You are on page 1of 13

Mechanics of Materials 103 (2016) 110–122

Contents lists available at ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mechmat

“Fly ash and GGBFS based powder-activated geopolymer binders: A


viable sustainable alternative of portland cement in concrete industry”
Kamal Neupane
Masters of Engineering by Research, B. E. Civil, University of Technology Sydney, NSW, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Geopolymer is a new binding material, synthesized by alkali activation of aluminosilicate compounds.
Received 15 December 2015 Previous researches around the world postulated that concrete from geopolymer binder exhibited supe-
Revised 5 September 2016
rior engineering, thermal and durability properties than ordinary Portland cement (OPC) concrete, such as
Available online 22 September 2016
higher mechanical strengths, higher resistivity to sulphate and acid attacks and higher thermal resistivity.
Keywords: Geopolymers are generally made from activation of aluminosilicate powders by highly concentrated
Powder-activated sodium hydroxide and or sodium silicate solutions, known as liquid-activated geopolymer. Recently, some
Liquid-activated cement and concrete companies around the world have commenced the production of liquid activator
Geopolymer binder based geopolymer binders. However, this type of geopolymer does not appear as a viable replacement of
Compressive strength Portland cement in concrete industry despite its substantial environmental benefits because of its lim-
Indirect tensile strength itation in mixing and handling process. Powder-activated geopolymer contains alkali activators in pow-
Flexural strength
der form which are blended together with source material. This geopolymer binder is physically similar
Modulus of elasticity and drying shrinkage
to OPC and mixing and handling process are also similar to conventional process. Experimental results
showed that concrete from this binder can set and harden in ambient temperature with significant early
age strength.
Two types of powder-activated geopolymer binders having different proportions of fly ash and slag
were used in this study. Engineering properties of geopolymer concrete of four different strength grades
(40, 50, 65 and 80 MPa) were investigated in detail at ambient curing condition and compared with OPC
concrete of same grade.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction ing geopolymer binder over OPC; potential reduced greenhouse


gases emissions and utilization of industrial by-products. Geopoly-
Production of ordinary Portland cement (OPC) contributes mer technology can utilize industrial by-products to produce a
around 5 to 7% of total global greenhouse gas emission (Hendriks good construction material. Considering those benefits, geopoly-
et al. 1998; Meyer 2009) which possesses a potential problem for mer is considered as a sustainable binding material (Li et al., 2004;
the global environment. Several researches were done in the past Lloyd and Rangan 2009).
in pursuit of environmental friendly alternative binding material. In last two decades, several studies have been carried out glob-
Geopolymer technology shows a considerable promises as an alter- ally on geopolymers; however, majority of these studies were fo-
native binder to Portland cement as proposed by Davidovits (1989). cused on material characterisation, physical and chemical prop-
Geopolymer is generally formed by reaction of an aluminosilicate erties of the material and the effects of source material (Duxson
compound of geological origins like clay and metakaolin or in- et al. 2007). Recently, more focus is given on investigation of com-
dustrial by-products, such as fly ash and ground granulated blast pressive strength and other mechanical properties of geopolymer
furnace slag (GGBFS or slag) with an alkaline solution. The reac- concrete (Diaz-Loya et al., 2011; Hardjito and Rangan 2005; Sofi
tion of aluminosilicate powder with alkali solution produces a syn- et al. 2007). However, these data are still not sufficient to iden-
thetic alkali aluminosilicate compound which is in amorphous to tify the relationship between different mechanical properties of
semi-crystalline state (Davidovits 1991) having good binding prop- geopolymer concrete. The application of geopolymer concrete in
erty like as calcium silicon hydrate (C-S-H) paste of Portland ce- structural elements is in very limited numbers so far because of
ment concrete. There are two major environmental benefits of us- insufficient proven data.
Geopolymers are generally made from activation of aluminosil-
icate powders by highly concentrated sodium hydroxide and or
E-mail address: kamalnep@hotmail.com sodium silicate solutions, thus called liquid-activated geopolymer.

http://dx.doi.org/10.1016/j.mechmat.2016.09.012
0167-6636/© 2016 Elsevier Ltd. All rights reserved.
K. Neupane / Mechanics of Materials 103 (2016) 110–122 111

Recently, some cement and concrete companies around the world used in concrete industry, geopolymer binder can be mixed in the
have begun the production of geopolymer binder under different similar way to Portland cement and capable of setting and harden-
brand names, such as E-CreteTM (Zeobond 2008), Earth Friendly ing at ambient conditions (Kidd 2009).
Concrete, (Wagners 2010) and BanahCEM (Banah 2011). However, Class F fly ash or metakaolin based geopolymer concrete takes
liquid-activated geopolymer does not appear as a viable replace- very long time to set and harden at ambient conditions, and there-
ment of Portland cement in concrete industry despite its substan- fore most previous research were carried out by heat curing of
tial environmental benefits and superior engineering properties be- geopolymer concrete and mortar (Ahmed et al., 2011; Fernandez-
cause of the difficulties in mixing and handling of this binder. Jimenez et al., 2006; Hardjito and Rangan 2005; Rovnaník 2010).
(Hajimohammadi, Provis & Van Deventer 2008). In order to be utilized in construction, concrete should able to set
and harden at ambient temperature. In general construction, con-
1.1. Limitations of liquid-activated geopolymer binder crete elements are cured in ambient temperature; elevated temper-
ature curing is not practicable and not cost effective. Till date, pub-
Sodium or potassium alkalis are the mostly used as activators lish engineering properties of geopolymer concrete cured at ambi-
in geopolymer binder (Diaz-Loya et al., 2011; Fernandez-Jimenez ent temperature are still limited. Replacement of fly ash by GGBFS
et al., 2006; Hardjito and Rangan 2005). There are two major lim- can make geopolymer concrete set and harden in normal tempera-
itations of liquid activated (two- part) geopolymer binder having ture with significantly reduced setting time and higher early age
sodium or potassium hydroxide and sodium silicate as activator. as well as later strength (Ganapati Naidu et al. 2012; Nath and
Both sodium and potassium hydroxide solutions are hazardous ma- Sarker 2012; Parthiban et al. 2013). It is suggested that presence
terials. They can cause severe burns to the human body on contact of calcium compound accelerate the geopolymerization process
(ERCOWorldwide 2012). Sodium silicate solution is also a moder- (Catalfamo et al. 1997) and there is the coexistence of geopoly-
ately hazardous material, that can cause chemical burns with con- meric gel and calcium-silicate-hydrate (C-S-H) in slag-fly ash based
tact or ingestion (Hartford 2010). Therefore, transporting and han- geopolymer paste, where C-S-H gel is responsible for developing
dling of these liquids can cause a potential physical risk of workers. early age strength (Oh et al. 2010; Yip et al., 2005)
Concentrated sodium hydroxide and potassium hydroxide solutions Fly ash and GGBFS based powder-activated geopolymer binder
are corrosive to materials such as tin, aluminium, zinc, copper, lead can offer following advantages over liquid –activated geopolymer.
and their alloys and also attack glass and some types of plastics
(Helmenstine 2013; Keith 2008). Liquid sodium hydroxide absorbs
• Concrete mixing and handling process is easier than using
carbon dioxide from air and gradually changes to sodium carbon- liquid-activated geopolymer binders and is similar to conven-
ate (OxyChem 2013). They should be stored very securely in air tional Portland cement.
tight vessels made of non-reactive martial.
• Storing and transporting are safer than liquid-activated
Change in concentration of alkali activator in the binder can geopolymer binder.
bring differences in the geopolymerization process and proper-
• Properties of concrete are more consistent due to fixed propor-
ties of concrete as an end product (Mishra et al. 2008; Somna tions of ingredients in binder.
et al. 2011). Increase in concentration of sodium or potassium hy-
• Lower workability loss (up to 2 hours) of concrete than from
droxide generally accelerates the geopolymerization reaction and liquid-activated geopolymers. This is due to slowly release of
results higher strength of concrete but decrease in workability alkali into the binder system and slower rate of initial reaction
(Chindaprasirt et al., 2007; Ryu et al. 2013). Therefore a skilled than in liquid-activated geopolymer.
manpower is needed for mixing and handling of this type of
geopolymer binder in order to maintain the desired concentration 2. Experimental programme
of liquid activator.
The limitations of liquid-activated geopolymer concrete have In this study engineering properties of geopolymer concrete of
prevented its common utilization in the concrete industry although four strength grades (40, 50, 65 and 80 MPa) were investigated
concrete from this binder possesses superior engineering proper- in two different states; fresh and hardened concrete. Mechanical
ties with added environmental benefits. A powder form of geopoly- properties such as compressive strength, indirect tensile strength
mer binder which can be handled and stored safely and mixed in and flexural strength were measured at different age up to 90 days
similar way to conventional OPC concrete is therefore needed for a under standard laboratory temperature (23 °C) curing. Modulus
potential repalcement of OPC (Hajimohammadi et al., 2008). of elasticity and drying shrinkage were measured as deformation
properties.
1.2. Powder-activated geopolymer binders The properties of geopolymer concrete were compared against
the OPC concrete of same grade. In addition, the relationships
In powder-activated geopolymer binder, both activators (pow- of mechanical properties of geopolymer concrete were evaluated
der form), and source materials are blended together in a fixed against equations suggested in concrete standards of current prac-
proportions to make one-part binder. This type of geopolymer tice, such AS 3600 (2009) and ACI 318 (2011).
binder can be safely mixed and handled in the similar way to Port-
land cement because of having chemical activator in solid (pow- 2.1. Materials
der) form. The mixing process of this binder is very similar with
conventional OPC concrete; addition of potable water for desired Two types of powder-activated geopolymers; Geopolymer 1
level of workability. Class F fly ash is the commonly used source (General purpose) and Geopolymer 2 (High early strength) recently
material in combination with ground granulated blast furnace slag developed by Cement Australia Pty Ltd were used as binding ma-
(GGBFS). terials. In these binders, activators (powder form), Class F fly ash
There were two major objectives of developing this binder. The and GGBFS were blended together in fixed proportions. The acti-
first objective was to develop a binding material that has lower vator was a combination of sodium silicate and sodium hydroxide
embodied energy by utilizing industrial by-products such as fly ash in powder form. The activator/source material ratio was around 0.3
and GGBFS. The second objective was to make a binder that would for both binders. The concentration of sodium hydroxide in binders
be physically similar to conventional Portland cement, such that changed with the of water/binder ratio of concrete. For exam-
mixing and handling process would be safe and easy. In order to be ple, the concentration was around 8 M (molarity) in Grade 40 MPa
112 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Table 1
Chemical compositions of fly ash, GGBFS and OPC (type GP).

LOI CaO SiO2 Al2 O3 Fe2 O3 MgO SO3 Na2 O K2 O Others

Fly ash 0.7 3.2 52.7 26.0 12.8 1.4 0.2 0.52 0.79 2.05
GGBFS 0.2 42.6 33.5 13.9 0.9 5.2 1.7 0.26 0.36 1.35
GP 3.5 64 19.2 4.96 3.07 1.14 2.5 0.14 0.41 1.08

Others: SrO, TiO2, Mn2O3 and P2O5; LOI: loss of ignition

Fig. 1. OPC (GP), Geopolymer 1 and Geopolymer 2 binders.

Fig. 2. Particles size distribution of coarse and fine aggregates.

concrete (water/binder ratio 0.45), whereas, it was as high as 13 M 2.2. Mix compositions of concrete
in 80 MPa concrete (water/binder ratio 0.26). Geopolymer 1 con-
tained 70% fly ash and 30% GGBFS by weight, whereas Geopoly- The mix design of both OPC and geopolymer concretes was
mer 2 contained 40% fly ash and 60% of GGBFS by weight. Table 1 done according to BRE mix design method (Teychenné et al. 1997).
shows the chemical compositions of low calcium fly ash and GG- However, amount of binder and water were adjusted for geopoly-
BFS used in geopolymer binders. mer concrete as necessary for a comparable strength and work-
The visual comparison of different binders; GP (General Purpose ability. Both OPC and geopolymer concretes were mixed in the
Portland cement), Geopolymer 1 and Geopolymer 2 is shown in same way in a rotating pan mixer. Concretes form geopolymer
Fig. 1. These geopolymer binders seem physically similar with GP; binders were mixed without addition of any chemical admixtures.
however, they are lighter in colour. Whereas, OPC concretes were produced with addition of normal
Crushed river course aggregates maximum sizes of 10 mm and water reducing (WR) admixture or high range water reducing ad-
20 mm collected in bed of Mary River, Queensland were used for mixture (HWR). The mix compositions of concrete of different
production of concrete for this study. A combination of medium grades are presented in Table 2. The mix design of concrete of
and fine sand was used as fine aggregate in concrete mix from the different grades and binders is based on comparable workability
same source. The particle size distribution of fine and coarse ag- (slump value) and 28 days compressive strength.
gregates is shown in Fig. 2.
Two types of chemical admixtures were used in OPC (control) 2.3. Casting, curing and testing of concrete specimens
concrete to control the water demand; normal water reducer (WR)
and high range water reducer (HWR). No chemical admixtures Sufficient numbers of cylinders having 100 mm diameter by
were used in production of geopolymer concrete of all grades. 200 mm height and flexural beams having 100 mm by 100 mm
K. Neupane / Mechanics of Materials 103 (2016) 110–122 113

Table 2
Mix compositions of concretes of different grades.

Binder type Concrete grade Binder content Aggregate (kg/m3 ) Water content Slump Chemical admixtures
(MPa) (kg/m3 ) (kg/m3 ) (mm) (litre/m3 )

20 mm 10 mm Coarse sand Fine sand

OPC 40 320 660 460 550 200 150 100 WR 1.12


OPC 50 395 640 440 515 195 174 100 WR 1.38
OPC 65 500 610 420 490 180 179 105 HWR 1.75
OPC 80 555 605 425 495 180 161 140 HWR 2.50
Geopolymer 1 40 310 700 490 585 210 112 120 –
Geopolymer 1 50 330 710 490 575 200 112 120 –
Geopolymer 1 65 360 705 480 565 190 109 110 –
Geopolymer 1 80 480 650 455 520 190 115 160 –
Geopolymer 2 40 280 705 490 595 210 132 110 –
Geopolymer 2 50 295 720 500 580 200 124 120 –
Geopolymer 2 65 360 705 475 565 190 112 120 –
Geopolymer 2 80 455 660 460 535 185 117 100 –

Fig. 3. Sealing of geopolymer concrete specimens (a) cylinders (b) flexural beams (c) shrinkage prisms.

cross section and 350 mm length were cast to determine compres- readings were taken on 7, 14, 21, 28, 56 and 112 days from the
sive strength, indirect tensile strength, flexural strength and modu- initial reading.
lus of elasticity of concrete from different binders and grades. Con- Both geopolymer and OPC concrete were tested for mechanical
crete specimens from both OPC and geopolymer concrete were left properties (compressive, indirect tensile and compressive strength)
in laboratory temperature (23 °C) until demoulding. Concrete cylin- according to relevant Australian standards at the ages of 1, 3, 7,
ders were demoulded after 24 hours, whereas, flexural beams were 28, 56 and 90 days, except modulus of elasticity of concretes was
demoulded after 48 hours. The concrete prisms for drying shrink- measured only for 28 days.
age measurement were prepared according to Australian Standard
1012.13 (1992) at temperature of 23 °C.
As water is a by-product of geopolymerization process 2.4. Testing of engineering properties of concrete
(Davidovits 1999), the initial moisture available in concrete may
be sufficient for further geopolymerization process. Therefore, after In this study, engineering properties of concrete were measured
the demouling, small amount of water was sprayed into geopoly- in both, plastic state (fresh concrete) and hardened state.
mer concrete specimens and they were sealed by plastic sheet im- Workability of fresh concrete was measured by slump measure-
mediately for curing as shown in Fig. 3. The sealed concrete cylin- ment according to AS 1379 (2007).
ders were curried at temperature of 23 °C until testing. On the Compressive strength, indirect tensile strength and flexural
other hand, OPC concrete specimens were cured by immersing in strength were investigated under hardened concrete properties.
lime saturated water at 23 °C until testing. Compressive strength of concrete from three different binders were
The initial curing (first 7 day after demoulding) of shrinkage measured for each grade of concrete (grades 40, 50 65 and 80 MPa)
prism from geopolymer concrete was also carried out by sealing. at 1, 3, 7, 28, 56 and 90 days of ages according to AS-1012.9
The prisms were made saturated by immersing in water for 2 to (1999) in order to study the development of strength in early as
3 minutes prior to sealing. In the case of OPC concrete, shrink- well as later age. Similarly, indirect tensile strength and flexural
age prisms were cured by submerging in lime saturated water in strength of concrete from each binder were measured for 3, 7, 28,
the similar way to the cylinders for 7 day. After the initial curing, 56 and 90 days of ages for all four grades of concrete according to
both OPC and geopolymer concrete prisms were taken out from AS-1012.10 (20 0 0) and AS-1012.11 (20 0 0), respectively. Minimum
curing for initial readings and then stored in a room having stan- 3 numbers of specimens were tested to determine the mechanical
dard temperature of 23 °C and 50% relative humidity according to properties at any age, whereas minimum 5 specimens were tested
Australian Standard 1012.13 (1992). Subsequent drying shrinkage at 28 days for all types and grades of concrete.
114 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Table 3 mer 1 required slightly less (5 to 7%) amount of water than from
Table of symbols.
Geopolymer 2 binder.
Ec Modulus of elasticity of concrete The higher proportion of fly ash present in source material of
f’c Characteristic compressive strength of concrete at 28 days geopolymer binder and the nature of geopolymerization process
fcm Mean compressive strength of concrete
to produce water as a by-product are the two major factors to
f’ct Characteristic axial tensile strength of concrete
fct Mean axial tensile strength of concrete decrease the water demand in geopolymer concrete. It is under-
f’r Characteristic flexural tensile strength of concrete stood that replacement of Portland cement by fly ash improves the
fr Mean flexural tensile strength of concrete workability of concrete because of round shaped fly ash particles
f’sp Characteristic indirect tensile strength of concrete (Neville 1995). In addition, fly ash can be dispersed easily in the
fsp Mean indirect tensile strength of concrete
alkaline environment without addition of admixture (Chindaprasirt
P Density of concrete cylinder
f3 Mean compressive strength of concrete at 3 days of age et al., 2007). The difference in proportion of fly ash in source ma-
f7 Mean compressive strength of concrete at 7 days of age terial of Geopolymer 1 and Geopolymer 2 is the major factor to
f28 Mean compressive strength of concrete at 28 days of age make the difference in water demand in concrete from these two
f90 Mean compressive strength of concrete at 90 days of age
binders. Lower water demand in geopolymer concrete can result
in lower porosity of harden concrete, which can make geopolymer
concrete more durable than OPC concrete.
Modulus of elasticity and drying shrinkage of geopolymer and Retention of workability is one of the important characteris-
OPC concretes of different grades were measured under deforma- tics of concrete which facilitates the transportation of concrete to
tion properties. Modulus of elasticity of concrete of all grades was a longer distance from batching plant and keep concrete work-
measured at 28 days of age according to AS-1012.17 (1997). Drying able. Generally Retention of workability is measured by slump loss;
shrinkage of concretes of all grades was measured for 7, 14, 21, 28, lower slump loss indicates higher workability retention. Slump loss
56 and 112 days (counted after completion of initial 7 days curing) of different grades of concrete from different binders with respect
according to AS-1012.13 (1992). to time for up to 2 hours is shown in Fig. 5. Data points in this
figure suggest that powder-activated geopolymer concrete possess
3. Results and analysis better workability retention than OPC concrete of same grade. It
is obvious that, higher grade concrete contains higher amount of
The mix design of concrete was based on comparable workabil- binder and shorter setting time than lower grade concrete, and
ity of concretes (average 120 mm slump) for all grades. Concretes therefore shows higher slump loss. Hence, Grade 40 MPa concrete
of from different binders and grades were produced with the same showed better retention of workability than Grade 50 and 65 MPa
source aggregates in same proportions. Therefore, it is assumed concrete in all binders. The slump loss in Grade 40 MPa geopoly-
that aggregate type was not a factor for the variation of results mer concrete was around 50% of initial slump when compared
of fresh and hardened concrete properties from different binders. with 80% loss in OPC concrete of same grade. It is assumed that
A list of symbols has been presented in Table 3 to define the the powder activated geopolymer has a slower release of alkali
parameters. into the binder system which results a slower rate of initial reac-
tion and therefore concrete remains workable for a longer period
3.1. Fresh concrete properties-workability (Collins and Sanjayan 1999).

The amount of water required in concretes from different 3.2. Compressive strength
binders for comparable workability is presented in Fig. 4; some
of the data from trial mix designs were also included in this plot. Compressive strength development with time of Grades 40, 50
There was a measurable difference between the water content in and 65 MPa concretes from different binders is shown in Fig. 6.
geopolymer and OPC concretes for same workability level, although Except early age (1 to 3 days) strength, compressive strength de-
OPC concrete were prepared with water reducing admixtures. The velopment pattern in concrete of Grades 80 MPa was also simi-
average water content in geopolymer concrete was 116 kg/m3 com- lar to these grades. Medium grade (40 and 50 MPa) geopolymer
pared to around 166 kg/m3 in OPC concrete. Geopolymer concrete concrete attained slightly lower compressive strength at early age
required around 30% less water than OPC concrete for the com- than OPC concrete of same strength grade. In Grade 40 MPa, both
parable workability, although OPC concrete were produced with Geopolymer 1 and Geopolymer 2 concrete exhibit relatively lower
addition of water reducing admixtures. Concrete from Geopoly- early age strength than OPC concrete which may be the results of

Fig. 4. Workability of fresh concrete of different grades.


K. Neupane / Mechanics of Materials 103 (2016) 110–122 115

Fig. 5. Workability loss with time of different concrete.

Fig. 6. Compressive strength development of different grades of concrete.

Table 4
Compressive strength and age factors of different grades concrete.

Grades (MPa) Concrete ID w/b ratio Compressive strength (MPa) Age factors

3 day 7 day 28 day 90 day f3 /f28 f7 /f28 f90 /f28

40 OPC 0.46 34.5 44.0 52.5 56.5 0.66 0.84 1.08


Geopolymer 1 0.36 18.0 37.0 58.0 67.0 0.31 0.64 1.16
Geopolymer 2 0.47 13.0 29.0 50.0 58.5 0.26 0.58 1.17
50 OPC 0.43 41.0 51.0 62.0 66.0 0.66 0.82 1.06
Geopolymer 1 0.34 26.0 45.5 67.0 80.5 0.39 0.68 1.20
Geopolymer 2 0.42 23.5 43.5 63.5 76.0 0.37 0.69 1.20
65 OPC 0.36 53.5 66.5 78.0 81.5 0.69 0.85 1.04
Geopolymer 1 0.30 35.5 53.0 73.5 85.5 0.48 0.72 1.16
Geopolymer 2 0.31 44.5 59.0 82.5 95.5 0.54 0.72 1.16
80 OPC 0.29 70.5 78.5 90.5 97.0 0.78 0.87 1.07
Geopolymer 1 0.24 54.5 73.5 97.0 115.5 0.56 0.76 1.19
Geopolymer 2 0.26 53.5 71.5 96.0 112.5 0.56 0.74 1.17

lower binder content in concrete mix and higher water/binder ra- The compressive strength results of different grades of concrete
tio. On the other hand, later age (after 7 days) strength develop- from each binder are presented in Table 4. Data points in this ta-
ment rate was significantly higher in geopolymer concrete than in ble suggest that the age factor (ratio of strength at any age to
OPC concrete, hence higher ultimate strength is achieved. The con- 28- day strength) of geopolymer concrete at early age (3 days) is
tinuous growth in compressive strength of geopolymer concrete is lower than OPC concrete of same grade; however, is gradually in-
the result of continuity of geopolymerization process for a longer creased with the increase in concrete grade. As such, high-strength
period. geopolymer concrete (65 and 80 MPa) attained significant early age
116 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Fig. 7. Indirect tensile strength development of Grade 50 and 65 MPa concretes.

strength equal as OPC concrete of same grade hence the age fac- chanical properties of OPC concrete are defined in several stan-
tor is similar with OPC concrete. The average growth in compres- dards and published papers.
sive strength from 28 to 90 days in geopolymer concretes was 18% ACI 363R (1992) suggested indirect tensile strength of concrete
while compared with average 6% growth in OPC concrete for the as follows.
same period. √  
f ’sp = 0.59 f ’c MPa for 21 MPa < fc < 83 MPa (1)
OPC concrete with only Portland cement (no supplementary ce-
mentitious materials added) is not common in general construc- where,
tion practice in Australia. A partial replacement of OPC cement by f’sp = characteristic indirect tensile strength of concrete
fly ash, ground granulated blast furnace slag and silica fume or by f’c = characteristic compressive strength of concrete
their combinations has been practiced in recent decades (Toutanji
et al. 2004). Partial replacement of Portland cement (20% to 40% Australian Standards 3600 (2009) suggested a relationship be-
by weight) by fly ash and slag or their combinations resulted sig- tween tensile and compressive strength of concrete as follows.

nificant reduction in early age strength in some previous studies f ’ct = 0.36 f ’c MPa (for f ’c : 20 MPa to 100 MPa) (2)
(Berry & Malhotra 1980; Siddique 2008; Johari et al. (2011)) sug-
where,
gested around 40% reduction in 1 day compressive strength and
f’ ct = characteristic axial tensile strength of concrete
20% in 3 days strength due to replacement of Portland cement by
Some researchers have proposed different models for the
20% of fly ash. Therefore, comparing with the early age results of
relationship between compressive strength and indirect tensile
OPC concrete having 20 to 40% of fly ash or slag, geopolymer con-
strength of geopolymer concrete based on their results. As the fi-
crete can attain same level of early age strength to OPC concrete
nal products of geopolymerization are influence by various factors
of same grade.
such as type and reactivity of source material and type and con-
The amount of binder required in geopolymer concrete was
centration of alkaline activators used (Duxson et al. 2006) there
significantly lower than OPC concrete of same strength grade
was a variation in result.
as shown in Table 2. For example, in Grade 50 MPa concrete, √
295 kg/m3 of Geopolymer 2 binder and 395 kg/m3 of OPC devel- Sofi et al. (2007 ) ) : f ’sp = 0.5 f ’c MPa (3)
oped compressive strength of 63.5 MPa and 62.0 MPa respectively,

at 28 days. Concrete from Geopolymer 1 required slightly higher Tempest (2010 ) ) : f ’sp = 0.616 f ’c MPa (4)
amount of binder than from Geopolymer 2 for the comparable
Concrete Institute of Australia (2011), suggests a similar equa-
strength because of the difference in slag content in source ma-
tion to AS 3600 (2009) for geopolymer concrete.
terials. √
fct = 0.4 fcm MPa (5)
The relationship between indirect tensile strength and compres-
3.3. Indirect tensile strength development sive strength of different types of concretes is plotted in Fig. 8.
Indirect tensile results at different ages (3 to 90 days) were in-
Development of tensile strength in Grade 50 and 65 MPa con- cluded in this plot. The line of ACI 363R is closely located to the
cretes is shown in Fig. 7. The development of tensile strength was data points of geopolymer concretes in lower range of compressive
in similar pattern for all grades of concretes. Unlike to compressive strength; however, there is a bigger gap in high strength range.
strength development, geopolymer concrete developed comparable Australian Standard 3600 (2009) estimates lower indirect tensile
early age (1–3 days) indirect tensile strength to OPC concrete of strength than geopolymer as well as OPC concretes of this study.
same grade and there was a higher increase in geopolymer con- Considering previous models and experimental results an equa-
crete in later age. As a result, geopolymer concrete attained around tion for Geopolymer 1 and Geopolymer 2 can be proposed as fol-
15% higher indirect tensile strength at 28 days and around 20% lowing concretes.
higher at 90 days then OPC concrete of same grade. √
As geopolymer is the new binding material, no equations has fsp = 0.67 fcm MPa (for fcm : 20 MPa to 110 MPa) (6)
been suggested to define the relationships of mechanical proper- In Fig. 8 all the data points of Geopolymer 1 and Geopolymer
ties of geopolymer concretes in structural concrete standards such 2 normal workable concretes were located within the ± 10% range
as Australian Standard 3600 (2009), ACI 318 (2011), and European from this proposed line. The line of ACI 363R (1992) is close with
Standard (EN) 1992-1-1 (2004). The relationships between me- lower error range (−10%) of proposed model.
K. Neupane / Mechanics of Materials 103 (2016) 110–122 117

Fig. 8. Relation between indirect tensile strength and compressive strength.

Fig. 9. Flexural strength development of Grade 50 and 65 MPa concretes.

3.4. Flexural strength development Most of the previous investigations reported higher flexural
strength of geopolymer concrete than OPC concrete of same com-
The development of flexural strength of Grades 50 MPa and pressive strength (Fernandez-Jimenez et al., 2006; Raijiwala and
65 MPa concretes are shown in Fig. 9. Flexural strength devel- Patil 2011).
opment in other grades of concrete was also in similar pattern. Sofi et al. (2007) proposed a very similar equation to AS 3600
Normal workable geopolymer concrete possessed relatively higher relationship to estimate flexural strength of geopolymer concrete.
flexural strength than OPC concrete of same grade. Similar to in-

direct tensile strength, there was significant growth in flexural f ’r = 0.6 f ’c MPa (9)
strength of geopolymer concrete at 3 days with an average of 60%
Diaz-Loya et al., (2011) proposed following relationship of flex-
of 28 days strength. Geopolymer concrete possessed 15% higher
ural and compressive strength of geopolymer concrete.
flexural strength than OPC concrete of same grade.

Similar to splitting tensile strength, there is no specific equa- fr = 0.69 fcm MPa (10)
tion suggested in concrete standards of current practice to define
The relationship between flexural strength and compressive
the relationship of flexural strength and compressive strength of
strength geopolymer concrete as well as OPC concretes of this
geopolymer concrete. There are some equations and relationship of
study is presented in Fig. 10.
compressive strength and flexural strength of OPC based concrete
A relationship between flexural strength and compressive
recommended in different concrete standards.
strength of geopolymer concrete can be proposed for the range of
According to ACI 363R (1992), the relation between flexural and
20 to 100 MPa of mean compressive strength as following.
compressive strength of concrete applicable from Grade 21 MPa to

83 MPa is as following. Flexural strength(fr ) = 0.88 fcm MPa (11)
√ In Fig. 10, all the data points of geopolymer normal workable
f ’r = 0.94 f ’c MPa (7)
concretes are located within ± 10% range from the proposed model.
where,
f’ r = characteristic flexural strength of concrete 3.5. Modulus of elasticity
According to AS 3600 (2009) the flexural strength of concrete
(Grades 20 MPa to 100 MPa) can be estimated as follows. The experimental results for modulus of elasticity of concretes
of different grades are presented in Fig. 11, each data represents

f ’r = 0.6 f ’c MPa (8) the average results of minimum three numbers of specimens.
118 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Fig. 10. Relationship of flexural strength and compressive strength.

Fig. 11. Experimental results of modulus of elasticity.

Fig. 11 shows that geopolymer concrete possessed similar mod- terials and activators. In most of the research, geopolymer concrete
ulus of elasticity to the OPC concrete of same strength grades. specimens were cured in elevated temperature in order to develop
In most of the concrete standards of current practice, modulus high early strength. Hardjito and Rangan (2005) and Fernandez-
of elasticity of concrete is estimated as a function of density and Jimenez et al., (2006) found that accelerated cured fly ash based
compressive strength. geopolymer concrete possessed relatively lower modulus of elas-
Modulus of Elasticity recommended by ACI 318–2011 (2011) is ticity then recommended by AS 3600. Sofi et al., (2007) suggested
as following. that equation recommended by AS 3600 was also applicable to fly
√ ash and slag based geopolymer concrete. Neupane et al. (2015) also
Ec = 0.043ρ 1.5 f ’c MPa (12)
reported that modulus of elasticity of flay ash based geopolymer
Modulus of elasticity recommended by Australian Standard concrete can be closely estimated using AS 3600 (2009).
360 0–20 09 (20 09) is as following Previous models and results are plotted in Fig. 13. Fernandez-
√ Jimenez et al., (2006) work on fly ash based geopolymer concrete
Ec = ρ 1.5 (0.024 fcm + 0.12) MPa; when fcm > 40 MPa (13)
found much lower modulus of elasticity then OPC concrete. Mod-
Where, ulus of elasticity of geopolymer concrete measured by Diaz-Loya
et al., (2011) was also relatively lower than modulus of elasticity
ρ = concrete cylinder density (kg/m3 ) and
of this study for the same compressive strength. The relationship
Ec = modulus of elasticity
model proposed by Tempest (2010) was located far below than
Above relationships are plotted in Fig. 12 with some additional data points of this study.
data of this study. In order to measure the compliance of results, a comparison of
In Fig. 12, the line representing equation of Australian Standard measured and estimated modulus of elasticity of geopolymer con-
360 0 (20 09) is very close with geopolymer as well as OPC con- crete was made for Australian Standard 3600 (2009) and ACI 318
cretes data points. The estimated modulus of elasticity according (2011). In this calculation, estimated values of modulus of elasticity
to Australian Standard 360 0 (20 09) has an error range of ± 20%. using Australian Standard 3600 (2009) were very close to the mea-
However, in this figure, data points of all geopolymer and OPC con- sured results of geopolymer concrete. The measured/estimated ra-
cretes are located well inside the ± 10% error range of that line. tios were between 0.93 to 1.02 with an average of 0.98 for geopoly-
Modulus of elasticity value is said to be overestimated by ACI 318 mer concrete. Hence, modulus of elasticity for geopolymer concrete
(2011) for geopolymer concrete because none of any data points can be closely estimated using AS 3600 equation. However, there
are located within this line. was a relatively bigger difference between the estimated and mea-
Investigations were done in the past to determine modulus of sured values of modulus of elasticity using ACI 318 (2011) equation
elasticity of geopolymer concrete made from different source ma- with an average of 0.86.
K. Neupane / Mechanics of Materials 103 (2016) 110–122 119

Fig. 12. Relationship of compressive strength and modulus of elasticity.

Fig. 13. Modulus of Elasticity of various geopolymer binder concretes.

3.6. Drying shrinkage concrete of same grade at ambient curing condition. There was a
common pattern of decreasing drying shrinkages with the increase
Drying shrinkage is one of the important serviceability prop- in concrete strength grades, hence drying shrinkage result of Grade
erties of structural concrete. Higher drying shrinkage may re- 80 MPa concrete was found to be lower than any other grade for
sult in surface cracking and curling of concrete structure which all binders. The rate of increase of drying shrinkage with age was
eventually leads to structural and durability problems. Wallah similar in concretes of all grades; higher at early ages and gradu-
and Rangan (2006) recorded significantly low drying shrinkage ally decreasing with time.
of heat cured liquid-activated fly ash based geopolymer con- This experiment found a relatively higher amount of drying
crete which was around 100 microstrain for 1 year. The lower shrinkage of geopolymer as well as OPC concrete than some pre-
shrinkage of heat cured concrete can be justified by the evapo- vious experiment in OPC concrete (Al-Attar 2008; Shah et al.,
ration of water from micro-pores during heat curing. Hence, less 1992). This may be due to different types of aggregates used and
water remains inside to contribute post-curing drying shrinkage other factors of concrete mix design such as cement type and wa-
(Davidovits 1999). Fernandez-Jimenez et al., (2006) also measured ter/binder ratio. Several previous investigations suggested that ag-
significantly lower (around 200 microstrains) drying shrinake of fly gregate properties (shrinkage and water absorptions) are also ma-
ash based geopolymer concrete up to 90 days of age. In their work, jor factors influencing drying shrinkage of concrete (Hansen and
geopolymer concrete samples were cured at 85 °C for 20 hours. Nielsen 1965; Hyodo et al. 2013). The drying shrinkage would be
Collins and Sanjayan (1999) reported a relatively higher amount different if different types of aggregates, such as limestone and
of drying shrinkage (around 1500 microstrain) of ambient cured basalt were used which usually give less drying shrinkage than
GGBFS based geopolymer concrete when compared to OPC con- sandstone aggregate. A detailed investigation in drying shrinkage
crete (around 700 microstrain) of same grade. On the other hand, of concrete from Greywacke sandstone aggregate by Mackechnie
Albitar et al. (2015) reported significantly lower (less than 100 mi- (2006)) showed an average 800 microstrain of drying shrinkage at
crostrain) drying shrinkage of fly ash alone and fly ash with granu- 56 days for concretes having water/binder ratio 0.5 to 0.7 which is
lated lead smelter slag based geopolymer concrete for 90 days un- similar to drying shrinkage results of the 40 MPa grade geopolymer
der ambient curing conditions. as well as OPC concrete of this study.
The drying shrinkage results of Grades 40, 50 and 65 MPa con- Australian Standard 1379 (2007) has specified the 56 days dry-
crete is shown in Fig. 14. Data points in this figure show that ing shrinkage of normal class concrete (50 MPa or less) should be
geopolymer concrete exhibited similar drying shrinkage to OPC less than 10 0 0 microstrain. Data points of Fig. 14 and results from
120 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Fig. 14. Drying shrinkge result of 50 and 65 MPa concretes at ambient curing.

Grade 80 MPa show that geopolymer concrete of any grade has not 6. Conclusions
exceeded this limit. Drying shrinkage in geopolymer concrete is
therefore within the acceptable limits suggested in current code Fly ash and GGBFS based powder-activated (one-part) geopoly-
of practice. mer binder is a viable sustainable alternative of Portland ce-
ment which can be used in general construction as a structural
or non-structural concrete. This binder can be mixed and han-
4. Durability properties dled in a similar way to OPC and concrete from this binder set
and harden in ambient condition, hence overcomes the limitations
Durability of geopolymer concrete is one of the major signifi- of fly ash based liquid-activated (two-part) geopolymer. Cost of
cance over OPC concrete because this binder system does not rely powder-activated geopolymer binder is slightly (15%) higher than
on calcuim compund and is free from C3 A (CIA 2011; Davidovits Portland cement; however, significantly lower amount (30%) of
1994). Several study in the past suggested that geopolymer con- binder is needed for the same strength grade concrete. Geopolymer
crete from this binder possess excellent ressistance against sul- concrete of any strength grade can be produced without addition
phate and acidic environment (Bakharev 2005; Wallah and Ran- of chemical admixtures.
gan 2006). There was no alklai silica reaction in both fly ash or Concrete with geopolymer binders required significantly less
slag based geopolymer concrete system (Fernandez-Jimenez et al., amount of water and binder for the comparable workability and
2006; Talling and Brandstetr 1989). A durabilty study of concretes 28-day compressive strength to OPC concrete. Geopolymer con-
from both Geopolymer 1 and Geopolymer 2 (same binders of crete showed better workability retention up to 2 hours than OPC
this study) showed excellent resistacne of geopolymer cocnrete to concrete of same grade.
acid and sulphate attack as well as free from alklai silica reaction Compressive strength development of medium grade geopoly-
(McKenzie 2011). mer concrete was slightly lower at early age compared to the OPC
(control) concrete of this study. However, comparing against the
early age results of OPC concrete having 20 to 40% fly ash or slag,
5. Economical issues geopolymer concrete can attain same level of early age strength.
High-strength geopolymer concrete exhibited similar early age
Industrial by-products, such as fly ash and GGBFS are the strength to control concrete. The later age growth in compressive
commonly used source materials (aluminosilicate compounds) for strength was significantly higher in geopolymer concrete with an
geopolymer binder because of their worldwide availability (Heath average of 18% for 28 to 90 days period.
et al. 2013). Duxson et al. (2006) suggested that fly ash or slag Geopolymer concrete exhibited 15 to 20% higher indirect tensile
based geopolymer can be 10–30% economical than Portland ce- and flexural strength than OPC concrete of same grade. Both indi-
ment because of using cheaper source materials. However, cost of rect tensile and flexural strength geopolymer concrete were higher
alkali activator and their transportation can increase the price of than estimated value using Australian Standard 3600 (2009).
geopolymer binder. Considering this factor in Australian context, Therefore, different relationships were proposed for geopolymer
price of geopolymer binder could be higher up to 14% than Port- concrete.
land cement (McLellan et al. 2011). Cost of the powder-activated Geopolymer concrete exihibited simialr modulus of elasticity to
geopolymer binders (Geopolymer 1 and Geopolymer 2) used in the OPC concrete of same strength grade. The measured modu-
this study could be around 15% higher than Portland cement lus of elasticity of geopolymer concretes of this study were very
(Kidd 2009). However, experimental result showed that around close (within 10%) to the estimated values using Australian Stan-
30% less amount of geopolymer binder is needed for the same dard 360 0 (20 09). Therefore, existing standard can be used to es-
grade concrete when compared with Portland cement. Further- timate the modulus of elasticity of geopolymer concrete.
more, geopolymer concrete does not need chemical admixtures to Drying shrinkage of geopolymer concrete was similar to OPC
control water demand for sufficient workability. These two factor concrete of same grade. The drying shrinkage of geopolymer con-
could make geopolymer concrete economical than OPC concrete crete of all grades was less than 10 0 0 microstrain at 56 days, hence
with added environmental benefits and superior concrete perfor- complied with Australian Standard 1379 (2007).
mance.
K. Neupane / Mechanics of Materials 103 (2016) 110–122 121

Experiment results showed that geopolymer concrete possess a Ganapati Naidu, P., Prasad, A.S.S.N., Adiseshu, S., Satayanarayana, P.V.V., 2012. ‘A
better engineering performance than OPC concrete, therefore it can study on strength properties of geopolymer concrete with addition of G.G.B.S’.
Int. J. Eng. Res. Develop. 2 (4), 19–28.
be used in structural concrete as a proven building material. Hajimohammadi, A., Provis, J.L., Van Deventer, J.S., 2008. ‘One-part geopolymer
mixes from geothermal silica and sodium aluminate’. Ind. Eng. Chem. Res. 47
Acknowledgement (23), 9396–9405.
Hansen, T.C., Nielsen, K.E.C., 1965. ‘Influence of aggregate properties on concrete
shrinkage’. In: ACI Journal Proceedings, 62. ACI.
The author like to express sincere acknowledgment to Cement Hardjito, D., Rangan, B.V., 2005. ‘Development and Properties of Low-Calcium Fly
Australia Pty. Ltd., Darra, QLD for the financial and material support Ash-Based Geopolymer Concrete’. Curtin University of Technology, Perth, Aus-
tralia.
as well as laboratory facility provided in this research. Hartford, J., 2010 ‘What Are the Dangers of Sodium Silicate?’ LIVE-
STRONG.com viewed 12/11/2013 http://www.livestrong.com/article/
References 293701- what- are- the- dangers- of- sodium- silicate.
Heath, A., Goodhew, S., Paine, K., Lawrence, M., Ramage, M., 2013. ‘The potential for
ACI-318, 2011. Building Code Requirements for Structural Concrete, ACI Committee using geopolymer concrete in the UK’. In: Proceedings of the ICE - Construction
318. American Concrete Institute, Farmington Hills, MI, USA. Materials, 166, pp. 195–203.
ACI-363R, 1992. State-of-the-Art Report on High-Strength Concrete, ACI Committee Helmenstine, A.M., 2013. ‘Drain cleaner can dissolve glass’ About.com
363. American Concrete Institute, Farmington Hills, MI, USA. Chemistry viewed 12/11/2013 http://chemistry.about.com/b/2013/08/18/
Ahmed, M.F., Nuruddin, M.F., Shafiq, N., 2011. ‘Compressive strength and workability drain- cleaner- can- dissolve- glass.htm.
characteristics of low-calcium fly ash-based self-compacting geopolymer con- Hendriks, C.A., Worrell, E., De Jager, D., Blok, K., Riemer, P., 1998. Emission reduction
crete’. Int. J. Civil, Environ., Struct., Constr. Archit. Eng. 5 (2), 64–70. of greenhouse gases from the cement industry. In: Proceedings of the Fourth
Al-Attar, T.S., 2008. ‘Effect of coarse aggregate characteristics on drying shrinkage of International Conference on Greenhouse Gas Control Technologies, pp. 939–944.
concrete’. J. Eng. Technol. 26 (2). Hyodo, H., Tanimura, M., Sato, R., Kawai, K., 2013. Evaluation of effect of aggregate
Albitar, M., Ali, M.M., Visintin, P., Drechsler, M., 2015. ‘Effect of granulated lead properties on drying shrinkage of concrete. Third International Conference on
smelter slag on strength of fly ash-based geopolymer concrete’. Constr. Build- Sustainable Construction Materials and Technologies, Kyoto, Japan, August 2013
ing Mater. 83, 128–135. paper presented to the.
AS-1012.9, 1999. Methods of Testing Concrete; Method 9: Determination of the Johari, M.A.M., Brooks, J.J., Kabir, S., Rivard, P., 2011. Influence of supplementary ce-
Compressive Strength of Concrete Specimens. Standards Australia International mentitious materials on engineering properties of high strength concrete. Con-
Ltd, Sydney, NSW. str. Building Mater. 25 (5), 2639–2648.
AS-1012.10, 20 0 0. Methods of Testing Concrete; Method 10: Determination of Indi- Keith., 2008. Sodium hydroxide/caustic soda/lye: useful stuff Practical Mech-
rect Tensile Strength of Concrete Cylinders (Brazil or splitting test). Standards anist.com viewed 12/11/2013 http://www.practicalmachinist.com/vb/
Australia International Ltd, Sydney, NSW. gunsmithing/sodium- hydroxide- caustic- soda- lye- useful- stuff- 158041/
AS-1012.11, 20 0 0. Methods of Testing Concrete; Method 11: Determination of the Kidd, P., 2009. One-Part Dry Mix Geopolymer Binder. Cement Australia Pty. Ltd,
Modulus of Rupture. Standards Australia International Ltd, Sydney, NSW. Darra, QLD.
AS-1012.13, 1992. Methods of testing concrete; Method 13: Determination of the Li, Z., Ding, Z., Zhang, Y., 2004. Development of sustainable cementitious materials.
drying shrinkage of the concrete samples prepared in the field or in the labora- In: Proceedings of the International Workshop on Sustainable Development and
tory. Standards Australia International Ltd, Sydney, NSW. Concrete Technology, pp. 55–76.
AS-1012.17, 1997. Methods of Testing Concrete; Method 17: Determination of the Lloyd, N., Rangan, V., 2009. Geopolymer concrete-sustainable cementless concrete.
Static Cord Modulus of Elasticity and Poisson’s Ratio of Concrete Specimens. ACI Special Pub. 261.
Standards Australia International Ltd, Sydney, NSW. Mackechnie, J.R., 2006. Shrinkage of concrete containing Greywacke sandstone ag-
AS-1379, 2007. Specification and Supply of Concrete. Standards Australia Interna- gregate. ACI Mater. J. 103 (5), 390–396.
tional Ltd, Sydney, NSW. McKenzie, C., 2011. A Report on Durability Characteristic of Geopolymer Concrete.
AS-360 0, 20 09. Concrete Structures. Standards Australia International Ltd, Sydney, Griffith University, Gold Coast, QLD Industrial Affiliates Program 2011.
NSW. McLellan, B.C., Williams, R.P., Lay, J., van Riessen, A., Corder, G.D., 2011. Costs and
Bakharev, T., 2005. Durability of geopolymer materials in sodium and magnesium carbon emissions for geopolymer pastes in comparison to ordinary portland ce-
sulfate solutions. Cement Concrete Res. 35 (6), 1233–1246. ment. J. Cleaner Prod. 19 (9-10), 1080–1090.
Banah, 2011. BanahCEM: Geopolymer Cement System-Technical Data Sheet. banah Meyer, C., 2009. The greening of the concrete industry. Cement Concrete Compos.
UK Ltd., Ballyclare, County Antrim, N. Ireland. 31 (8), 601–605.
Berry, E.E., Malhotra, V.M., 1980. Fly ash for use in concrete-a critical review. In: ACI Mishra, A., Choudhary, D., Jain, N., Kumar, M., Sharda, N., Dutt, D., 2008. Effect of
Journal Proceedings, 77. ACI. concentration of alkaline liquid and curing time on strength and water absorp-
Catalfamo, P., Pasquale, S.D., Corigliano, F., Mavilia, L., 1997. Influence of the calcium tion of geopolymer concrete. ARPN J. Eng. Appl. Sci. 3 (1), 14–18.
content on the coal fly ash features in some innovative applications. Resour. Nath, P., Sarker, P.K., 2012. Geopolymer concrete for ambient curing condition. Pro-
Conser. Recycl. 20 (2), 119–125. ceeding of the Australasian Structural Engineering Conference 2012.
Chindaprasirt, P., Chareerat, T., Sirivivatnanon, V., 2007. Workability and strength Neupane, K., Sriravindrarajah, R., Baweja, D., Chalmers, D., 2015. Effect of curing on
of coarse high calcium fly ash geopolymer. Cement. Concrete Compos. 29 (3), the compressive strength development in structural grades of geocement con-
224–229. crete. Constr. Building Mater. 94, 241–248.
CIA, 2011. Recommended Practice: Geopolymer Concrete (Z16). Concrete Institute of Neville, A.M., 1995. Properties of Concrete, 4th edn Addison Wesley Longman lim-
Australia, National Office, North Sydney NSW. ited, London, England.
Collins, F.G., Sanjayan, J.G., 1999. ‘Workability and mechanical properties of alkali Oh, J.E., Monteiro, P.J.M., Jun, S.S., Choi, S., Clark, S.M., 2010. The evolution of
activated slag concrete’. Cement Concrete Res. 29 (3), 455–458. strength and crystalline phases for alkali-activated ground blast furnace slag
Davidovits, J., 1989. ‘Geopolymers and geopolymeric materials’. J. Thermal Anal. and fly ash-based geopolymers. Cement Concrete Res. 40 (2), 189–196.
Calorimetry 35 (2), 429–441. OxyChem, 2013. ‘OxyChem Material Safety Data Sheet-Potassium Hydroxide’. Occi-
Davidovits, J., 1991. ‘Geopolymers- Inorganic polymeric new materials’. J. Thermal dental Chemical Corporation (OxyChem), Dallas, Texas, USA.
Anal. Calorimetry 37 (8), 1633–1656. Parthiban, K., Saravanarajamohan, K., Shobana, S., Bhaskar, A.A., 2013. Effect of re-
Davidovits, J., 1994. ‘Properties of geopolymer cements’. In: First International Con- placement of slag on the mechanical properties of fly ash based geopolymer
ference on Alkaline Cements and Concretes, 1, pp. 131–149. concrete. Int. J. Eng. Technol. (IJET) 5 (3), 2555–2559.
Davidovits, J., 1999. ‘Chemistry of geopolymeric systems, terminology’. In: Geopoly- Raijiwala, D.B., Patil, H.S., 2011. Geopolymer concrete: a concrete of the next decade.
mere ’99 Proceedings, 99, pp. 9–39. J. Eng. Res. Stud. 2 (1), 19–25.
Diaz-Loya, I.E., Allouche, E.N., Vaidya, S., 2011. ‘Mechanical properties of Fly- Rovnaník, P., 2010. Effect of curing temperature on the development of hard struc-
-Ash-based geopolymer concrete’. ACI Mater. J. 108 (3), 300–306. ture of metakaolin-based geopolymer. Constr. Building Mater. 24 (7), 1176–1183.
Duxson, P., Fernández-Jiménez, A., Provis, J.L., Lukey, G.C., Palomo, A., Deventer, J.S.J., Ryu, G.S., Lee, Y.B., Koh, K.T., Chung, Y.S., 2013. The mechanical properties of fly
2006. ‘Geopolymer technology: the current state of the art’. J. Mater. Sci. 42 (9), ash-based geopolymer concrete with alkaline activators. Constr. Building Mater.
2917–2933. 47, 409–418.
Duxson, P., Provis, J.L., Lukey, G.C., Van Deventer, J.S., 2007. ‘The role of inorganic Shah, S.P., Karaguler, M.E., Sarigaphuti, M., 1992. Effects of shrinkage-reducing ad-
polymer technology in the development of ‘green concrete’’. Cement Concrete mixtures on restrained shrinkage cracking of concrete. ACI Mater. J. 89 (3),
Res. 37 (12), 1590–1597. 289–295.
EN-1992.1.1, 2004. Eurocode 2, Design of Concrete Structures- Part 1-1: General Siddique, R., 2008. Waste Materials and By- Products in Concrete. Springer,
Rules and Rules for Buildings. European Committee for Standardization, Brus- Berlin,Germany.
sels, Belgium. Sofi, M., Van Deventer, J.S.J., Mendis, P.A., Lukey, G.C., 2007. Engineering properties
ERCOWorldwide, 2012. Material Safety Data Sheet-Sodium Hydroxide Solution. of inorganic polymer concretes (IPCs). Cement Concrete Res. 37 (2), 251–257.
ERCO Worldwide, A division of Superior Plus LP, Toronto, Canada. Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., Chindaprasirt, P., 2011. NaOH-ac-
Fernandez-Jimenez, A., García-Lodeiro, I., Palomo, A., 2006. ‘Durability of alkali-ac- tivated ground fly ash geopolymer cured at ambient temperature. Fuel 90 (6),
tivated fly ash cementitious materials’. J. Mater. Sci. 42 (9), 3055–3065. 2118–2124.
Fernandez-Jimenez, A.M., Palomo, A., Lopez-Hombrados, C., 2006. ‘Engineering Talling, B., Brandstetr, J., 1989. Present state and future of alkali-activated slag con-
properties of alkali-activated fly ash concrete’. ACI Mater. J. 103 (2), 106–112. cretes. ACI Special Pub. 114, 1519–1546.
122 K. Neupane / Mechanics of Materials 103 (2016) 110–122

Tempest, B., 2010. Engineering Characterization of Waste Derived Geopolymer Ce- Wagners, 2010. Sustainable Alternatives in Construction: Earth Friendly Concrete
ment Concrete for Structural Applications Ph. D. thesis. The University of North (EFC). Wagners Global, Toowoomba QLD.
Carolina at Charlotte, Ann Arbor, United States. Wallah, S.E., Rangan, B.V., 2006. Low Calcium Fly Ash Based Geopolymer Concrete—
Teychenné, D.C., Franklin, R.E., Erntroy, H.C., Marsh, B.K., 1997. Building Research Es- Long Term Properties. Faculty of Engineering, Curtin University, Perth, Australia
tablishment-Design of Normal Concrete Mixes, 2nd ed. Construction Research Res. Report-GC2.
Communications Ltd., London, England. Yip, C.K., Lukey, G.C., van Deventer, J.S.J., 2005. The coexistence of geopolymeric gel
Toutanji, H., Delatte, N., Aggoun, S., Duval, R., Danson, A., 2004. Effect of supple- and calcium silicate hydrate at the early stage of alkaline activation. Cement
mentary cementitious materials on the compressive strength and durability of Concrete Res. 35 (9), 1688–1697.
short-term cured concrete. Cement Concrete Res. 34 (2), 311–319. Zeobond, 2008. E-Crete Engineering Summary. Zeobond Pty. Ltd., Somerton, Victo-
ria.

You might also like