Vous êtes sur la page 1sur 13

Construction and Building Materials 85 (2015) 7890

Contents lists available at ScienceDirect

Construction and Building Materials


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

Review

Geopolymer concrete: A review of some recent developments


B. Singh , Ishwarya G., M. Gupta, S.K. Bhattacharyya
CSIR-Central Building Research Institute, Roorkee 247667, India

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

 An overview of geopolymer is

Conversion of y ash into geopolymers/concrete.

presented alongwith its processing


parameters.
 The hardened properties and
durability of geopolymer concrete are
discussed.
 The design guidelines for OPC
concrete are applicable to
geopolymer concrete also.
 Geopolymeric building products
developed at CSIR-CBRI are
highlighted.
 Ambient cured single component
geopolymer may enhance its wider
use in the eld.

a r t i c l e

i n f o

Article history:
Received 26 November 2013
Received in revised form 16 February 2015
Accepted 4 March 2015
Available online 31 March 2015
Keywords:
Geopolymer concrete
Activator
Bond strength
Compressive strength
Durability

a b s t r a c t
An overview of advances in geopolymers formed by the alkaline activation of aluminosilicates is presented alongwith opportunities for their use in building construction. The properties of mortars/concrete
made from geopolymeric binders are discussed with respect to fresh and hardened states, interfacial
transition zone between aggregate and geopolymer, bond with steel reinforcing bars and resistance to
elevated temperature. The durability of geopolymer pastes and concrete is highlighted in terms of their
deterioration in various aggressive environments. R&D works carried out on heat and ambient cured
geopolymers at CSIR-CBRI are briey outlined alongwith the product developments. Research ndings
revealed that geopolymer concrete exhibited comparative properties to that of OPC concrete which
has potential to be used in civil engineering applications.
2015 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.
4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
An overview of geopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Constituents effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
C-S-H phase effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Effect of admixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Curing conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Geopolymer mortars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Geopolymer concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Corresponding author.
E-mail address: singhb122000@yahoo.com (B. Singh).
http://dx.doi.org/10.1016/j.conbuildmat.2015.03.036
0950-0618/ 2015 Elsevier Ltd. All rights reserved.

79
79
80
80
80
81
81
81

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

5.

6.
7.

4.1.
Fresh and hardened properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Interfacial transition zone. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Bond between reinforcing bars and geopolymer concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.
Fire behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Durability studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.
Alkali-silica reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Effect of acid attack. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.
Effect of sulphate attack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.4.
Carbonation and permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.5.
Corrosion of steel reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Research and development at CSIR-CBRI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
The concrete industry faces challenges to meet the growing
demand of Portland cement due to limited reserves of limestone,
slow manufacturing growth and increasing carbon taxes. It is
reported that the requirement of cement in India is likely to touch
550 million tonnes by 2020 with a shortfall of 230 million tonnes (58%) and the demand for cement has been constantly
increasing due to increased infra-structural activities of the country [1]. One effort to combat shortfall is the development of alternate binders to Portland cement aiming at to reduce the
environmental impact of construction, use of greater proportion
of waste pozzolan, and also to improve concrete performance.
Search for several alternatives such as alkali-activated cement, calcium sulphoaluminate cement, magnesium oxy carbonate cement
(carbon negative cement), supersulphated cement etc. are being
made with the advantages of Portland cement [2]. As the family
of the alkali-activated cement is growing, the alkaline cement is
classied based on a phase composition of the hydration products:
R-A-S-H (R = Na+ or K+) in the aluminosilicate based systems and RC-A-S-H in the alkali-activated slag or alkaline Portland cements
[3]. In recent years, geopolymer has attracted considerable attention among these binders because of its early compressive
strength, low permeability, good chemical resistance and excellent
re resistance behaviour [49]. Because of these advantageous
properties, the geopolymer is a promising candidate as an alternative to ordinary Portland cement for developing various sustainable products in making building materials, concrete, re
resistant coatings, bre reinforced composites and waste
immobilization solutions for the chemical and nuclear industries.

79

81
83
83
84
84
84
85
85
86
86
86
88
88
88

amorphous and possess sufcient reactive glassy content, low


water demand and be able to release aluminium easily. The alkaline activators such as sodium hydroxide (NaOH), potassium
hydroxide (KOH), sodium silicate (Na2SiO3) and potassium silicate
(K2SiO3) are used to activate aluminosilicate materials. Compared
to NaOH, KOH showed a greater level of alkalinity. But in reality,
it has been found that NaOH possesses greater capacity to liberate
silicate and aluminate monomers [4]. The properties of geopolymers can be optimised by proper selection of raw materials, correct
mix and processing design to suit a particular application [4].
Viewing the importance of the subject, a collaborative project
sponsored by the European Commission BRITE was undertaken
jointly by France, Spain and Italy on development of Cost-effective
geopolymeric cement for innocuous stabilization of toxic elements
(GEOCISTEM). The project was aimed at manufacturing geopolymeric cement by replacing potassium silicate with cheaper alkaline
volcanic tuffs [9].
Geopolymers are synthesized by the reaction of a solid
aluminosilicate powder with alkali hydroxide/alkali silicate [8]. A
schematic representation on formation of y ash-based geopolymers/concrete is shown in Fig. 1. Under highly alkaline conditions,
polymerisation takes place when reactive aluminosilicates are
rapidly dissolved and free [SiO4] and [AlO4] tetrahedral units
are released in solution. The tetrahedral units are alternatively
linked to polymeric precursor by sharing oxygen atom, thus forming polymeric SiOAlO bonds. The following reactions occur during geopolymerisation [7].

Si2 O5 Al2 O2 n H2 O OH ! SiOH4 AlOH4

2
2. An overview of geopolymers
Geopolymer is considered as the third generation cement after
lime and ordinary Portland cement. The term geopolymer is
generically used to describe a amorphous alkali aluminosilicate
which is also commonly used for to as inorganic polymers,
alkali-activated cements, geocements, alkali-bonded ceramics, hydroceramics etc. Despite this variety of nomenclature,
these terms all describe materials synthesized utilising the same
chemistry [4]. It essentially consists of a repeating unit of sialate
monomer (SiOAlO). A variety of aluminosilicate materials
such as kaolinite, feldspar and industrial solid residues such as
y ash, metallurgical slag, mining wastes etc. have been used as
solid raw materials in the geopolymerization technology. The
reactivity of these aluminosilicate sources depends on their chemical make-up, mineralogical composition, morphology, neness and
glassy phase content. The main criteria for developing stable
geopolymer are that the source materials should be highly

This process releases water that is normally consumed during


dissolution. The water, expelled from geopolymer during the reaction provides workability to the mixture during handling. This is in
contrast to the chemical reaction of water in Portland cement mixture during the hydration process. It is reported that the hydration
products of metakaolin/y ash activation are zeolite type: sodium
aluminosilicate hydrate gels with different Si/Al ratio whereas the
major phase produced in slag activation is calcium silicate hydrate
with a low Ca/Si ratio. Though many physical properties of
geopolymers prepared from various aluminosilicate sources may
appear to be similar, their microstructures and chemical properties
vary to a large extent. The metakaolin-based geopolymer has an
advantage that it can be manufactured consistently, with predictable properties both during the preparation and development.
However, its plate-shaped particles lead to rheological problems,
increasing the complexity of processing as well as the water
demand of the system [6]. Contrary to this, the y ash-based

80

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

Fig. 1. Conversion of y ash into geopolymers/concrete.

particles and formation of reaction products (Fig. 2). The reduced


porosity enhanced the strength of geopolymer pastes [13].
Typically, the optimum geopolymer strength was reported with
SiO2/Al2O3 ratio in the range of 3.03.8 and Na2O/Al2O3 ratio of
1 [14,15]. Changes in SiO2/Al2O3 ratio beyond this range have
been found to result in low strength. The setting time of geopolymer pastes increased with increasing SiO2/Al2O3 ratio of the initial
mixture.
2.2. C-S-H phase effect

Fig. 2. Pore size distribution of y ash-based geopolymer pastes at different


activator dosages [13].

geopolymer is generally more durable and stronger than that of


metakaolin-based geopolymer [4]. The slag-based geopolymer is
considered to have high early strength and greater acid resistance
than those of metakaolin and y ash-based systems.
2.1. Constituents effect
The most important factors affecting the properties of geopolymer pastes are: SiO2/Al2O3 ratio, R2O/Al2O3 ratio, SiO2/R2O ratio
(R = Na+ or K+) and liquidsolid ratio. The majority of research concluded that an amorphous structure of geopolymers is preferable
in order to achieve desired mechanical strength [1015]. The
relationship between the compressive strength and SiO2/R2O ratio
showed that an increase in alkali content or decrease in silicate
content increases the compressive strength of geopolymers attributable to the formation of aluminosilicate network structures
[10,11]. Geopolymer activated with NaOH alone with Si/Na of 4/4
or less formed the crystalline zeolite (Na96Al96Sr96O384216H2O)
but at a ratio >4/4, nanosized crystals of another zeolite
(Na6[AlSiO4]64H2O) were formed [12]. The addition of even small
amount of sodium silicate to the NaOH signicantly reduces crystallite formation due to templating function of silicate units. At low
activator dosage (18%), the pores developed in the y ash-based
paste were larger and exhibited wider distributions (19.8
2342 A) whereas at higher activator dosage (30%), the pores were
smaller and showed a narrow distribution (19.81155 A) mainly
due to the pore renement as a result of more dissolution of

The effect of C-S-H phase on the geopolymerization of


aluminosilicates has been studied with a view to know its role in
early age strength [1622]. In metakaoin/slag blends, both C-S-H
phase and aluminosilicate gel (N-A-S-H) co-exist in the paste
[16] as similar to NaOH activated high calcium y ash-based
geopolymer [17] which are responsible for the strength increase.
The little dissolution of calcium occurs in the case of adding natural
calcium silicate minerals at lower alkalinity, resulting in less C-S-H
gel formation and subsequent strength reduction of geopolymer
pastes [18]. In the case of y ash/slag blends, the reaction at
27 C is dominated by the slag activation, whereas the reaction at
60 C is due to combined activation of y ash and slag. The
improvement in compressive strength of pastes with slag addition
is attributed to its compactness of the microstructure [19]. The
initiation of hardening in y ash/slag geopolymer made with
potassium silicate and potassium hydroxide was due to C-S-H/CA-S-H formation and the hardening continues due to a rapid formation of a C-A-S-H, K-A-S-H and (Ca, K)-A-S-H depending on
the availability of calcium ions and pH of the system. A slower dissolution rate of calcium ions effectively increased the compressive
strength as rapid geopolymerization continues for a longer duration [20]. The low pH and limited calcium ion environment facilitate the polymerisation reaction between silicate and aluminate
species in high calcium y ash-based geopolymers producing NA-S-H gel [21]. Guo et al. [22] reported 63.4 MPa compressive
strength of class C y ash-based geopolymer paste showing the
role of calcium participation in the strength development.
2.3. Effect of admixtures
Kusbiantora et al. [23] reported from their studies that admixtures such as sucrose and citric acid which act as retarder in OPC
have different mechanism in y ash-based geopolymers. Sucrose
acted as a retarder since it is absorbed by Ca, Al and Fe ions to form
insoluble metal complexes. On the other hand, citric acid acted as
an accelerator reducing the setting time by 9 and 16 min

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

respectively. Amongst the commercial superplasticizers, the naphthalene based superplasticizer was effective when single activator
was used rendering 136% increase in relative slump without any
decrease in compressive strength. Modied polycarboxylate based
superplasticizer was efcient one when multi-compound activator
was used with a decrease in compressive strength of 29% [24].
However, retarding effect of polycarboxylate based superplasticizer was also reported in y ash/slag blended system though
the improvement in workability was signicant compared to naphthalene based superplasticizer [25].
2.4. Curing conditions
Several attempts [2631] have been made to study the effect of
different curing conditions on the properties of geopolymer pastes.
The curing temperatures were reported in the range between 40 C
and 85 C for complete geoplymerisation reactions. Palomo et al.
[26] studied curing of alkali activated y ash (0.25 and 0.30 liquid/solid ratio) at 65 C and 85 C. They indicated that the compressive strength of geopolymers (812 M) cured at 85 C for
24 h was much higher than those cured at 65 C. The rise of
strength was much smaller when curing time was extended after
24 h. Perera et al. [27] studied the curing of metakaolin-based
geopolymers under ambient (2123 C) and heat conditions (40
60 C) with a controlled relative humidity (RH) for 24 h and found
that curing at 30% RH was preferable to that at 70% RH. Heah et al.
[28] concluded that the curing of metakaolin-based geopolymers
at ambient temperature was not feasible while increase in temperature (40 C, 60 C, 80 C, 100 C) favored the strength gain after
13 days. However, curing at higher temperature for a longer period of time caused failure of samples at a later age due to the thermolysis of SiOAlO bond. Rovnanik [29] reported that curing
of metakaolin based geopolymer at elevated temperature (40
80 C) accelerated the strength development but in 28 days, the
mechanical properties deteriorated in comparison with results
obtained for an ambient or slightly decreased temperature.
Ebrahim and Ali [30] prepared three mixes with different formulations and cured hydrothermally at different temperatures
(45, 65, 85 C) and time (520 h) after 1 and 7 days of procuring.
Longer procuring at room temperature, before the application of
heat is benecial for higher strength development. In general, adequate curing of geopolymeric materials is required to achieve optimal mechanical and durability performance to maintain their
structural integrity [31].

81

(14 M activator solution) with 1030 wt% aggregate exhibited an


acceptable owability, while the mortars containing 40 & 50 wt%
aggregate were stiff and difcult to pack in the mould. Increasing
aggregate content in the mortar mixes leads to insufcient activator for complete geopolymerization of y ash/slag. The activator
may also be utilised for wetting of aggregate leaving less availability for dissolution of these y ash or slag particles. The compressive
strength of geopolymer mortars with high level of aggregate can be
achieved by optimising the amount of activator dosage [34].
Khandelwal et al. [35] summarised that the compressive strength,
modulus of elasticity and Poissons ratio of y ash-based geopolymer mortars increased logarithmically with the increase of strain
rate. These engineering properties of geopolymer mortars compared favourably with those predicted by Standards/Codes for concrete mixtures. When bottom ash was used, the geopolymer
mortars exhibited a low compressive strength (20 MPa). With
10% replacement of sand by bottom ash, the mix exhibited a comparable compressive strength to those made with sand only. The
increase in strength (50100%) of bottom ash mortar was also
reported when the specimens were exposed at 800 C probably
due to activation of bottom ash [36]. When lignite bottom ash
was ground to a mean particle size of 15.7 lm (3% retained on
sieve No. 325), the compressive strength of mortars activated with
sodium hydroxide/sodium silicate was 2458 MPa [37].
Brough and Atkinson [38] prepared geopolymer mortars using
slag, sand and activator in a ratio of 1:2.33:0.5. At water-to-total
solid ratio of 0.42, the mortar gained strength of 40 MPa. The
sodium silicate activated mortars exhibited higher compressive
strength with low levels of porosity at the interface while KOH
activated mortars were highly porous in the interfacial zone giving
low compressive strength values. Yang et al. [39,40] found that the
ow of alkali-activated mortars increased with the increase of
water-binder ratio and decrease of aggregate-binder ratio. When
the aggregate-binder ratio was larger than 2.5, the ow of mortars
decreased sharply. They also found that slag-based geopolymer
mortars exhibited much higher compressive strength but exhibited
slightly less ow than the y ash-based geopolymer mortars for
the same mixing condition. The poor compressive strength of y
ash-based mortars cured at low temperatures is attributed to the
presence of unreacted y ash particles and large number of voids.
As the aggregate-binder ratio increased, the compressive strength
increased up to a ratio of 2.5 which indicated that the threshold
of aggregates in geopolymer mortars were slightly lower than
OPC mortars. The shrinkage strain of alkali-activated mortars was
also found to be lower than the OPC mortars.

3. Geopolymer mortars
4. Geopolymer concrete
Various studies [3240] were conducted on ow and mechanical properties of geopolymer mortars because of their more relevant applications in building construction. The properties of
mortars were optimised with respect to initial ow, aggregate-binder ratio, activator-binder ratio and activator molarity.
Chindaprasirt et al. [32] reported that the compressive strength
of class C y ash-based geopolymer mortar was 52 MPa when
cured at 70 C for 3 days using sand-y ash ratio of 2.75 at workable ow of 135 5%. Prolonged curing at high temperatures led
to the reduction in the compressive strength because of weakening
of microstructure and increased porosity due to the loss of moisture. In another attempt [33], they produced geopolymer mortar
with a compressive strength of 86 MPa at 28 days with the help
of air classied class C y ash (4500 cm2/g neness) activated with
sodium silicate and NaOH (10 M) at 1:1 mass ratio. The dimensional change in terms of drying shrinkage (161  106 mm/
mm) was insignicant when compared with the Portland cement
mortar (700850  106 mm/mm). The geopolymer mortars

4.1. Fresh and hardened properties


Various mix proportioning of geopolymer concrete (GPC) were
reported with target strength up to 80 MPa. The typical properties
of geopolymer concrete mixes used by the various authors (41, 45,
48, 49, 53) are summarised in Table 1. The properties of mixes
were studied with respect to water-geopolymer solid ratio, activator strength, water/Na2O ratio, curing time, curing temperature,
and age hardening. The slump of mixes varied depending on the
molarity of activator, workability aids and extra water added to
the mix [41]. The rheological parameters such as yield stress and
plastic viscosity were attempted over slump test of concrete to
assess its workability loss and ow behaviour. Yield stress gives
initial resistance to ow arose from the friction among the solid
particles while plastic viscosity governs the ow after it is initiated
resulting from viscous dissipation due to the movement of water in
the sheared material. Laskar & Bhattacharjee [42] studied the

82

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

Table 1
Typical properties of geopolymer concrete mixes.

Hardjito et al. [41]


Jimenez et al. [45]
So et al. [48]
Diaz-loya et al. [49]
Pan et al. [53]

Density
(kg/m3)

Molarity
(M)

Slump
(mm)

CS
(MPa)

STS
(MPa)

FS
(MPa)

MOE
(GPa)

Poissons
ratio

Activator/
binder
ratio

Curing
temperature
and time

23302430
NR
21472408
18902371
18762555

1016
8 & 12.5
NR
14
8

60215
NR
NR
100150
NR

3080
2943.5
4756.5
1080
65.177.9

3.746
NR
2.84.1
NR
2.85.1

512
6.86
4.96.2
2.246.41
NR

2331
10.718.4
2339
1.942
11.241.2

0.120.16
NR
0.230.26
0.080.22
0.150.19

0.350.4
0.4 & 0.55
0.450.59
0.40.94
0.40.65

6080 C for 24 h
85 C for 20 h
23 C till testing
60 C for 72 h
60 C for 24 h

CS: compressive strength; STS: splitting tensile strength; FS: exural strength; MOE: modulus of elasticity; NR: not reported.

rheology of y ash-based geopolymer concrete with slump varying


from 25 mm to owing concrete with 120 M activator strength.
They found that the yield stress and plastic viscosity were affected
by the molar strength of the sodium hydroxide solution and the
ratio of silicate to hydroxide solution. The setting time of geopolymer concrete was reported up to 120 min. Like Portland cement
pastes and mortars, geopolymers also behave like Bingham uid
and have a history dependent rheological prole, i.e., geopolymers
may be kept in a uid form, if subjected to constant shearing for a
certain period of time before initial setting starts [43]. The setting
could be enhanced up to 180 min with the use of naphthalene
based admixture and extended mixing time especially in the case
of slag-based geopolymer which has the potential for a wide range
of technological applications [44].
Hardjito et al. [41] produced y ash-based GPC with the compressive strength ranging between 30 and 80 MPa with the slump

Fig. 3. Correlations within the mechanical properties of y ash-based geopolymer


concrete. (a) Splitting tensile strength vs compressive strength. (b) Modulus of
elasticity vs compressive strength [47].

varied from 100 to 250 mm (activator strength: 814 M). The optimum strength was obtained at 0.18 water-geopolymer solid ratio
cured at 90 C. As the water-geopolymer solids ratio increased,
the compressive strength of GPC decreased analogous to the well
known relationship between compressive strength and water-cement ratio for OPC concrete. The compressive strength of GPC
remained unchanged with the age when tested after 24 h curing
at elevated temperature. Fernandez-Jimenez et al. [45] made y
ash-based geopolymer concrete with a compressive strength of
45 MPa at 0.55 liquid/solid ratio cured at 85 C for 20 h. The development of high early strength in GPC was explained by its compact
microstructure, formation of adequate reaction products, smaller
mean size of the pores and good aggregate-paste bond. They
observed that GPC has a much lower modulus of elasticity
(18.4 GPa) than the OPC concrete (30.3 GPa). Olivia and Nikraz
[46] proportioned y ash-based geopolymer concrete mix with a
compressive strength of 55 MPa at 28 days and cured at different
temperatures in the range of 6075 C. The hardened mix had
higher tensile and exural strengths, produced less expansion
and showed modulus of elasticity that were 1529% lower than
that of OPC concrete mix. The drying shrinkage (0.025%) of GPC
was less than the OPC concrete (0.09%) after 12 weeks. The minimal shrinkage of GPC may also be due to the signicant resistance
offered by its zeolitic microstructure towards drying loss of the
water incorporated during casting [45].
Several attempts [4753] have also been made to establish
correlations within the mechanical properties of geopolymer concrete. It was reported that the experimental splitting tensile
strength of y ash-based GPC was higher than the OPC concrete
(Fig. 3). The increased strength is accounted for a denser interfacial
zone established between the aggregate and geopolymer paste.
The modulus of elasticity increased as the compressive strength
of GPC increased. The modulus of elasticity of GPC was found to
be lower than the values predicted by ACI guidelines for OPC concrete. So et al. [48] studied the engineering properties of y ash/
slag-based GPC. The splitting tensile strength and exural strength
of GPC were comparable to those models presented by the
Australian Standard (AS 3600) for OPC concrete. Although, the difference between splitting tensile and exural strength of GPC
mixes has been found to be approximately 2 MPa, similarities
between the strength gain was apparent. Diaz-Loya et al. [49] prop
posed the equation fr = 0.69 fc0 MPa for correlation between the
exural strength (fr) and compressive strength and the equation
p
Ec = 580 fc MPa for correlation between elastic modulus (Ec)
and compressive strength of GPC (fc = compressive strength).
When compared with the typical Poissons ratio value of OPC concrete (0.150.22), the values of GPC appeared to reside toward the
low end of range (0.080.22). Ryu et al. [50] suggested a model for
relationship between compressive strength and splitting tensile
strength (fsp = 0.17 (f0 c)3/4) for y ash-based GPC. Bondar et al.
[51] reported a relationship between ultrasonic pulse velocity
and compressive strength of GPC. They found that GPC showed a
lower ultrasonic pulse velocity than the OPC concrete even those

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

with the same or higher compressive strength. It was also reported


[52,53] that GPC was brittle as compared to its OPC counterpart
due to the highly cross-linked framework. The fracture energy of
GPC was also low because of its higher bond with aggregates as
compared to OPC concrete [54].
4.2. Interfacial transition zone
It is well known that the interfacial zone (ITZ) between aggregate and matrix is the weakest link in OPC concrete at which
micro-cracks usually rst develop under loads [55]. Investigation
of this zone is very crucial since it is known to have different
microstructure from the bulk of the hardened paste. The high
porosity of ITZ allows the easier penetration of external agents
such as chlorides, oxygen, sulphates, etc. into concrete structure.
Contrary to this, ITZ of GPC has been identied as being dense
and much less microstructurally distinct from the bulk of binder
region [56,57]. The stronger ITZ contributes to higher splitting tensile strength, bond strength and durability of the GPC.
Lee and Deventer [56,57] discussed interface between the natural siliceous aggregates and paste in GPC using kaolin and albite as
precursors. The increase in concentration of the activating solution
increased the binding capacity of the gel with natural aggregates.
The presence of chloride salts decreased the interfacial bonding
strength between the paste and aggregate probably by causing
gel crystallisation near the aggregate surfaces which resulted in
debonding. In another attempt, they found that the addition of
0.5 M soluble silicate into an activating solution (10 M NaOH and
2.5 M sodium silicate) facilitates the formation of an aluminiumenriched aluminosilicate surface onto the aggregates through
accelerated Si-preferential dissolution of kaolin and albite. The surface formed during albite leaching was found to possess a similar
Si/Al ratio to the real interface between a silicious aggregate and
y ash/metakaolin geopolymer paste activated with 10 M NaOH
solution. Without soluble silicates, no deposited aluminosilicate
interface was observed. This suggested that both high concentration of alkali and soluble silicate are essential for the formation
of a strong interface between silicious aggregates and geopolymer
pastes. Zhang et al. [58] reported that at the beginning, there were
many large voids in the fresh ITZ in potassium poly(sialate)
geopolymer concrete. As hydration proceeded, these voids were
completely lled with the hydration products. At this stage, the
difference in the microstructure between the ITZ and matrix was
hardly distinguishable. The contents of K/Al and Si/Al in the ITZ
were higher than those in the matrix. Demei et al. [59] presented

Fig. 4. Bond strength of y ash-based geopolymer concrete as a function of steel bar


diameter [47].

83

FESEM analysis of ITZ in the y ash-based self compacting geopolymer concrete with varying superplasticizer dosages. They reported
that relatively a loose and porous ITZ was found at low superplasticizer dosages (3%) whereas a dense ITZ was found between
the aggregate and geopolymer paste at higher dosage (7%). They
also found that the compressive strength increased with decrease
in the thickness of ITZ and this relationship depends on the superplasticizer dosage.
4.3. Bond between reinforcing bars and geopolymer concrete
The transfer of forces across the interface between concrete and
reinforcing steel bar is of fundamental importance in the structural
design [60]. Bond stresses in the reinforced concrete arise from two
distinct situations. The rst is anchorage or development where
bars are terminated. The second is exural bond or the change of
force along a bar due to a change in bending moment along the
member. The bond strength of reinforcing bars with concrete is
governed by several factors such as the strength of the concrete,
the thickness of the concrete surrounding the reinforcing bar, the
connement of the concrete due to transverse reinforcement and
the bar geometry. Generally, the bond strength between the
reinforcing bar and matrix increases with increasing steel bar
diameter and compressive strength of GPC (Fig. 4). There is a
greater amount of slip for larger size rebars in GPC.
Sarker [61] found that the bond strength of y ash-based GPC
increased with the increase of concrete cover-bar diameter ratio
(1.713.62) and the concrete compressive strength (2529 MPa).
He also observed that GPC has higher bond strength than the
OPC concrete because of higher splitting tensile strength and dense
interfacial transition zone between the aggregate and geopolymer
paste. Bond-slip behaviour [45] of GPC showed that the embedded
steel bar of 8 mm dia broke before slipping and concrete cracking
whereas the bar embedded in OPC concrete slipped. For
16 mm bar, GPC failed by matrix cracking while the bars in OPC
concrete were again observed to slip. So et al. [62] reported that
the values of bond strength of steel bars in y ash-based GPC were
comparable in both beam-end as well as direct pullout specimen
tests. The normalised bond strength increased with a reduction
in rebar size. The bond strength tested according to AS 3600, ACI
318-02 and EC2 recommendations showed that these Codes are
applicable and also safe to predict the developmental length for
GPC.
Attempts were also made to study behaviour of reinforced y
ash-based GPC beams and columns with respect to longitudinal
tensile reinforcement ratio and concrete compressive strength as
test variables [6366]. Sumajouw et al. [63] reported that the
exural capacity of beams increased with the increase in tensile
reinforcement (0.642.69%) but the effect of concrete compressive
strength was marginal. The ductility index increased signicantly
for beams having longitudinal reinforcement ratio less than 2%.
They also studied the strength of reinforced GPC slender columns
with respect to the compressive strength of concrete, longitudinal
reinforcement ratio and load eccentricity. The design provisions
mentioned in the Standards for OPC concrete can be used for
designing geopolymer concrete columns also. Dattatreya et al.
[64] found that the load carrying capacity of reinforced slag-based
GPC beams was 17.7% more than the Portland pozzolana cement
concrete beams at 2.68% tension reinforcement. Yost et al. [65]
indicated that loaddeection behaviour of GPC beam was identical to OPC beam. The maximum strain obtained for under-reinforced beam was less than 3000 microstrains which is generally
assumed for design work. The predicted neutral axis depth was
15% less than the experimentally achieved value for GPC. The
Whitneys stress block for strength calculation was found applicable for GPC also. Ng et al. [66] investigated potential use of steel

84

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

bres (up to 1.5 wt%) to replace conventional shear reinforcement


in GPC beams of 2250 mm span length. They found that the
increase in bre volume led to an increase in the cracking load
and the ultimate shear strength. A good correlation of test data
was observed with the predictive b Model Code 2010.

carbonation, alkali-silica reaction and freezethaw attack. In view of


this, several studies are being carried out to understand the behaviour of geopolymers exposed to these conditions.

4.4. Fire behaviour

Alkali-silica reaction (ASR) causes gradual but severe deterioration of hardened Portland cement concrete in terms of its
strength loss, cracking, volume expansion etc. It involves the reaction between the hydroxyl ion in the pore solution within the concrete matrix and reactive silica of the aggregate. In general terms,
the reactions will proceed in stages, with the rst stage being the
hydrolysis of reactive silica by hydroxyl ions to form alkali-silica
gel and a later secondary overlapping stage being the absorption
of water by the gel, which will result in increase of volume [72].

In general, concrete has good property with respect to re resistance. However, it is known that the residual strength of OPC concrete after ring between 800 C and 1000 C does not exceed 20
30% normally because of dehydration and destruction of C-S-H &
other crystalline hydrates, aggregate types, permeability etc. Fire
introduces high temperature gradient and as a result, the hot layer
tends to separate and spall from the cooler interior layer of the
body [67]. Contrary to this, geopolymers possess good re resistance at elevated temperature because of the existence of highly
distributed nano-pores in the ceramic like microstructure that
allows physically and chemically bonded water to migrate and
evaporate without damaging the aluminosilicate network [4].
During re, several events such as evaporation of water adsorbed
by N-A-S-H gel, formation of anhydrous products, crystallization
of stable anhydrous phases and melting (sintering) leading to
destruction generally occurred. The phase transformation of
geopolymers during re is depicted below.

Kong et al. [68] found that the residual strength of y ash-based


geopolymer pastes increased by 6% after exposure to 800 C,
whereas the strength of metakaolin-based geopolymer pastes was
reduced by 34%. During heating, the high permeability of y ashbased geopolymer provides the escape route for moisture in the
matrix, thereby decreasing the damage. The strength increase is
also partly attributed to the sintering reaction of unreacted y ash
particles. Geopolymer pastes made with metakaolin and potassium
based activator showed an enhanced post-elevated temperature
performance compared to sodium based activator system. The
strength deterioration reduced with increasing Si/Al ratio (>1.5)
[69]. Aggregate size larger than 10 mm resulted in good strength
performance in both ambient and elevated temperature (800 C).
The strength loss in y ash-based geopolymer concrete at elevated
temperatures is attributed to thermal mismatch between the
geopolymer paste and aggregate [70]. No spalling was reported in
the samples by Zhao and Sanjayan [71] when y ash-based GPC
with compressive strength ranging from 40 to 100 MPa was
exposed to 850 C. They also found that at the same strength level,
GPC possessed higher spalling resistance under re than the OPC
concrete due to its increased porosity.

5.1. Alkali-silica reaction

(i) Acid-based reaction

H0:38 SiO2:19 0:38NaOH ! Na0:38 SiO2:19 0:38H2 O

(ii) Attack of the siloxane bridges and disintegration of the silica


2

Na0:38 SiO2:19 1:62NaOH ! 2Na2 H2 SiO4

In geopolymer concrete, the un-utilised alkali after geopolymerization of aluminosilicates is expected to react with the silica of the
aggregates causing disruption of their siloxane bridges. It is
reported that geopolymer mortars using aggregates of different
reactivities expanded less than the corresponding Portland cement
mortars [73]. The geopolymer mortars appeared to be sound without any surface cracking. The cause of expansion in slag-based
geopolymer mortars is the formation of sodium calcium silicate
hydrate reaction product with rosette-type morphology [74].
Contrary to this, there was no signicant expansion in y ashbased geopolymer mortars. The formation of crystalline zeolites
was very slow and since these minerals are usually found in the
gaps of the matrix, the existence of stress that might generate
cracking is unlikely [75]. Geopolymer mortar bars made with y
ash/slag blends expanded less than 0.1% limit prescribed in ASTM
C1260-07 after 16 days (Fig. 5). At 90 days exposure, these mortars
failed to meet the specied criteria. Increasing slag content in y
ash/slag mix increased the expansion of resulting systems [76].
ASR has also been claimed to be helpful in providing a strong bond
at the paste-aggregate interface, thus enhancing the tensile
strength of GPC [8]. Patil et al. [73] indicated that sandstone, quartz
and limestone aggregates in geopolymer concrete were not prone

5. Durability studies
One of the major problems associated with OPC concrete is its
long term durability which had always been an issue against
aggressive environments. The deterioration of concrete is usually
assessed for sulphate attack, chloride induced corrosion, atmospheric

Fig. 5. Alkali-silica reaction in various geopolymer and OPC mortars under an


accelerated condition (1 M NaOH) at 80 C [76].

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

85

to ASR. During accelerated mortar bar test, a slight expansion was


noticed because of re-initiation of the geopolymerization process
of unreacted y ash particles leading to lower porosity and higher
strength. The lower sensitivity of reactive aggregates in GPC provides economic advantages in areas where high quality deposits
of aggregates have been depleted.

5.2. Effect of acid attack


The acid resistance of geopolymer pastes/concrete was studied
by several authors [7784]. The extent of degradation depends on
the concentration of acid solution and period of exposure.
Davidovits et al. [8] indicated that metakaolin-based geopolymer
pastes showed only 7% mass loss when sample was immersed in
5% H2SO4 for 30 days. It was also reported that y ash-based
geopolymer pastes retained a dense microstructure after 3 months
exposure in HNO3. Temuujin et al. [77] concluded that acid and
alkaline resistance of y ash-based geopolymer strongly depend
on its mineralogical composition. High solubility of Al, Si and Fe
ions was obtained in both strong alkali and acid solutions. The performance of y ash-based geopolymer pastes when exposed to 5%
acetic acid and 5% H2SO4 solutions was superior to ordinary
Portland cement pastes. The deterioration in pastes was connected
to depolymerisation of the aluminosilicate network and formation
of zeolites [78].

5
Wallah and Rangan [41] found that the reduction in compressive strength of y ash-based GPC in 0.5% H2SO4 solution was
20% after 12 months exposure. This value was 52% and 65%
respectively when samples exposed to 1% and 2% H2SO4 solution.
Pitting and erosion on the surface of the concrete were also
observed. The loss in strength of concrete is mainly due to the
degradation in the geopolymer matrix rather than the aggregate.
They concluded that the acid resistance of GPC was superior to
OPC concrete. Arifn et al. [79] exposed GPC made with a blend
of pulverized fuel ash and palm oil fuel ash in 2% solution of sulphuric acid for 18 months. The weight loss in GPC was 8% while
OPC concrete exhibited 20% weight loss. The strength reduction
in GPC was 35% in 18 months as against 68% strength loss in OPC
concrete after 30 days and was severely deteriorated after
18 months. The C-S-H could have severe deleterious effect on
OPC concrete while N-A-S-H gel appeared to have little effect on
the structure of GPC. Sathia et al. [80] reported the weight loss in
concrete samples was less than 5% after 3 months exposure in 3%
H2SO4 solution. Bakharev et al. [81] found that slag-based GPC
(40 MPa) exhibited 33% reduction in strength compared to 47%
in OPC concrete when exposed in acetic acid solution (pH 4) for
12 months. The slag particles and low calcium C-S-H with average
Ca/Si ratio of 1 were more stable in the acid solution than the constituents of the OPC pastes. During immersion in 2% H2SO4 solution, the strength loss was 11% compared to 36.2% for OPC
concrete.

Fig. 6. Atomic force microscope images of y ash-based geopolymer exposed under


sulphate after 4 months [86].

5.3. Effect of sulphate attack


Fly ash-based geopolymer pastes did not deteriorate signicantly, under the inuence of water, sodium sulphate (4.4%)
and ASTM sea water [82]. Only some uctuations in exural
strength were observed between 7 days and 3 months exposures.
The least strength change was observed in the pastes exposed in
the 5% Na2SO4 and 5% MgSO4 solutions while most signicant
deterioration was observed in the 5% mixed sulphate solution

86

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

(Na2SO4 + MgSO4) after 5 months exposure [83]. In y ash/slag


system, the extensive physical deterioration of pastes was
observed during immersion in MgSO4 solution after 3 months
exposure but not in Na2SO4 solution. The calcium sulphate dihydrate formed in paste was identied as being particularly damaging to the materials in MgSO4 [84]. Atomic force microscopic
images of y ash-based geopolymer pastes exposed to sulphate
environment are shown in Fig. 6. In the case of Na2SO4 solution,
only exposition of grains was clearly visible while in MgSO4 solution, both exposition of grains and dissolved aluminosilicate
matrix were observed showing severity of MgSO4 attack [85].
The deterioration is considered mainly due to the destruction of
aluminosilicate skeleton, liberation of silicic acid, leaching of
sodium ion etc. [86]. These reactions seem to have signicant
effect on the mechanical strength. The geopolymer prepared with
NaOH activator had the best performance over those made with a
synergistically used sodium silicate and NaOH/KOH activators,
which is attributed to its stable cross-linked aluminosilicate polymer structure.
Several attempts [41,87] have been made to study sulphate
resistance of GPC. The deterioration in concrete was evaluated in
terms of its visual appearance, weight loss and change in compressive strength. Hardjito et al. [41] observed that there was no signicant effect of 5% Na2SO4 solution in the compressive strength,
the weight loss and the dimension of y ash-based GPC after
3 months exposure. Rajamane et al. [87] reported sulphate resistance of y ash-based GPC for 3 months in 5% Na2SO4 and 5%
MgSO4 solutions. The weight loss in samples was 2.4% only.
There was 229% loss of compressive strength as compared to 9
38% in the OPC concrete. The deterioration of OPC concrete can
be attributed to the formation of expansive gypsum and ettringite
which can cause expansion, cracking and spalling in the concrete.
Contrary to this, GPC in general do not contain Ca(OH)2 and monosulphoaluminate in the matrix to cause expansion.

5.4. Carbonation and permeability


Bernal et al. [88] studied slag/metakaolin-based GPC (w/b ratio
0.47) under an accelerated carbonation test using CO2 concentration of 3.0 0.2% at 20 C for 28 days. They found that the compressive strength decreased monotonically as the carbonation
proceeds. The relationship between the pore volume and extent
of carbonation was much more similar with samples with different percentages of metakaolin contrary to the slag-based samples.
This suggested that porosity is not the only parameter controlling
the strength loss of the carbonated binder. There must be a
convoluting effect due to the binder gel chemistry, which determines the residual level of strength after an accelerated carbonation. Olivia and Nikraz [46] reported lower water permeability
(2.464.67  1011 m/s) of GPC (activator-y ash ratio, 0.300.40
cured at 60 C for 24 h) than the OPC concrete due to its denser
paste and smaller pore inter-connectivity. They also reported that
the water-geopolymer solids ratio was the most inuential
parameter that affects the properties of GPC. Bondar et al. [51]
studied the oxygen and chloride permeability of alkali-activated
concrete made with the Iranian natural pozzolan (Taftan andesite
and Shahindej dacite). They concluded that alkali-activated natural pozzalona concrete has 1035% lower oxygen permeability at
normal curing conditions for 90 days compared with the OPC concrete. The rapid chloride permeability test gave high values for
the alkali-activated concrete. This is probably due to the very
high alkali ion concentration in the pore solution promoting
higher electrical conductivity in the GPC. This effect seems to
reduce with age due to a change in the porosity of the GPC
microstructure.

5.5. Corrosion of steel reinforcement


Corrosion potential is a technique used to detect the state of
reinforcement without disturbing the structures. This is important
because the intensity of corrosion of steel in concrete is generally
known only after the concrete has cracked or disrupted. Various
studies [46,80,89] were reported to estimate the corrosion potential of steel within the GPC as per ASTM C876. Olivia and Nikraz
[46] reported that the half cell potential of GPC was lower than
the specied value of 404 mV mentioned in the Standard for severe corrosion after 91 days. Sathia et al. [80] also reported corrosion potential up to 300 mV which showed a probable
corrosion indication due to the lower pH of concrete during the
half-cell potential measurement. Accelerated corrosion results
showed that GPC mixes exhibited low level corrosion activity
and time to failure that were 3.865.70 times longer than those
of the OPC concrete. Under impressed voltage, a crack appeared
suddenly in the concrete when time to failure was reached and this
was followed immediately by high current reading. The large
amounts of y ash and alkaline activators in the GPC mix increased
the availability of ions that can produce high electrical resistance at
high impressed voltage. This enhanced the cathodic reaction and
reduces the rate of corrosion, which in turn, reduces the tensile
stress of the specimens, thus decreasing the risk of cracking and
clearly extending the time to failure [46]. Reddy et al. [89] compared the durability of GPC with that of OPC concrete exposed to
marine environment for a period of 21 days. The initial corrosion
current measured for GPC (7191 mA) was much lower than that
of OPC concrete (772 mA). The OPC specimens initially recorded
decrease in the current but later started increasing while the GPC
current never showed signicant increase.

6. Research and development at CSIR-CBRI


A systematic R&D work is initiated at CSIR-Central Building
Research Institute, Roorkee on the development of heat and ambient cured geopolymers using y ash, slag and other aluminosilicates as precursors. In view of variability in the constituents of
y ash, the property optimisation of geopolymeric pastes was carried out as a function of activator concentration and its dosage,
water-geopolymer solid ratio, curing time and curing temperature
[13]. Geopolymerisation reaction, thermal stability, identication
of bond linkages and microstructural features were analysed by
various techniques such as quasi isothermal DSC, TGA, FTIR and
FESEM. The durability of geopolymer pastes/mortars was also
studied in terms of alkali-silica reaction and also in acidic and sulphate environments for 4 months [76,86]. The suitability of these
geopolymer pastes was assessed in making various geopolymeric
products such as mortars & concrete, bricks, solid & hollow blocks,
insulation concrete, foam, sandwich composites and temperature
resistant coatings (Fig. 7(ac)). Attempt was also made to utilise
lime sludge, a waste from paper industry with the geopolymeric
binders for making paving blocks.
Fly ash-based GPC mixes were made with the compressive
strength of 2555 MPa using absolute volume method adopted
for OPC concrete mixes. The strength of GPC increased with
decreasing water-geopolymer solid ratio as it is said analogous to
the water-cement ratio of the OPC concrete. The compressive
strength increased with increasing molarity of the activator (10
16 M) probably due to the formation of stable aluminosilicate networks following the dissolution of silica and alumina in the solution from the y ash. It was found that the splitting tensile
strength of GPC was more than those of predicted values as per
ACI 318 guideline and other existing empirical equations. A trend
line curve between the compressive strength and modulus of

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

87

Fig. 7a. Light weight y ash-based geopolymer concrete sheets using EPS beads and in-situ foaming [85].

Fig. 7b. Fly ash-based geopolymer bricks [85].

Fig. 7c. Fly ash-based geopolymer solid and hollow blocks [85].

elasticity showed that the elastic modulus was lower (17%) than
the one predicted by Ivan Diaz-Loya et al. for GPC and also the values obtained with ACI guidelines. As expected, the bond strength of
steel bar embedded in GPC increased with increasing steel bar
diameter and compressive strength of concrete. It was noted that
the bond strength between geopolymer paste and reinforcing bars
was found to be higher than the OPC concrete [47].
Light weight geopolymer concrete was proportioned with the
help of y ash, activators, expanded polystyrene beads (EPS up
to 3 wt% or 91 vol%), admixtures and ne & coarse aggregates
(Fig. 7(a)). It was noted that a decrease in the strength was more

when larger size of EPS beads (<4.75 mm) were added in the mortars probably due to their less surface area/volume ratio. By adding
coarse aggregate, the compressive strength (18 MPa) and density
(1500 and 1840 kg/m3) of EPS geopolymer concrete, comply the
minimum specied criteria of ACI 213R-03 guidelines for structural light weight concrete (compressive strength, 17 MPa; density
11201920 kg/m3). To meet the requirement of insulation concrete, the addition of 20% coarse aggregate (10 mm maximum size)
into EPS/geopolymer mix exceeds its compressive strength

88

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

Fig. 8. Fire test of y ash-based geopolymers as per BS 476 showing surface of very low spread of ame [90].

(15 MPa) as specied (13.1 MPa) in ASTM C 90. Regarding re


performance, the samples were non-ignitable and exhibited Class
I-very low spread of ame as per BS EN-476 part 7 (Fig. 8). It
was noted that the re propagation index of the samples was <3
exhibiting no support to re growth. Flammability data obtained
from a cone calorimeter showed that the insulation concrete had
insignicant heat release rate (9.63 kW/m2) and effective heat of
combustion (3.75 MJ/kg). The thermal conductivity of insulation
concrete was found in the range of 0.4270.852 W/mK. It was concluded that light weight concrete can be engineered by proper
selection of variables in making insulating materials for use in
buildings [90].
A geopolymer foam composition has been developed using y
ash, activator, ller, surfactant, buffer and strengthening agent. It
sets at room temperature within 2 h and completely cured after
24 h. The density of foam was lying in the range of 600800 kg/
m3. It can be easily prepared by a simple mixing followed by pouring into mould. SEM examination revealed that pores in the samples were uniformly distributed. The ammability test carried
out by a cone calorimeter showed that the total heat released, mass
loss, smoke release and CO/CO2 yield were insignicant. It resists
against re between 700 C and 800 C. The foam can be used as
a core material in the sandwich and insulation panels [85].
Geopolymer bricks of size 230  115  75 mm were produced
using y ash-based pastes, coarse y ash and natural sand
(Fig. 7(b)). The bricks were cured at 80 C for 2 h. The bricks were
obtained with density ranged between 1920 and 2100 kg/m3,
water absorption, 1015% and dry compressive strength, 12
25 MPa depending on activator concentration. These bricks can
be easily jointed with ordinary cement mortars [85].
The solid geopolymer blocks of size 300  200  150 mm were
produced on a machine using coarse aggregate (1180 kg/m3), ne
aggregate (296 kg/m3), yash (494 kg/m3), activator (111 kg/m3)
and water (55.55 kg/m3). The properties of blocks are: density,
2100 kg/m3; compressive strength, 9.87 MPa; water absorption
(24 h), <10%; drying shrinkage, <1%. The hollow blocks of size
400  300  200 mm were also produced on a block making
machine. The properties of blocks are: density, 1200 kg/m3; compressive strength, 5 MPa; net weight, 20 kg. It was noted that the
cost of solid and hollow geopolymer blocks was about 15% and
10% higher than the OPC concrete blocks (Fig. 7(c)).
7. Conclusions
Based on the discussions, it is concluded that geopolymer concrete has considerable potential to be used as a construction

material in several applications. A number of key properties have


been investigated and very high strengths have been attained.
The design provisions mentioned in ACI guidelines and other
National Code for OPC concrete are reported to be applicable for
geopolymer concrete also. The production of ready mixed geopolymer concrete can be achieved which represents the successful
implementation of a technically very challenging product.
However, it presents signicant scientic challenges associated
with the need for a better understanding of the setting reactions
involved, the relationship between mix design characteristics, the
short and long term mechanical properties and overall durability.
Although, signicant progress was made, there is an immense need
to work out generalisation of water-geopolymer solids ratio, bond
between reinforcement and geopolymer paste, structural behaviour of reinforced GPC members, corrosion of reinforcement in
geopolymer concrete etc. An appropriate code of practice for
geopolymers and their products need to be formulated based on
research data and eld data for mass adaptation by the users. It
is felt that the widespread uptake of geopolymer technology is hindered by a number of factors, in particular issues to do with a lack
of long term-durability data. In this relatively new research eld,
there are also difculties in compliance with regulatory standard,
specically those dening chemical composition in cement.
Conventionally, geopolymer binders require heat curing, high
pH and also have difculty in eld handling. Therefore, efforts
are needed to develop a room temperature cured one component
geopolymer system using solid activators instead of alkaline solutions in view of its wider acceptance in the eld.

Acknowledgements
This paper forms part of a Supra Institutional Project of CSIR
R&D program (Govt. of India) and is published with the permission
of Director, CSIR-Central Building Research Institute, Roorkee.
Authors are also thankful to Ms. Sarika Sharma, Mr. Ankur Singh,
Mr. Rakesh Paswan and Mr. Md. Reyazur Rahman for their help
during the work.

References
[1] http://www.worldcement.com/news/cement/articles/Cement_India_demand_
price_capacity_160.aspx#.UoIUv3DItSk, 12 November 2013.
[2] Shi C, Fernandez-Jiminez A, Palomo A. New cements for the 21st century: the
pursuits of an alternative to Portland cement. Cem Concr Res 2002;32:86579.
[3] Krivenko PV, Kovalachuk GYu. Directed synthesis of alkaline aluminosilicate
minerals in a geocement matrix. J Mater Sci 2007;42:294453.

B. Singh et al. / Construction and Building Materials 85 (2015) 7890


[4] Duxon P, Fernandez-Jiminez A, Provis JL, Luckey GC, Palomo A, Van Deventure
JSJ. Geopolymer technology: the current state of the art. J Mater Sci
2007;42:291733.
[5] Provis JL, VanDeventer JSJ, editors. Geopolymers, structure, processing,
properties and application. UK: Woodhead Publishing Limited; 2009.
[6] Li C, Sun H, Li L. A review: the comparison between alkali-activated slag (Si+Ca)
and metakaolin (Si+Al) cements. Cem Concr Res 2010;40:13419.
[7] Komnitsas KA. Potential of geopolymer technology towards green buildings
and sustainable cities. In: International conference on green buildings and
sustainable cities, procedia engineering, vol. 21; 2011: p.102332.
[8] Davidovits J. Geopolymers: inorganic polymeric new materials. J Therm Anal
1991;37:163356.
[9] Davidovits J, Buzzi L, Rocher P, Marini DGC, Tocco S. Geopolymeric cement
based on low cost geologic materials geocistem. In: second international
conference geopolymer 99, France; 1999.
[10] Van Jaarsveld J, Van Deventure J. Effect of alkali metal activators on the
properties of y ash-based geopolymer. Ind Eng Chem Res 1999;38:3932.
[11] Xu H, Van Deventure JSV. Geopolymerisation of multiple minerals. Min. Eng.
2002;15:11319.
[12] Zhang B, MacKenzie KJD, Brown IWM. Crystalline phase formation in
metakaolinite geopolymers activated with NaOH and sodium silicate. J
Mater Sci 2009;44:466876.
[13] Ishwarya G. Development of geopolymer concrete cured at ambient
temperature [M. Tech thesis]. Roorkee, India: CSIR-Central Building Research
Institute (AcSIR); 2013.
[14] Silva PD, Crenstil KS, Sirivivatnanon V. Kinetics of geopolymerization: role of
Al2O3 and SiO2. Cem Concr Res 2007;37:5128.
[15] de Vargas AS, DalMolin DCC, Vilela ACV, da Silva FJ, Pavo B, Veit H. The effects
of Na2O/SiO2 molar ratio, curing temperature and age on compressive
strength, morphology and microstructure of alkali-activated y ash-based
geopolymers. Cem Concr Comp 2011;33:65360.
[16] Yip CK, Luckey GC, Provis JL, van Deventure JSV. The coexistence of
geopolymeric gel and calcium silicate hydrate at the early stage of alkaline
activation. Cem Concr Res 2005;35:168897.
[17] Somna K, Jaturapitakkul C, Kajitvichyanukul P, Chindaprasirt P. NaOHactivated ground y ash geopolymer cured at ambient temperature. Fuel
2011;90:211824.
[18] Yip CK, Luckey GC, Provis JL, van Deventure JSV. Effect of calcium silicate
sources on geopolymerisation. Cem Concr Res 2008;38:55464.
[19] Kumar S, Kumar R, Melhotra SP. Inuence of granulated blast furnace slag on
the reaction, structure and properties of y ash-based geopolymer. J Mater Sci
2010;45:60715.
[20] Puligilla S, Mondal P. Role of slag in microstructural development and
hardening of y ash-slag geopolymer. Cem Concr Res 2013;43:7080.
[21] Chindaprasirt P, Silva PD, Crentsil KS, Hanjitsuwan S. Effect of SiO2 and Al2O3
on the setting and hardening of high calcium y ash-based geopolymer
systems. J Mater Sci 2012;47:487683.
[22] Guo X, Shi H, Dick WA. Compressive strength and microstructural
characteristics of class C y geopolymer. Cem Concr Comp 2010;32:1427.
[23] Kusbiantoro A, Ibrahim MS, Muthusamy K, Alias A. Development of sucrose
and citric acid as natural based admixture for y ash based geopolymer. Proc
Environ Sci 2013;17:596602.
[24] Nemotallahi B, Sanjayan JG. Effect of different superplasticizer and activator
combinations on workability and strength of y ash based geopolymer. Mater
Des 2014;57. 667-67.
[25] Jang JG, Lee NK, Lee HK. Fresh and hardened properties of alkali-activated
y ash/slag pastes with superplasticizers. Constr Build Mater
2014;50:16976.
[26] Palomo A, Grutzeck MW, Blanco MT. Alkali-activated y ashes: a cement for
the future. Cem Concr Res 1999;29:13239.
[27] Perera DS, Uchida O, Vance ER, Finnie KS. Inuence of curing schedule on the
integrity of geopolymer. J Mater Sci 2007;42:3099106.
[28] Heah CY, Kamarudin H, Al Bakri AMM, Binhussain M, Luqman M, Nizar IK, et al.
Effect of curing prole on Kaolin-based geopolymers. Phys Procedia
2011;22:30511.
[29] Rovnanik P. Effect of curing temperature on the development of hard structure
of metakaolin-based geopolymer. Constr Build Mater 2010;24:117683.
[30] Kani EN, Allahverdi A. Effects of curing time and temperature on strength
development of inorganic polymeric binder based on natural pozzolan. J Mater
Sci 2009;44:308897.
[31] Van Jaarsveld JGS, Van Deventer JSJ, Lukey GC. The effect of composition and
temperature on the properties of y ash and kaolinite based geopolymers.
Chem Eng J 2002;89:6373.
[32] Chindaprasirt P, Chareerat T, Sirivivatnanon V. Workability and strength of
coarse high calcium y ash geopolymer. Cem Concr Comp 2007;29:2249.
[33] Chindaprasirt P, Chareerat T, Hatanaka S, Cao T. High strength geopolymer
using ne high-calcium y ash. J Mater Civ Eng 2011;23:26470.
[34] Temuujin J, van Riessen A, Mackenzie. Preparation and characterisation of y
ash-based geopolymer mortars. Constr Build Mater 2010;24:190610.
[35] Khandelwal M, Ranjith PG, Pan Z, Sanjayan JG. Effect of strain rate on strength
properties of low-calcium y ash-based geopolymer mortar under dry
condition. Arabian J Geosci 2013;6:23839.
[36] Hardjito D, Fung SS. Fly ash-based geopolymer mortar incorporating bottom
ash. Mode Appl Sci 2010;4:4452.
[37] Sathonsaowaphak A, Chindaprasirt P, Pimraksa K. Workability and strength of
lignite bottom ash geopolymer mortar. J Hazard Mater 2009;168:4450.

89

[38] Brough AR, Atkinson A. Sodium silicate-based alkali-activated slag mortars Part
I. Strength, hydration and microstructure. Cem Concr Res 2002;32:86579.
[39] Yang KH, Song JK, Lee KS, Ashour AF. Flow and compressive strength of alkaliactivated mortars. ACI Mater J 2009;106:508.
[40] Yang KH, Song JK. Workability loss and compressive strength development of
cementless mortars activated by combination of sodium silicate and sodium
hydroxide. J Mater Civ Eng ASCE 2009;21:11927.
[41] Hardjito D, Wallah SE, Sumajouw DMJ, Rangan BV. On the development of y
ash-based geopolymer concrete. ACI Mater J 2004;101:46772.
[42] Laskar AI, Bhattacharjee R. Rheology of y ash-based geopolymer concrete. ACI
Mater J 2011;108:53642.
[43] Montes C, Allouche EN. Rheological behaviour of y ash-based geopolymers.
STP 1566 on Geopolymer Binder Systems 2013, ASTM: p. 7284.
[44] Palacios M, Banll Phillip FG, Puertas F. Rheology and setting of alkaliactivated slag pastes and mortars: effect of organic admixture. ACI Mater J
2008;105:1408.
[45] Fernandez-Jiminez AM, Palomo A, Lopez-Hombrados C. Engineering properties
of alkali-activated y ash concrete. ACI Mater J 2006;103:10612.
[46] Olivia M, Nikraz H. Properties of y ash geopolymer concrete designed by
taguchi method. Mater Des 2012;36:1918.
[47] Singh A. Engineering properties of reinforced geopolymer concrete [B. Tech
project report]. Roorkee, India. CSIR-Central Building Research Institute; 2012.
[48] So M, Van Deventer JSJ, Mendis PA, Lukey GC. Engineering properties of
inorganic polymer concretes. Cem Concr Res 2007;37:2517.
[49] Diaz-Loya EI, Allouche EN, Vaidya S. Mechanical properties of y ash-based
geopolymer concrete. ACI Mater J 2011;108:3006.
[50] Ryu GS, Lee YB, Koh KT, Chung YS. The mechanical properties of y ash-based
geopolymer concrete with alkaline activators. Constr Build Mater
2013;47:40918.
[51] Bondar D, Lynsdale CY, Milestone NB, Hassani N, Ramezanianpour AA.
Engineering properties of alkali-activated natural pozzolan concrete. ACI
Mater J 2011;108:6472.
[52] Yost JR, Radlinska A, Ernst S, Salera M. Structural behavior of alkali activated
y ash concrete. Part 1: mixture design, material properties and sample
fabrication. Mater Struct 2012;46:43547.
[53] Pan Z, Sanjayan JG, Rangan BV. Fracture properties of geopolymer paste and
concrete. Mag Concr Res 2011;63:76377.
[54] Sarker PK, Haque R, Ramgolam KV. Fracture properties of heat cured y ashbased geopolymer concrete. Mater Des 2013;44:5806.
[55] Mehta PK, Monteiro PJM. Concrete: microstructure, properties and
materials. USA: The McGraw Hill companies, Inc.; 2006.
[56] Lee WKW, Van Deventure JSJ. The interface between natural siliceous
aggregates and geopolymers. Cem Concr Res 2004;34:195206.
[57] Lee WKW, Van Deventure JSJ. Chemical interactions between siliceous
aggregates and low-Ca alkali-activated cements. Cem Concr Res
2007;37:84455.
[58] Zhang YS, Sun W, Li JZ. Hydration process of interfacial transition in potassium
polysialate (K-PSDS) geopolymer concrete. Mag Concr Res 2005;57:338.
[59] Demie S, Nuruddin MF, Shaq N. Effects of micro-structure characteristics of
interfacial transition zone on the compressive strength of self-compacting
geopolymer concrete. Constr Build Mater 2013;41:918.
[60] ACI 408R-03. Bond and development of straight reinforcing bars in tension.
[61] Sarker PK. Bond strength of reinforcing steel embedded in y ash-based
geopolymer concrete. Mater Struct 2011;44:102130.
[62] So M, Van Deventer JSJ, Mendis PA, Lukey GC. Bond performance of
reinforcing bars in inorganic polymer concrete (IPCs). J Mater Sci
2007;42:310716.
[63] Sumajouw DMJ, Hardjito D, Wallah SE, Rangan BV. Fly ash-based geopolymer
concrete: study of slender reinforced columns. J Mater Sci 2007;42:312430.
[64] Dattatreya JK, Rajamane NP, Ambily PS. Structural behaviour of reinforced
geopolymer concrete beams and columns. CSIR-SERC Research Report RR6;2009.
[65] Yost JR, Radlinska A, Ernst S, Salera M. Structural behaviour of alkali activated
y ash concrete. Part 2: structural testing and experimental ndings. Mater
Struct 2012;46:44962.
[66] Ng TS, Amin A, Foster SJ. The behaviour of steel-bre-reinforced geopolymer
concrete beams in shear. Mag Concr Res 2013;65:30818.
[67] Neville AM. Properties of concrete. 4th ed. India: Dorling Kindersley
Publishing, Inc.; 1997.
[68] Kong DLK, Sanjayan JG, Crentsil KS. Comparative performance of geopolymers
made with metakaolin and y ash after exposure to elevated temperature.
Cem Concr Res 2007;37:15839.
[69] Kong DLK, Sanjayan JG, Crentsil KS. Factors affecting the performance of
metakaolin geopolymers exposed to elevated temperatures. J Mater Sci
2008;43:82431.
[70] Kong DLY, Sanjayan JG. Effect of elevated temperatures on geopolymer paste,
mortar and concrete. Cem Concr Res 2010;40:3349.
[71] Zhao R, Sanjayan JG. Geopolymer and Portland cement concretes in simulated
re. Mag Concr Res 2011;63:16373.
[72] Swamy RN, editor. The alkali-silica reaction in concrete. UK: Blackie and Sons
Limited; 1992.
[73] Patil KK, Allouche EN. Impact of alkali silica reaction on y ash-based
geopolymer concrete. J Mater Civ Eng ASCE 2013;25:1319.
[74] Fernandez-Jimenez A, Puertas F. The alkali-silica reaction in alkali-activated
granulated slag mortars with reactive aggregate. Cem Concr Res
2002;32:101924.

90

B. Singh et al. / Construction and Building Materials 85 (2015) 7890

[75] Garcia-Lodeiro I, Palomo A, Fernandez-Jimenez A. The alkali-aggregate


reaction in alkali activated y ash mortars. Cem Concr Res 2007;37:17583.
[76] Singh B, Ishwarya G, Gupta M, Bhattacharyya SK. Performance evaluation of
geopolymer concrete through alkali-silica reaction. In: Advances in chemically
activated materials, Changsha, China; Jun 13, 2014.
[77] Temuujin J, Minjigmaa A, Lee M. Characterisation of class F y ash geopolymer
pastes immersed in acid and alkaline solutions. Cem Concr Comp
2011;33:108691.
[78] Bakharev T. Resistance of geopolymer materials to acid attack. Cem Concr Res
2005;35:65870.
[79] Arifn MAM, Bhutta MAR, Hussin MW, Mohd Tahir M, Aziah N. Sulfuric acid
resistance of blended ash geopolymer concrete. Constr Build Mater
2013;43:806.
[80] Sathia R, Ganesh Babu K, Santhanam M. Durability study of low calcium y ash
geopolymer concrete. In: 3rd ACF international conference, Ho chi minh city,
Vietnam; 2008.
[81] Bakharev T, Sanjayan JG, Cheng Y-B. Resistance of alkali-activated slag
concrete to acid attack. Cem Concr Res 2003;33:160711.
[82] Fernandez-Jimenez A, Garcia-Lodeiro I, Palomo A. Durability of alkali-activated
y ash cementitious material. J Mater Sci 2009;42:305565.
[83] Bakharev T. Durability of geopolymer materials in sodium and magnesium
sulfate solutions. Cem Concr Res 2005;35:123346.
[84] Ismail I, Bernal SA, Provis JL, Hamdan S, van Deventer JSJ. Microstructural
changes in alkali activated y ash/slag geopolymers with sulphate exposure.
Mater Struct 2013;46:36173.

[85] Bhattacharyya SK, Singh B. High performance materials and construction


technologies for sustainable built space. Supra Institutional Project Report (SIP
29), CSIR-Central Building Research Institute, Roorkee, India; 2012.
[86] Singh B, Sharma S, Gupta M, Bhattacharyya SK. Performance of y ash-based
geopolymer pastes under chemical environment. In: International conference
on advances in construction materials through science and engineering, Hong
Kong; 57 September, 2011.
[87] Rajamane NP, Natraja MC, Dattatreya JK, Lakshmanan N, Sabitha D. Sulphate
resistance and eco-friendliness of geopolymer concretes. Ind Concr J
2012;86:1321.
[88] Bernal SA, Gutierrez RM, Provis JL. Engineering and durability properties of
concrete based on alkali-activated granulated blast furnace slag/metakaolin
blends. Constr Build Mater 2012;33:99108.
[89] Reddy DV, Edouard JB, Sobhan K, Tipni A. Experimental evaluation of the
durability of y ash-based geopolymer concrete in the marine environment.
In: 9th Latin American and Caribbean conference on engineering for a smart
planet, innovation, information technology and computational tools for
sustainable development Colombia, Australia; 2011.
[90] Singh B, Gupta M, Chauhan M, Bhattacharyya SK. Lightweight geopolymer
concrete with EPS. In: CIB World Building Congress 2013, Brisbane, Australia;
59 May, 2013.

Vous aimerez peut-être aussi