Geopolymer
63

For the GGBS mortar, the XRD pattern shows a large amount of amorphous phase with additional
CSH. With GGBS, the amorphous phases were easily detected as broad hump around 2835 2
due to the formation of amorphous components in geopolymer gel. Calcium element also reacted
to form CSH which coexisted with geopolymer gel (Ismail et al. 2014) and high 28-days
compressive strength was obtained. The presence of CSH and relatively high compressive strength
is in line with the previous research (Kumar et .al 2010). The incorporation of OPC provided
additional calcium to react and form CSH in the geopolymer system.

64

CHAPTER 5
CONCLUSIONS
In this study, the mechanical properties of FA and GGBS based geopolymers were studied. These
properties includes compressive strength, slant shear strength, bending strength and cylinder split
tensile strength. Microstructural properties were also studied using SEM and XRD analysis. Based
on the test results, the following conclusions are drawn:

1. The compressive strength of FA and GGBS based geopolymer mortars were


comparable with the commercially available repair mortars.

2. The bond strength, bending strength and split tensile strength was found to be directly
proportional to the compressive strength of the mortar.

3. The compressive strength for FA and GGBS based geopolymer mortars is observed to
increase with increase in molarity of the activator and increase in the OPC content.
Among the FA based GPMs, F12M20OPC and for GGBS, mix G12M20OPC showed
mix showed the maximum compressive strength.

4. There was an increase of 5 times in 3 days compressive strength for 12M FA samples
incorporating 20% OPC replacement.

5. GGBS based mortar gave superior compressive strength than the fly ash based mortars
for every molarity. The highest compressive strength achieved was 62MPa for
G12M20OPC specimen.

6. The slant shear strength of GGBS geopolymer mortars is superior than FA based
geopolymers. The 12M GGBS mix gave strength greater than average slant shear
strength of repair mortar. The maximum slant shear strength of 31.51MPa was

65

obtained for GGBS sample G12M20OPC which was more than average slant shear
strength of the two repair mortars (26.42MPa).

7. Except for mix F8M0OPC, both geopolymers provided bending strength greater than
base line bending strength. While there was an average improvement of 87% by using
RM, the average improvement using 12M GGBS and FA mortars were 98% and 45%
respectively from the base line.

8. The cylinder split tensile test to check bond strength proved that GGBS mixes
G12M15OPC and G12M15OPC provided excellent bond strength quality whereas the
bond strength quality of fly ash mixes F12M15OPC and G12M20OPC varied from
fair to good with increasing molarity.

9. SEM images showed that the geopolymer mortar samples with 12M activator and high
percentage of OPC were denser as compared with rest of the samples. The specimen
G12M20OPC and F12M20OPC appeared most dense in SEM images among their
series respectively.

10. The results indicated that the performances of GPM containing OPC with high alkaline
concentrations as a repair material was at least equal to those of commercially
available ones according to the slant shear bond test and beam notch filled bending
test. The performance of GGBS based geopolymers was better than low calcium FA
base geopolymers.

66

FUTURE SCOPE OF WORK

The comparison between the suggested geopolymer mortars with other pavement repair materials
in the market shows the superior bond strength of the suggested fly ash as well as GGBS based
geopolymer to PCC substrate. The bond strength between the PCC and the geopolymer mortar
through the Slant shear test, Cylindrical split tensile test, Bending stress test were also conducted.
It is found that the failure started at the mortar substrate and most of failure modes occurred
through mortar substrate, which indicates an excellent bond interface formed by geopolymer
mortar with PCC substrate. However, this study has a limitation on the process of creating alkaline
sodium silicates solution. Alternate materials can be used to make Alkaline activator solutions and
its effect on geopolymerization process can be studied. This study also not includes the effects of
elevated temperatures on bond strength. Hence, further studies must be conducted to investigate
the effect of these variations on geopolymers. Experiments can be focused on higher curing days
(90 and 365 days).

67

REFERENCES
Alanazi, H., Yang, M., Zhang, D., & Gao, Z. (2016). Bond strength of PCC pavement repairs using
metakaolin-based geopolymer mortar. Cement and Concrete Composites, 65, 7582.
ASTM C496 / C496M - 11 Standard Test Method for Splitting Tensile Strength of Cylindrical
Concrete Specimens
ASTM C882 / C882M Standard Test Method for Bond Strength of Epoxy-Resin Systems Used
With Concrete By Slant Shear
ASTM C882, Standard Test Method for Bond Strength of Epoxy-Resin Systems used With
Concrete by Slant Shear, vol. 04.02, Annual Book of ASTM Standard,2005.
B. Bharatkumar, B. Raghuprasad, D. Ramachandramurthy, R. Narayanan, S. Gopalakrishnan,
Effect of fly ash and slag on the fracture characteristics of high performance concrete, Mater.
Struct. 38 (1) (2005) 6372.
Bakharev, T. (2005a). Durability of geopolymer materials in sodium and magnesium sulfate
solutions. Cement And Concrete Research, 35(6), 1233-1246.
Bakharev, T. (2005b). Geopolymeric materials prepared using Class F fly ash and elevated
temperature curing. Cement And Concrete Research, 35(6), 1224-1232.
Bakharev, T. (2005c). Resistance of geopolymer materials to acid attack. Cement And Concrete
Research, 35(4), 658-670.
Cement.

In

Wikipedia.

Retrieved

May

16,

2016,

from

https://en.wikipedia.org/wiki/List_of_countries_by_cement_production
chemical composition, temperature and aggregate:binder ratio
Cheng, T. W., & Chiu, J. P. (2003). Fire-resistant geopolymer produced by granulated blast
furnace slag. Minerals Engineering, 16(3), 205-210.
Criado, M., Criado, M., Fernandez-Jimenez, a, Fernandez-Jimenez, a, de la Torre, a G., de la
Torre, a G., Palomo, a. (2007). An XRD study of the effect of the SiO2/Na2O ratio on the
alkali activation of fly ash.

68

D.G. Geissert, S.E. Li, G.C. Frantz, E.J. Stephens, Splitting prism test method to evaluate concreteto-concrete bond strength, ACI Mater. J. 96 (3) (1999).
D.G. Geissert, S.E. Li, G.C. Frantz, E.J. Stephens, Splitting prism test method to evaluate concreteto-concrete bond strength, ACI Mater. J. 96 (3) (1999) 359e366.
Davidovits, J. (1984). Synthetic Mineral Polymer Compound of The Silicoaluminates Family and
Preparation Process, United States Patent - 4,472,199 (pp. 1-12). USA.
Davidovits, J. (1988a). Soft Mineralurgy and Geopolymers. Paper presented at the Geopolymer
88, First European Conference on Soft Mineralurgy, Compiegne, France.
Davidovits, J. (1988a). Soft Mineralurgy and Geopolymers. Paper presented at the Geopolymer
88, First European Conference on Soft Mineralurgy, Compiegne, France.
Davidovits, J. (1988b). Geopolymer Chemistry and Properties. Paper presented at the Geopolymer
88, First European Conference on Soft Mineralurgy, Compiegne, France.ent And Concrete
Research, 29(8), 1323-1329.
Davidovits, J. (1988b). Geopolymer Chemistry and Properties. Paper presented at the Geopolymer
88, First European Conference on Soft Mineralurgy, Compiegne, France.
Davidovits, J. (1988c). Geopolymers of the First Generation: SILIFACE-Process. Paper presented
at the Geopolymer 88, First European Conference on Soft Mineralurgy, Compiegne, France.
Davidovits, J. (1988d). Geopolymeric Reactions in Archaeological Cements and in Modern
Blended Cements. Paper presented at the Geopolymer 88, First European Conference on Soft
Mineralurgy, Compiegne, France.
Davidovits, J. (1991). Geopolymers: Inorganic Polymeric New Materials. Journal of Thermal
Analysis, 37, 1633-1656.
Davidovits, J. (1991). Geopolymers: Inorganic Polymeric New Materials. Journal of Thermal
Analysis, 37, 1633-1656.
Davidovits, J. (1994a). High-Alkali Cements for 21st Century Concretes. Paper presented at the
V. Mohan Malhotra Symposium on Concrete Technology: Past, Present And Future, University
of California, Berkeley.
69

Davidovits, J. (1994a). High-Alkali Cements for 21st Century Concretes. Paper presented at the
V. Mohan Malhotra Symposium on Concrete Technology: Past, Present And Future, University
of California, Berkeley.
Davidovits, J. (1994b). Properties of Geopolymer Cements. In Kiev (Ed.), First International
Conference on Alkaline Cements and Concretes (pp. 131-149).Kiev, Ukraine: Kiev State
Technical University.
Davidovits, J. (1994c). Global Warming Impact on the Cement and Aggregates Industries. World
Resource Review, 6(2), 263-278.
Davidovits, J. (1999, 30 June - 2 July 1999). Chemistry of Geopolymeric Systems, Terminology.
Paper presented at the Geopolymere 99 International Conference, Saint-Quentin, France.
Davidovits, J. (1999, 30 June - 2 July 1999). Chemistry of Geopolymeric Systems, Terminology.
Paper presented at the Geopolymere 99 International Conference, Saint-Quentin, France.
Davidovits, J. (2005). Green-Chemistry and Sustainable Development Granted and False Ideas
About Geopolymer-Concrete. Paper presented at the International Workshop on Geopolymers and
Geopolymer Concrete (GGC), Perth, Australia.
De Vargas, A. S., Dal Molin, D. C. C., Vilela, A. C. F., Silva, F. J. D., Pavo, B., & Veit, H. (2011).
The effects of Na2O/SiO2 molar ratio, curing temperature and age on compressive strength,
morphology and microstructure of alkali-activated fly ash-based geopolymers. Cement and
Concrete Composites, 33(6), 653660.
Diaz EI, Allouche EN, Eklund S. Factors affecting the suitability of fly ash as source material for
geopolymers. Fuel 2010;89(5):9926.
Duxson, P., Mallicoat, S. W., Lukey, G. C., Kriven, W. M., & van Deventer, J. S. J. (2007). The
effect of alkali and Si/Al ratio on the development of mechanical properties of metakaolin-based
geopolymers. Colloids and Surfaces A: Physicochemical and Engineering Aspects,
F. Pacheco-Torgal, J.A. Labrincha, C. Leonelli, A. Palomo, P. Chindaprasirt, Handbook of AlkaliActivated Cements Mortars and Concretes, Wood Head Publishing Limited Elsevier Science and
Technology, 2014.

70

FM 3-C 882, Florida Test Method for Performance of epoxy-resin systems with concrete by slant
shear and compressive strength, In Florida Department of Transportation Standard Specifications
for Road and Bridge Construction, 2015, July.
Garca-Lodeiro I, Fernndez-Jimnez A, Blanco MT, Palomo A. FTIR study of the solgel
synthesis of cementitious gels: CSH and NASH. J SolGel Sci Technol 2008;45(1):6372.
Garcia-Lodeiro I, Palomo A, Fernandez-Jimenez A, Macphee DE. Compatibility studies between
N-A-S-H and C-A-S-H gels. Study in the ternary diagram Na2OCaOAl2O3SiO2H2O. Cem
Concr Res 2011;41(9):92331.
Grhan, G., & Krkl, G. (2014). The influence of the NaOH solution on the properties of the fly
ash-based geopolymer mortar cured at different temperatures. Composites Part B: Engineering,
58, 371377.
Grhan, G., & Krkl, G. (2014). The influence of the NaOH solution on the properties of the fly
ash-based geopolymer mortar cured at different temperatures. Composites Part B
Gourley, J. T. (2003). Geopolymers; Opportunities for Environmentally Friendly Construction
Materials. Paper presented at the Materials 2003 Conference: Adaptive Materials for a Modern
Society, Sydney.
IS 14858:2000 Compression Testing Machine Used For Testing Of Concrete And Mortar Requ
IS 4031: Methods of physical tests for hydraulic cement
IS 8112:2013 Ordinary Portland Cement, 43 Grade Specification
IS: 12089-1987, Specification For Granulated Slag For The Manufacture of Portland Slag Cement.
IS: 3812-2003, Specification for Fly ash for Use as Pozzolana and Admixture, Bureau of Indian
Standards, New Delhi, India.
IS:2250-1981;Code Of Practice For Preparation And Use Of Masonry Mortars
IS:383-1970;Specification For Coarse And Fine Aggregates From Natural Sources For Concrete
IS:460-1962. Indian standard specification for test sieves

71

IS:460-1962. Indian standard specification for test sieves. - New Delhi Indian Standards Institution
1967 - 27p.
Ismail I, Bernal SA, Provis JL, San Nicolas R, Hamdan S, van Deventer JSJ. Modification of phase
evolution in alkali-activated blast furnace slag by the incorporation of fly ash. Cement Concr
Compos 2014;45:12535.
K. Dombrowski, A. Buchwald, M. Weil, The influence of calcium content on the structure and
thermal performance of fly ash based geopolymers, J. Mater. Sci. 42 (9) (2007) 30333043.
Krkl, G. (2016). The effect of high temperature on the design of blast furnace slag and coarse
fly ash-based geopolymer mortar. Composites Part B: Engineering, 92, 918.
Kumar S, Kumar R, Mehrotra SP. Influence of granulated blast furnace slag on the reaction,
structure and properties of fly ash based geopolymer. J Mater Sci 2010;45(3):60715.
Lecomte I, Henrist H, Liegeois M, Maseri F, Rulmont A, Cloots R. (Micro)-structural comparison
between geopolymers, alkali-activated slag cement and Portland cement. J Eur Ceram Soc
2006;26:378997.
Malhotra, V. M. (1999). Making Concrete "Greener" With Fly Ash. ACI Concrete International,
21(5), 61-66.
Malhotra, V. M., & Mehta, P. K. (2002). High-Performance, High-Volume Fly Ash Concrete:
Materials, Mixture Proportioning, Properties, Construction Practice, and Case Histories. Ottawa:
Supplementary Cementing Materials for Sustainable Development Inc.
McCaffrey, R. (2002). Climate Change and the Cement Industry. Global Cement and Lime
Magazine (Environmental Special Issue), 15-19.
Mijarsh, M. J. A., Megat Johari, M. A., & Ahmad, Z. A. (2015b). Compressive strength of treated
palm oil fuel ash based geopolymer mortar containing calcium hydroxide, aluminum hydroxide
and silica fume as mineral additives. Cement and Concrete Composites, 60, 6581.
Mijarsh, M. J. A., Megat Johari, M. A., & Ahmad, Z. A. (2015c). Effect of delay time and Na2SiO3
concentrations on compressive strength development of geopolymer mortar synthesized from
TPOFA. Construction and Building Materials, 86, 6474.
72

Palomo, A., M.W.Grutzeck, & M.T.Blanco. (1999). Alkali-activated fly ashes A cement for the
future.
Phoo-ngernkham, T., Maegawa, A., Mishima, N., Hatanaka, S., & Chindaprasirt, P. (2015). Effects
of sodium hydroxide and sodium silicate solutions on compressive and shear bond strengths of FA
GBFS geopolymer. Construction and Building Materials, 91, 18.
Phoo-ngernkham, T., Sata, V., Hanjitsuwan, S., & Ridtirud, C. (2015). Journal of Materials in
Civil Engineering High calcium fly ash geopolymer mortar containing Portland cement for use as
repair material High calcium fly ash geopolymer mortar containing Portland cement for use as
repair material, 98, 482488 properties. Research report GC 2, Perth, Australia: Faculty of
Engineering,
Pradip Nath , Prabir Kumar Sarker ,(2014) Effect of GGBFS on setting, workability and early
strength properties of fly ash geopolymer concrete cured in ambient condition
Puertas F, Martnez-Ramrez S, Alonso S, Vazquez T. Alkali-activated fly ash/slag cements:
strength behaviour and hydration products. Cem Concr Res 2000;30(10):162532.
Rangan, B. V., Hardjito, D., Wallah, S. E., & Sumajouw, D. M. J. (2005a). Fly ashbased
geopolymer concrete: a construction material for sustainable development. Concrete in Australia,
31, 25-30.
Rangan, B. V., Hardjito, D., Wallah, S. E., & Sumajouw, D. M. J. (2005b). Studies of fly ash-based
geopolymer concrete. Paper presented at the World Congress Geopolymer 2005, Saint-Quentin,
France.
Ral Arellano-Aguilar a, Oswaldo Burciaga-Daz a, Alexander Gorokhovsky b, Jos Ivn
Escalante-Garca , (2014), Geopolymer mortars based on a low grade metakaolin: Effects of the
Roy, D. M. (1999). Alkali-activated cements Opportunities and Challenges. Cement & Concrete
Research, 29(2), 249-254.
S. Achtemichuk, J. Hubbard, R. Sluce, M.H. Shehata, The utilization of recycled concrete
aggregate to produce controlled low-strength materials without using Portland cement, Cement
Concr. Compos. 31 (8) (2009) 564569.

73

Sarker PK. Analysis of geopolymer concrete columns. Mater Struct 2009;42(6):71524.


Somna K, Jaturapitakkul C, Kajitvichyanukul P, Chindaprasirt P. NaOHactivated ground fly ash
geopolymer cured at ambient temperature. Fuel 2011;90(6):211824.
Soutsos, M., Boyle, A. P., Vinai, R., Hadjierakleous, A., & Barnett, S. J. (2016b). Factors
influencing the compressive strength of fly ash based geopolymers. Construction and Building
Materials, 110, 355368.
Soutsos, M., Boyle, A. P., Vinai, R., Hadjierakleous, A., & Barnett, S. J. (2016b). Factors
influencing the compressive strength of fly ash based geopolymers. Construction and Building
Materials, 110, 355368.
Swanepoel, J. C., & Strydom, C. A. (2002). Utilisation of fly ash in a geopolymeric material.
Applied Geochemistry, 17(8), 1143-1148.
T. Phoo-ngernkham, A. Maegawa, N. Mishima, S. Hatanaka, P. Chindaprasirt, Effects of sodium
hydroxide and sodium silicate solutions on compressive and shear bond strengths of FAGBFS
geopolymer, Constr. Build. Mater. 91 (2015) 18.
Tchakoute, H. K., Roscher, C. H., Kong, S., Kamseu, E., & Leonelli, C. (2016). Geopolymer
binders from metakaolin using sodium waterglass from waste glass and rice husk ash as alternative
activators: A comparative study. Construction and Building Materials, 114, 276289.
Temuujin, J., Van Riessen, A., & MacKenzie, K. J. D. (2010a). Preparation and characterisation
of fly ash based geopolymer mortars. Construction and Building Materials, 24(10), 19061910.
Thokchom, S., Dutta, D., & Ghosh, S. (2011). Effect of Incorporating Silica Fume in Fly Ash.
World Academy of Science, Engineering and Technology, 60(12), 243247.
van Jaarsveld, J. G. S., van Deventer, J. S. J., & Lukey, G. C. (2002a). The effect of composition
and temperature on the properties of fly ash- and kaolinite-based geopolymers. Chemical
Engineering Journal, 89(1-3), 63-73.
van Jaarsveld, J. G. S., van Deventer, J. S. J., & Lukey, G. C. (2002a). The effect of composition
and temperature on the properties of fly ash- and kaolinite-based geopolymers. Chemical
Engineering Journal, 89(1-3), 63-73.
74

van Jaarsveld, J. G. S., van Deventer, J. S. J., & Lukey, G. C. (2002a). The effect of composition
and temperature on the properties of fly ash- and kaolinite-based geopolymers. Chemical
Engineering Journal, 89(1-3), 63-73.
van Jaarsveld, J. G. S., van Deventer, J. S. J., & Lukey, G. C. (2002b). The characterisation of
source materials in fly ash-based geopolymers. Materials Letters, 3975(Article in press).
van Jaarsveld, J. G. S., van Deventer, J. S. J., & Schwartzman, A. (1999). The potential use of
geopolymeric materials to immobilise toxic metals: Part II. Material and leaching characteristics.
Minerals Engineering, 12(1), 75-91..
Wallah SE, Rangan BV. Low-calcium fly ash-based geopolymer concrete: longterm
X.S. Shi, F.G. Collins, X.L. Zhao, Q.Y. Wang, Mechanical properties and microstructure analysis
of fly ash geopolymeric recycled concrete, J. Hazard Mater. 237238 (2012) 2029.
Xu, H., & Deventer, J. S. J. V. (2000). The geopolymerisation of alumino-silicate minerals.
International Journal of Mineral Processing, 59(3), 247-266.
Zhou, W., Yan, C., Duan, P., Liu, Y., Zhang, Z., Qiu, X., & Li, D. (2016). A comparative study of
high- and low-Al2O3 fly ash based-geopolymers: The role of mix proportion factors and curing
temperature. Materials and Design, 95, 6374.

75

S-ar putea să vă placă și