Вы находитесь на странице: 1из 13

Construction and Building Materials 85 (2015) 78–90

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 fly 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 field.

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

Article history: An overview of advances in geopolymers formed by the alkaline activation of aluminosilicates is pre-
Received 26 November 2013 sented alongwith opportunities for their use in building construction. The properties of mortars/concrete
Received in revised form 16 February 2015 made from geopolymeric binders are discussed with respect to fresh and hardened states, interfacial
Accepted 4 March 2015
transition zone between aggregate and geopolymer, bond with steel reinforcing bars and resistance to
Available online 31 March 2015
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
Keywords:
geopolymers at CSIR-CBRI are briefly outlined alongwith the product developments. Research findings
Geopolymer concrete
Activator
revealed that geopolymer concrete exhibited comparative properties to that of OPC concrete which
Bond strength has potential to be used in civil engineering applications.
Compressive strength Ó 2015 Elsevier Ltd. All rights reserved.
Durability

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
2. An overview of geopolymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
2.1. Constituents effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2.2. C-S-H phase effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2.3. Effect of admixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2.4. Curing conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3. Geopolymer mortars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4. Geopolymer concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

⇑ Corresponding author.
E-mail address: singhb122000@yahoo.com (B. Singh).

http://dx.doi.org/10.1016/j.conbuildmat.2015.03.036
0950-0618/Ó 2015 Elsevier Ltd. All rights reserved.
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 79

4.1. Fresh and hardened properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81


4.2. Interfacial transition zone. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.3. Bond between reinforcing bars and geopolymer concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.4. Fire behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5. Durability studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.1. Alkali-silica reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.2. Effect of acid attack. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.3. Effect of sulphate attack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.4. Carbonation and permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.5. Corrosion of steel reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
6. Research and development at CSIR-CBRI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
7. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

1. Introduction amorphous and possess sufficient reactive glassy content, low


water demand and be able to release aluminium easily. The alka-
The concrete industry faces challenges to meet the growing line activators such as sodium hydroxide (NaOH), potassium
demand of Portland cement due to limited reserves of limestone, hydroxide (KOH), sodium silicate (Na2SiO3) and potassium silicate
slow manufacturing growth and increasing carbon taxes. It is (K2SiO3) are used to activate aluminosilicate materials. Compared
reported that the requirement of cement in India is likely to touch to NaOH, KOH showed a greater level of alkalinity. But in reality,
550 million tonnes by 2020 with a shortfall of 230 million ton- it has been found that NaOH possesses greater capacity to liberate
nes (58%) and the demand for cement has been constantly silicate and aluminate monomers [4]. The properties of geopoly-
increasing due to increased infra-structural activities of the coun- mers can be optimised by proper selection of raw materials, correct
try [1]. One effort to combat shortfall is the development of alter- mix and processing design to suit a particular application [4].
nate binders to Portland cement aiming at to reduce the Viewing the importance of the subject, a collaborative project
environmental impact of construction, use of greater proportion sponsored by the European Commission – BRITE was undertaken
of waste pozzolan, and also to improve concrete performance. jointly by France, Spain and Italy on development of ‘‘Cost-effective
Search for several alternatives such as alkali-activated cement, cal- geopolymeric cement for innocuous stabilization of toxic elements
cium sulphoaluminate cement, magnesium oxy carbonate cement (GEOCISTEM)’’. The project was aimed at manufacturing geopoly-
(carbon negative cement), supersulphated cement etc. are being meric cement by replacing potassium silicate with cheaper alkaline
made with the advantages of Portland cement [2]. As the family volcanic tuffs [9].
of the alkali-activated cement is growing, the alkaline cement is Geopolymers are synthesized by the reaction of a solid
classified based on a phase composition of the hydration products: aluminosilicate powder with alkali hydroxide/alkali silicate [8]. A
R-A-S-H (R = Na+ or K+) in the aluminosilicate based systems and R- schematic representation on formation of fly ash-based geopoly-
C-A-S-H in the alkali-activated slag or alkaline Portland cements mers/concrete is shown in Fig. 1. Under highly alkaline conditions,
[3]. In recent years, geopolymer has attracted considerable atten- polymerisation takes place when reactive aluminosilicates are
tion among these binders because of its early compressive rapidly dissolved and free [SiO4] and [AlO4] tetrahedral units
strength, low permeability, good chemical resistance and excellent are released in solution. The tetrahedral units are alternatively
fire resistance behaviour [4–9]. Because of these advantageous linked to polymeric precursor by sharing oxygen atom, thus form-
properties, the geopolymer is a promising candidate as an alterna- ing polymeric Si–O–Al–O bonds. The following reactions occur dur-
tive to ordinary Portland cement for developing various sustain- ing geopolymerisation [7].
able products in making building materials, concrete, fire
resistant coatings, fibre reinforced composites and waste ðSi2 O5 Al2 O2 Þn þ H2 O þ OH ! SiðOHÞ4 þ AlðOHÞ4 ð1Þ
immobilization solutions for the chemical and nuclear industries.

ð2Þ
2. An overview of geopolymers
This process releases water that is normally consumed during
Geopolymer is considered as the third generation cement after dissolution. The water, expelled from geopolymer during the reac-
lime and ordinary Portland cement. The term ‘‘geopolymer’’ is tion provides workability to the mixture during handling. This is in
generically used to describe a amorphous alkali aluminosilicate contrast to the chemical reaction of water in Portland cement mix-
which is also commonly used for to as ‘‘inorganic polymers’’, ture during the hydration process. It is reported that the hydration
‘‘alkali-activated cements’’, ‘‘geocements’’, ‘‘alkali-bonded cera- products of metakaolin/fly ash activation are zeolite type: sodium
mics’’, ‘‘hydroceramics’’ etc. Despite this variety of nomenclature, aluminosilicate hydrate gels with different Si/Al ratio whereas the
these terms all describe materials synthesized utilising the same major phase produced in slag activation is calcium silicate hydrate
chemistry [4]. It essentially consists of a repeating unit of sialate with a low Ca/Si ratio. Though many physical properties of
monomer (–Si–O–Al–O–). A variety of aluminosilicate materials geopolymers prepared from various aluminosilicate sources may
such as kaolinite, feldspar and industrial solid residues such as appear to be similar, their microstructures and chemical properties
fly ash, metallurgical slag, mining wastes etc. have been used as vary to a large extent. The metakaolin-based geopolymer has an
solid raw materials in the geopolymerization technology. The advantage that it can be manufactured consistently, with pre-
reactivity of these aluminosilicate sources depends on their chemi- dictable properties both during the preparation and development.
cal make-up, mineralogical composition, morphology, fineness and However, its plate-shaped particles lead to rheological problems,
glassy phase content. The main criteria for developing stable increasing the complexity of processing as well as the water
geopolymer are that the source materials should be highly demand of the system [6]. Contrary to this, the fly ash-based
80 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

Fig. 1. Conversion of fly 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 3.0–3.8 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 geopoly-
mer pastes increased with increasing SiO2/Al2O3 ratio of the initial
mixture.

2.2. C-S-H phase effect

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 [16–22]. 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 fly ash-based
Fig. 2. Pore size distribution of fly ash-based geopolymer pastes at different geopolymer [17] which are responsible for the strength increase.
activator dosages [13]. 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
geopolymer is generally more durable and stronger than that of pastes [18]. In the case of fly ash/slag blends, the reaction at
metakaolin-based geopolymer [4]. The slag-based geopolymer is 27 °C is dominated by the slag activation, whereas the reaction at
considered to have high early strength and greater acid resistance 60 °C is due to combined activation of fly ash and slag. The
than those of metakaolin and fly ash-based systems. improvement in compressive strength of pastes with slag addition
is attributed to its compactness of the microstructure [19]. The
2.1. Constituents effect initiation of hardening in fly ash/slag geopolymer made with
potassium silicate and potassium hydroxide was due to C-S-H/C-
The most important factors affecting the properties of geopoly- A-S-H formation and the hardening continues due to a rapid for-
mer pastes are: SiO2/Al2O3 ratio, R2O/Al2O3 ratio, SiO2/R2O ratio mation of a C-A-S-H, K-A-S-H and (Ca, K)-A-S-H depending on
(R = Na+ or K+) and liquid–solid ratio. The majority of research con- the availability of calcium ions and pH of the system. A slower dis-
cluded that an amorphous structure of geopolymers is preferable solution rate of calcium ions effectively increased the compressive
in order to achieve desired mechanical strength [10–15]. The strength as rapid geopolymerization continues for a longer dura-
relationship between the compressive strength and SiO2/R2O ratio tion [20]. The low pH and limited calcium ion environment facili-
showed that an increase in alkali content or decrease in silicate tate the polymerisation reaction between silicate and aluminate
content increases the compressive strength of geopolymers attri- species in high calcium fly ash-based geopolymers producing N-
butable to the formation of aluminosilicate network structures A-S-H gel [21]. Guo et al. [22] reported 63.4 MPa compressive
[10,11]. Geopolymer activated with NaOH alone with Si/Na of 4/4 strength of class C fly ash-based geopolymer paste showing the
or less formed the crystalline zeolite (Na96Al96Sr96O384216H2O) role of calcium participation in the strength development.
but at a ratio >4/4, nanosized crystals of another zeolite
(Na6[AlSiO4]64H2O) were formed [12]. The addition of even small 2.3. Effect of admixtures
amount of sodium silicate to the NaOH significantly reduces crys-
tallite formation due to templating function of silicate units. At low Kusbiantora et al. [23] reported from their studies that admix-
activator dosage (18%), the pores developed in the fly ash-based tures such as sucrose and citric acid which act as retarder in OPC
paste were larger and exhibited wider distributions (19.8– have different mechanism in fly ash-based geopolymers. Sucrose
2342 A) whereas at higher activator dosage (30%), the pores were acted as a retarder since it is absorbed by Ca, Al and Fe ions to form
smaller and showed a narrow distribution (19.8–1155 A) mainly insoluble metal complexes. On the other hand, citric acid acted as
due to the pore refinement as a result of more dissolution of an accelerator reducing the setting time by 9 and 16 min
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 81

respectively. Amongst the commercial superplasticizers, the naph- (14 M activator solution) with 10–30 wt% aggregate exhibited an
thalene based superplasticizer was effective when single activator acceptable flowability, while the mortars containing 40 & 50 wt%
was used rendering 136% increase in relative slump without any aggregate were stiff and difficult to pack in the mould. Increasing
decrease in compressive strength. Modified polycarboxylate based aggregate content in the mortar mixes leads to insufficient activa-
superplasticizer was efficient one when multi-compound activator tor for complete geopolymerization of fly ash/slag. The activator
was used with a decrease in compressive strength of 29% [24]. may also be utilised for wetting of aggregate leaving less availabil-
However, retarding effect of polycarboxylate based super- ity for dissolution of these fly ash or slag particles. The compressive
plasticizer was also reported in fly ash/slag blended system though strength of geopolymer mortars with high level of aggregate can be
the improvement in workability was significant compared to naph- achieved by optimising the amount of activator dosage [34].
thalene based superplasticizer [25]. Khandelwal et al. [35] summarised that the compressive strength,
modulus of elasticity and Poisson’s ratio of fly ash-based geopoly-
mer mortars increased logarithmically with the increase of strain
2.4. Curing conditions
rate. These engineering properties of geopolymer mortars com-
pared favourably with those predicted by Standards/Codes for con-
Several attempts [26–31] have been made to study the effect of
crete mixtures. When bottom ash was used, the geopolymer
different curing conditions on the properties of geopolymer pastes.
mortars exhibited a low compressive strength (20 MPa). With
The curing temperatures were reported in the range between 40 °C
10% replacement of sand by bottom ash, the mix exhibited a com-
and 85 °C for complete geoplymerisation reactions. Palomo et al.
parable compressive strength to those made with sand only. The
[26] studied curing of alkali activated fly ash (0.25 and 0.30 liq-
increase in strength (50–100%) of bottom ash mortar was also
uid/solid ratio) at 65 °C and 85 °C. They indicated that the com-
reported when the specimens were exposed at 800 °C probably
pressive strength of geopolymers (8–12 M) cured at 85 °C for
due to activation of bottom ash [36]. When lignite bottom ash
24 h was much higher than those cured at 65 °C. The rise of
was ground to a mean particle size of 15.7 lm (3% retained on
strength was much smaller when curing time was extended after
sieve No. 325), the compressive strength of mortars activated with
24 h. Perera et al. [27] studied the curing of metakaolin-based
sodium hydroxide/sodium silicate was 24–58 MPa [37].
geopolymers under ambient (21–23 °C) and heat conditions (40–
Brough and Atkinson [38] prepared geopolymer mortars using
60 °C) with a controlled relative humidity (RH) for 24 h and found
slag, sand and activator in a ratio of 1:2.33:0.5. At water-to-total
that curing at 30% RH was preferable to that at 70% RH. Heah et al.
solid ratio of 0.42, the mortar gained strength of 40 MPa. The
[28] concluded that the curing of metakaolin-based geopolymers
sodium silicate activated mortars exhibited higher compressive
at ambient temperature was not feasible while increase in tem-
strength with low levels of porosity at the interface while KOH
perature (40 °C, 60 °C, 80 °C, 100 °C) favored the strength gain after
activated mortars were highly porous in the interfacial zone giving
1–3 days. However, curing at higher temperature for a longer per-
low compressive strength values. Yang et al. [39,40] found that the
iod of time caused failure of samples at a later age due to the ther-
flow of alkali-activated mortars increased with the increase of
molysis of –Si–O–Al–O– bond. Rovnanik [29] reported that curing
water-binder ratio and decrease of aggregate-binder ratio. When
of metakaolin based geopolymer at elevated temperature (40–
the aggregate-binder ratio was larger than 2.5, the flow of mortars
80 °C) accelerated the strength development but in 28 days, the
decreased sharply. They also found that slag-based geopolymer
mechanical properties deteriorated in comparison with results
mortars exhibited much higher compressive strength but exhibited
obtained for an ambient or slightly decreased temperature.
slightly less flow than the fly ash-based geopolymer mortars for
Ebrahim and Ali [30] prepared three mixes with different for-
the same mixing condition. The poor compressive strength of fly
mulations and cured hydrothermally at different temperatures
ash-based mortars cured at low temperatures is attributed to the
(45, 65, 85 °C) and time (5–20 h) after 1 and 7 days of procuring.
presence of unreacted fly ash particles and large number of voids.
Longer procuring at room temperature, before the application of
As the aggregate-binder ratio increased, the compressive strength
heat is beneficial for higher strength development. In general, ade-
increased up to a ratio of 2.5 which indicated that the threshold
quate curing of geopolymeric materials is required to achieve opti-
of aggregates in geopolymer mortars were slightly lower than
mal mechanical and durability performance to maintain their
OPC mortars. The shrinkage strain of alkali-activated mortars was
structural integrity [31].
also found to be lower than the OPC mortars.

3. Geopolymer mortars
4. Geopolymer concrete
Various studies [32–40] were conducted on flow and mechani-
cal properties of geopolymer mortars because of their more rele- 4.1. Fresh and hardened properties
vant applications in building construction. The properties of
mortars were optimised with respect to initial flow, aggregate-bin- Various mix proportioning of geopolymer concrete (GPC) were
der ratio, activator-binder ratio and activator molarity. reported with target strength up to 80 MPa. The typical properties
Chindaprasirt et al. [32] reported that the compressive strength of geopolymer concrete mixes used by the various authors (41, 45,
of class C fly ash-based geopolymer mortar was 52 MPa when 48, 49, 53) are summarised in Table 1. The properties of mixes
cured at 70 °C for 3 days using sand-fly ash ratio of 2.75 at work- were studied with respect to water-geopolymer solid ratio, activa-
able flow of 135 ± 5%. Prolonged curing at high temperatures led tor strength, water/Na2O ratio, curing time, curing temperature,
to the reduction in the compressive strength because of weakening and age hardening. The slump of mixes varied depending on the
of microstructure and increased porosity due to the loss of mois- molarity of activator, workability aids and extra water added to
ture. In another attempt [33], they produced geopolymer mortar the mix [41]. The rheological parameters such as yield stress and
with a compressive strength of 86 MPa at 28 days with the help plastic viscosity were attempted over slump test of concrete to
of air classified class C fly ash (4500 cm2/g fineness) activated with assess its workability loss and flow behaviour. Yield stress gives
sodium silicate and NaOH (10 M) at 1:1 mass ratio. The dimen- initial resistance to flow arose from the friction among the solid
sional change in terms of drying shrinkage (1–61  106 mm/ particles while plastic viscosity governs the flow after it is initiated
mm) was insignificant when compared with the Portland cement resulting from viscous dissipation due to the movement of water in
mortar (700–850  106 mm/mm). The geopolymer mortars the sheared material. Laskar & Bhattacharjee [42] studied the
82 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

Table 1
Typical properties of geopolymer concrete mixes.

Density Molarity Slump CS STS FS MOE Poisson’s Activator/ Curing


(kg/m3) (M) (mm) (MPa) (MPa) (MPa) (GPa) ratio binder temperature
ratio and time
Hardjito et al. [41] 2330–2430 10–16 60–215 30–80 3.74–6 5–12 23–31 0.12–0.16 0.35–0.4 60–80 °C for 24 h
Jimenez et al. [45] NR 8 & 12.5 NR 29–43.5 NR 6.86 10.7–18.4 NR 0.4 & 0.55 85 °C for 20 h
Sofi et al. [48] 2147–2408 NR NR 47–56.5 2.8–4.1 4.9–6.2 23–39 0.23–0.26 0.45–0.59 23 °C till testing
Diaz-loya et al. [49] 1890–2371 14 100–150 10–80 NR 2.24–6.41 1.9–42 0.08–0.22 0.4–0.94 60 °C for 72 h
Pan et al. [53] 1876–2555 8 NR 65.1–77.9 2.8–5.1 NR 11.2–41.2 0.15–0.19 0.4–0.65 60 °C for 24 h

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

rheology of fly ash-based geopolymer concrete with slump varying varied from 100 to 250 mm (activator strength: 8–14 M). The opti-
from 25 mm to flowing concrete with 1–20 M activator strength. mum strength was obtained at 0.18 water-geopolymer solid ratio
They found that the yield stress and plastic viscosity were affected cured at 90 °C. As the water-geopolymer solids ratio increased,
by the molar strength of the sodium hydroxide solution and the the compressive strength of GPC decreased analogous to the well
ratio of silicate to hydroxide solution. The setting time of geopoly- known relationship between compressive strength and water-ce-
mer concrete was reported up to 120 min. Like Portland cement ment ratio for OPC concrete. The compressive strength of GPC
pastes and mortars, geopolymers also behave like Bingham fluid remained unchanged with the age when tested after 24 h curing
and have a history dependent rheological profile, i.e., geopolymers at elevated temperature. Fernandez-Jimenez et al. [45] made fly
may be kept in a fluid form, if subjected to constant shearing for a ash-based geopolymer concrete with a compressive strength of
certain period of time before initial setting starts [43]. The setting 45 MPa at 0.55 liquid/solid ratio cured at 85 °C for 20 h. The devel-
could be enhanced up to 180 min with the use of naphthalene opment of high early strength in GPC was explained by its compact
based admixture and extended mixing time especially in the case microstructure, formation of adequate reaction products, smaller
of slag-based geopolymer which has the potential for a wide range mean size of the pores and good aggregate-paste bond. They
of technological applications [44]. observed that GPC has a much lower modulus of elasticity
Hardjito et al. [41] produced fly ash-based GPC with the com- (18.4 GPa) than the OPC concrete (30.3 GPa). Olivia and Nikraz
pressive strength ranging between 30 and 80 MPa with the slump [46] proportioned fly ash-based geopolymer concrete mix with a
compressive strength of 55 MPa at 28 days and cured at different
temperatures in the range of 60–75 °C. The hardened mix had
higher tensile and flexural strengths, produced less expansion
and showed modulus of elasticity that were 15–29% 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 mini-
mal shrinkage of GPC may also be due to the significant resistance
offered by its zeolitic microstructure towards drying loss of the
water incorporated during casting [45].
Several attempts [47–53] have also been made to establish
correlations within the mechanical properties of geopolymer con-
crete. It was reported that the experimental splitting tensile
strength of fly 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 con-
crete. Sofi et al. [48] studied the engineering properties of fly ash/
slag-based GPC. The splitting tensile strength and flexural strength
of GPC were comparable to those models presented by the
Australian Standard (AS 3600) for OPC concrete. Although, the dif-
ference between splitting tensile and flexural strength of GPC
mixes has been found to be approximately 2 MPa, similarities
between the strength gain was apparent. Diaz-Loya et al. [49] pro-
p
posed the equation ‘‘fr = 0.69 fc0 MPa’’ for correlation between the
flexural 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 Poisson’s ratio value of OPC con-
crete (0.15–0.22), the values of GPC appeared to reside toward the
low end of range (0.08–0.22). Ryu et al. [50] suggested a model for
relationship between compressive strength and splitting tensile
strength (fsp = 0.17 (f0 c)3/4) for fly ash-based GPC. Bondar et al.
[51] reported a relationship between ultrasonic pulse velocity
Fig. 3. Correlations within the mechanical properties of fly ash-based geopolymer
and compressive strength of GPC. They found that GPC showed a
concrete. (a) Splitting tensile strength vs compressive strength. (b) Modulus of
elasticity vs compressive strength [47]. lower ultrasonic pulse velocity than the OPC concrete even those
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 83

with the same or higher compressive strength. It was also reported FESEM analysis of ITZ in the fly ash-based self compacting geopoly-
[52,53] that GPC was brittle as compared to its OPC counterpart mer concrete with varying superplasticizer dosages. They reported
due to the highly cross-linked framework. The fracture energy of that relatively a loose and porous ITZ was found at low super-
GPC was also low because of its higher bond with aggregates as plasticizer dosages (3%) whereas a dense ITZ was found between
compared to OPC concrete [54]. 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 super-
4.2. Interfacial transition zone
plasticizer dosage.
It is well known that the interfacial zone (ITZ) between aggre-
4.3. Bond between reinforcing bars and geopolymer concrete
gate and matrix is the weakest link in OPC concrete at which
micro-cracks usually first develop under loads [55]. Investigation
The transfer of forces across the interface between concrete and
of this zone is very crucial since it is known to have different
reinforcing steel bar is of fundamental importance in the structural
microstructure from the bulk of the hardened paste. The high
design [60]. Bond stresses in the reinforced concrete arise from two
porosity of ITZ allows the easier penetration of external agents
distinct situations. The first is anchorage or development where
such as chlorides, oxygen, sulphates, etc. into concrete structure.
bars are terminated. The second is flexural bond or the change of
Contrary to this, ITZ of GPC has been identified as being dense
force along a bar due to a change in bending moment along the
and much less microstructurally distinct from the bulk of binder
member. The bond strength of reinforcing bars with concrete is
region [56,57]. The stronger ITZ contributes to higher splitting ten-
governed by several factors such as the strength of the concrete,
sile strength, bond strength and durability of the GPC.
the thickness of the concrete surrounding the reinforcing bar, the
Lee and Deventer [56,57] discussed interface between the natu-
confinement of the concrete due to transverse reinforcement and
ral siliceous aggregates and paste in GPC using kaolin and albite as
the bar geometry. Generally, the bond strength between the
precursors. The increase in concentration of the activating solution
reinforcing bar and matrix increases with increasing steel bar
increased the binding capacity of the gel with natural aggregates.
diameter and compressive strength of GPC (Fig. 4). There is a
The presence of chloride salts decreased the interfacial bonding
greater amount of slip for larger size rebars in GPC.
strength between the paste and aggregate probably by causing
Sarker [61] found that the bond strength of fly ash-based GPC
gel crystallisation near the aggregate surfaces which resulted in
increased with the increase of concrete cover-bar diameter ratio
debonding. In another attempt, they found that the addition of
(1.71–3.62) and the concrete compressive strength (25–29 MPa).
0.5 M soluble silicate into an activating solution (10 M NaOH and
He also observed that GPC has higher bond strength than the
2.5 M sodium silicate) facilitates the formation of an aluminium-
OPC concrete because of higher splitting tensile strength and dense
enriched aluminosilicate surface onto the aggregates through
interfacial transition zone between the aggregate and geopolymer
accelerated Si-preferential dissolution of kaolin and albite. The sur-
paste. Bond-slip behaviour [45] of GPC showed that the embedded
face formed during albite leaching was found to possess a similar
steel bar of 8 mm dia broke before slipping and concrete cracking
Si/Al ratio to the real interface between a silicious aggregate and
whereas the bar embedded in OPC concrete slipped. For
fly ash/metakaolin geopolymer paste activated with 10 M NaOH
16 mm bar, GPC failed by matrix cracking while the bars in OPC
solution. Without soluble silicates, no deposited aluminosilicate
concrete were again observed to slip. Sofi et al. [62] reported that
interface was observed. This suggested that both high concentra-
the values of bond strength of steel bars in fly ash-based GPC were
tion of alkali and soluble silicate are essential for the formation
comparable in both beam-end as well as direct pullout specimen
of a strong interface between silicious aggregates and geopolymer
tests. The normalised bond strength increased with a reduction
pastes. Zhang et al. [58] reported that at the beginning, there were
in rebar size. The bond strength tested according to AS 3600, ACI
many large voids in the fresh ITZ in potassium poly(sialate)
318-02 and EC2 recommendations showed that these Codes are
geopolymer concrete. As hydration proceeded, these voids were
applicable and also safe to predict the developmental length for
completely filled with the hydration products. At this stage, the
GPC.
difference in the microstructure between the ITZ and matrix was
Attempts were also made to study behaviour of reinforced fly
hardly distinguishable. The contents of K/Al and Si/Al in the ITZ
ash-based GPC beams and columns with respect to longitudinal
were higher than those in the matrix. Demei et al. [59] presented
tensile reinforcement ratio and concrete compressive strength as
test variables [63–66]. Sumajouw et al. [63] reported that the
flexural capacity of beams increased with the increase in tensile
reinforcement (0.64–2.69%) but the effect of concrete compressive
strength was marginal. The ductility index increased significantly
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 load–deflection behaviour of GPC beam was identi-
cal to OPC beam. The maximum strain obtained for under-rein-
forced 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
Fig. 4. Bond strength of fly ash-based geopolymer concrete as a function of steel bar Whitney’s stress block for strength calculation was found applic-
diameter [47]. able for GPC also. Ng et al. [66] investigated potential use of steel
84 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

fibres (up to 1.5 wt%) to replace conventional shear reinforcement carbonation, alkali-silica reaction and freeze–thaw attack. In view of
in GPC beams of 2250 mm span length. They found that the this, several studies are being carried out to understand the beha-
increase in fibre volume led to an increase in the cracking load viour of geopolymers exposed to these conditions.
and the ultimate shear strength. A good correlation of test data
was observed with the predictive fib Model Code 2010. 5.1. Alkali-silica reaction

4.4. Fire behaviour Alkali-silica reaction (ASR) causes gradual but severe deteri-
oration of hardened Portland cement concrete in terms of its
In general, concrete has good property with respect to fire resis- strength loss, cracking, volume expansion etc. It involves the reac-
tance. However, it is known that the residual strength of OPC con- tion between the hydroxyl ion in the pore solution within the con-
crete after firing between 800 °C and 1000 °C does not exceed 20– crete matrix and reactive silica of the aggregate. In general terms,
30% normally because of dehydration and destruction of C-S-H & the reactions will proceed in stages, with the first stage being the
other crystalline hydrates, aggregate types, permeability etc. Fire hydrolysis of reactive silica by hydroxyl ions to form alkali-silica
introduces high temperature gradient and as a result, the hot layer gel and a later secondary overlapping stage being the absorption
tends to separate and spall from the cooler interior layer of the of water by the gel, which will result in increase of volume [72].
body [67]. Contrary to this, geopolymers possess good fire resis-
tance at elevated temperature because of the existence of highly (i) Acid-based reaction
distributed nano-pores in the ceramic like microstructure that
H0:38 SiO2:19 þ 0:38NaOH ! Na0:38 SiO2:19 þ 0:38H2 O ð3Þ
allows physically and chemically bonded water to migrate and
evaporate without damaging the aluminosilicate network [4].
During fire, several events such as evaporation of water adsorbed (ii) Attack of the siloxane bridges and disintegration of the silica
by N-A-S-H gel, formation of anhydrous products, crystallization
2
of stable anhydrous phases and melting (sintering) leading to Na0:38 SiO2:19 þ 1:62NaOH ! 2Na2þ þ H2 SiO4 ð4Þ
destruction generally occurred. The phase transformation of
geopolymers during fire is depicted below.
In geopolymer concrete, the un-utilised alkali after geopolymer-
ization 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 with-
out 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 significant expansion in fly ash-
based 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
Kong et al. [68] found that the residual strength of fly ash-based cracking is unlikely [75]. Geopolymer mortar bars made with fly
geopolymer pastes increased by 6% after exposure to 800 °C, ash/slag blends expanded less than 0.1% limit prescribed in ASTM
whereas the strength of metakaolin-based geopolymer pastes was C1260-07 after 16 days (Fig. 5). At 90 days exposure, these mortars
reduced by 34%. During heating, the high permeability of fly ash- failed to meet the specified criteria. Increasing slag content in fly
based geopolymer provides the escape route for moisture in the ash/slag mix increased the expansion of resulting systems [76].
matrix, thereby decreasing the damage. The strength increase is ASR has also been claimed to be helpful in providing a strong bond
also partly attributed to the sintering reaction of unreacted fly ash at the paste-aggregate interface, thus enhancing the tensile
particles. Geopolymer pastes made with metakaolin and potassium strength of GPC [8]. Patil et al. [73] indicated that sandstone, quartz
based activator showed an enhanced post-elevated temperature and limestone aggregates in geopolymer concrete were not prone
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 fly 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 fly 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 fire than the OPC
concrete due to its increased porosity.

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 Fig. 5. Alkali-silica reaction in various geopolymer and OPC mortars under an
assessed for sulphate attack, chloride induced corrosion, atmospheric accelerated condition (1 M NaOH) at 80 °C [76].
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 85

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


noticed because of re-initiation of the geopolymerization process
of unreacted fly ash particles leading to lower porosity and higher
strength. The lower sensitivity of reactive aggregates in GPC pro-
vides 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 [77–84]. 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 fly 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 fly 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 per-
formance of fly 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].

ð5Þ
Wallah and Rangan [41] found that the reduction in compres-
sive strength of fly 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. Ariffin et al. [79] exposed GPC made with a blend
of pulverized fuel ash and palm oil fuel ash in 2% solution of sul-
phuric 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 Fig. 6. Atomic force microscope images of fly ash-based geopolymer exposed under
18 months. The C-S-H could have severe deleterious effect on sulphate after 4 months [86].

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% 5.3. Effect of sulphate attack
H2SO4 solution. Bakharev et al. [81] found that slag-based GPC
(40 MPa) exhibited 33% reduction in strength compared to 47% Fly ash-based geopolymer pastes did not deteriorate signifi-
in OPC concrete when exposed in acetic acid solution (pH 4) for cantly, under the influence of water, sodium sulphate (4.4%)
12 months. The slag particles and low calcium C-S-H with average and ASTM sea water [82]. Only some fluctuations in flexural
Ca/Si ratio of 1 were more stable in the acid solution than the con- strength were observed between 7 days and 3 months exposures.
stituents of the OPC pastes. During immersion in 2% H2SO4 solu- The least strength change was observed in the pastes exposed in
tion, the strength loss was 11% compared to 36.2% for OPC the 5% Na2SO4 and 5% MgSO4 solutions while most significant
concrete. deterioration was observed in the 5% mixed sulphate solution
86 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

(Na2SO4 + MgSO4) after 5 months exposure [83]. In fly ash/slag 5.5. Corrosion of steel reinforcement
system, the extensive physical deterioration of pastes was
observed during immersion in MgSO4 solution after 3 months Corrosion potential is a technique used to detect the state of
exposure but not in Na2SO4 solution. The calcium sulphate dihy- reinforcement without disturbing the structures. This is important
drate formed in paste was identified as being particularly dama- because the intensity of corrosion of steel in concrete is generally
ging to the materials in MgSO4 [84]. Atomic force microscopic known only after the concrete has cracked or disrupted. Various
images of fly ash-based geopolymer pastes exposed to sulphate studies [46,80,89] were reported to estimate the corrosion poten-
environment are shown in Fig. 6. In the case of Na2SO4 solution, tial of steel within the GPC as per ASTM C876. Olivia and Nikraz
only exposition of grains was clearly visible while in MgSO4 solu- [46] reported that the half cell potential of GPC was lower than
tion, both exposition of grains and dissolved aluminosilicate the specified value of 404 mV mentioned in the Standard for sev-
matrix were observed showing severity of MgSO4 attack [85]. ere corrosion after 91 days. Sathia et al. [80] also reported corro-
The deterioration is considered mainly due to the destruction of sion potential up to 300 mV which showed a probable
aluminosilicate skeleton, liberation of silicic acid, leaching of corrosion indication due to the lower pH of concrete during the
sodium ion etc. [86]. These reactions seem to have significant half-cell potential measurement. Accelerated corrosion results
effect on the mechanical strength. The geopolymer prepared with showed that GPC mixes exhibited low level corrosion activity
NaOH activator had the best performance over those made with a and time to failure that were 3.86–5.70 times longer than those
synergistically used sodium silicate and NaOH/KOH activators, of the OPC concrete. Under impressed voltage, a crack appeared
which is attributed to its stable cross-linked aluminosilicate poly- suddenly in the concrete when time to failure was reached and this
mer structure. was followed immediately by high current reading. The large
Several attempts [41,87] have been made to study sulphate amounts of fly ash and alkaline activators in the GPC mix increased
resistance of GPC. The deterioration in concrete was evaluated in the availability of ions that can produce high electrical resistance at
terms of its visual appearance, weight loss and change in compres- high impressed voltage. This enhanced the cathodic reaction and
sive strength. Hardjito et al. [41] observed that there was no sig- reduces the rate of corrosion, which in turn, reduces the tensile
nificant effect of 5% Na2SO4 solution in the compressive strength, stress of the specimens, thus decreasing the risk of cracking and
the weight loss and the dimension of fly ash-based GPC after clearly extending the time to failure [46]. Reddy et al. [89] com-
3 months exposure. Rajamane et al. [87] reported sulphate resis- pared the durability of GPC with that of OPC concrete exposed to
tance of fly ash-based GPC for 3 months in 5% Na2SO4 and 5% marine environment for a period of 21 days. The initial corrosion
MgSO4 solutions. The weight loss in samples was 2.4% only. current measured for GPC (71–91 mA) was much lower than that
There was 2–29% loss of compressive strength as compared to 9– of OPC concrete (772 mA). The OPC specimens initially recorded
38% in the OPC concrete. The deterioration of OPC concrete can decrease in the current but later started increasing while the GPC
be attributed to the formation of expansive gypsum and ettringite current never showed significant increase.
which can cause expansion, cracking and spalling in the concrete.
Contrary to this, GPC in general do not contain Ca(OH)2 and mono-
sulphoaluminate in the matrix to cause expansion. 6. Research and development at CSIR-CBRI

A systematic R&D work is initiated at CSIR-Central Building


5.4. Carbonation and permeability Research Institute, Roorkee on the development of heat and ambi-
ent cured geopolymers using fly ash, slag and other aluminosili-
Bernal et al. [88] studied slag/metakaolin-based GPC (w/b ratio cates as precursors. In view of variability in the constituents of
0.47) under an accelerated carbonation test using CO2 concentra- fly ash, the property optimisation of geopolymeric pastes was car-
tion of 3.0 ± 0.2% at 20 °C for 28 days. They found that the com- ried out as a function of activator concentration and its dosage,
pressive strength decreased monotonically as the carbonation water-geopolymer solid ratio, curing time and curing temperature
proceeds. The relationship between the pore volume and extent [13]. Geopolymerisation reaction, thermal stability, identification
of carbonation was much more similar with samples with differ- of bond linkages and microstructural features were analysed by
ent percentages of metakaolin contrary to the slag-based samples. various techniques such as quasi isothermal DSC, TGA, FTIR and
This suggested that porosity is not the only parameter controlling FESEM. The durability of geopolymer pastes/mortars was also
the strength loss of the carbonated binder. There must be a studied in terms of alkali-silica reaction and also in acidic and sul-
convoluting effect due to the binder gel chemistry, which deter- phate environments for 4 months [76,86]. The suitability of these
mines the residual level of strength after an accelerated carbona- geopolymer pastes was assessed in making various geopolymeric
tion. Olivia and Nikraz [46] reported lower water permeability products such as mortars & concrete, bricks, solid & hollow blocks,
(2.46–4.67  1011 m/s) of GPC (activator-fly ash ratio, 0.30–0.40 insulation concrete, foam, sandwich composites and temperature
cured at 60 °C for 24 h) than the OPC concrete due to its denser resistant coatings (Fig. 7(a–c)). Attempt was also made to utilise
paste and smaller pore inter-connectivity. They also reported that lime sludge, a waste from paper industry with the geopolymeric
the water-geopolymer solids ratio was the most influential binders for making paving blocks.
parameter that affects the properties of GPC. Bondar et al. [51] Fly ash-based GPC mixes were made with the compressive
studied the oxygen and chloride permeability of alkali-activated strength of 25–55 MPa using absolute volume method adopted
concrete made with the Iranian natural pozzolan (Taftan andesite for OPC concrete mixes. The strength of GPC increased with
and Shahindej dacite). They concluded that alkali-activated natu- decreasing water-geopolymer solid ratio as it is said analogous to
ral pozzalona concrete has 10–35% lower oxygen permeability at the water-cement ratio of the OPC concrete. The compressive
normal curing conditions for 90 days compared with the OPC con- strength increased with increasing molarity of the activator (10–
crete. The rapid chloride permeability test gave high values for 16 M) probably due to the formation of stable aluminosilicate net-
the alkali-activated concrete. This is probably due to the very works following the dissolution of silica and alumina in the solu-
high alkali ion concentration in the pore solution promoting tion from the fly ash. It was found that the splitting tensile
higher electrical conductivity in the GPC. This effect seems to strength of GPC was more than those of predicted values as per
reduce with age due to a change in the porosity of the GPC ACI 318 guideline and other existing empirical equations. A trend
microstructure. line curve between the compressive strength and modulus of
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 87

Fig. 7a. Light weight fly 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 val-
ues obtained with ACI guidelines. As expected, the bond strength of when larger size of EPS beads (<4.75 mm) were added in the mor-
steel bar embedded in GPC increased with increasing steel bar tars probably due to their less surface area/volume ratio. By adding
diameter and compressive strength of concrete. It was noted that coarse aggregate, the compressive strength (18 MPa) and density
the bond strength between geopolymer paste and reinforcing bars (1500 and 1840 kg/m3) of EPS geopolymer concrete, comply the
was found to be higher than the OPC concrete [47]. minimum specified criteria of ACI 213R-03 guidelines for struc-
Light weight geopolymer concrete was proportioned with the tural light weight concrete (compressive strength, 17 MPa; density
help of fly ash, activators, expanded polystyrene beads (EPS – up 1120–1920 kg/m3). To meet the requirement of insulation con-
to 3 wt% or 91 vol%), admixtures and fine & coarse aggregates crete, the addition of 20% coarse aggregate (10 mm maximum size)
(Fig. 7(a)). It was noted that a decrease in the strength was more into EPS/geopolymer mix exceeds its compressive strength
88 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

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

(15 MPa) as specified (13.1 MPa) in ASTM C 90. Regarding fire material in several applications. A number of key properties have
performance, the samples were non-ignitable and exhibited Class been investigated and very high strengths have been attained.
I-very low spread of flame as per BS EN-476 part 7 (Fig. 8). It The design provisions mentioned in ACI guidelines and other
was noted that the fire propagation index of the samples was <3 National Code for OPC concrete are reported to be applicable for
exhibiting no support to fire growth. Flammability data obtained geopolymer concrete also. The production of ready mixed geopoly-
from a cone calorimeter showed that the insulation concrete had mer concrete can be achieved which represents the successful
insignificant heat release rate (9.63 kW/m2) and effective heat of implementation of a technically very challenging product.
combustion (3.75 MJ/kg). The thermal conductivity of insulation However, it presents significant scientific challenges associated
concrete was found in the range of 0.427–0.852 W/mK. It was con- with the need for a better understanding of the setting reactions
cluded that light weight concrete can be engineered by proper involved, the relationship between mix design characteristics, the
selection of variables in making insulating materials for use in short and long term mechanical properties and overall durability.
buildings [90]. Although, significant progress was made, there is an immense need
A geopolymer foam composition has been developed using fly to work out generalisation of water-geopolymer solids ratio, bond
ash, activator, filler, surfactant, buffer and strengthening agent. It between reinforcement and geopolymer paste, structural beha-
sets at room temperature within 2 h and completely cured after viour of reinforced GPC members, corrosion of reinforcement in
24 h. The density of foam was lying in the range of 600–800 kg/ geopolymer concrete etc. An appropriate code of practice for
m3. It can be easily prepared by a simple mixing followed by pour- geopolymers and their products need to be formulated based on
ing into mould. SEM examination revealed that pores in the sam- research data and field data for mass adaptation by the users. It
ples were uniformly distributed. The flammability test carried is felt that the widespread uptake of geopolymer technology is hin-
out by a cone calorimeter showed that the total heat released, mass dered by a number of factors, in particular issues to do with a lack
loss, smoke release and CO/CO2 yield were insignificant. It resists of long term-durability data. In this relatively new research field,
against fire between 700 °C and 800 °C. The foam can be used as there are also difficulties in compliance with regulatory standard,
a core material in the sandwich and insulation panels [85]. specifically those defining chemical composition in cement.
Geopolymer bricks of size 230  115  75 mm were produced Conventionally, geopolymer binders require heat curing, high
using fly ash-based pastes, coarse fly ash and natural sand pH and also have difficulty in field handling. Therefore, efforts
(Fig. 7(b)). The bricks were cured at 80 °C for 2 h. The bricks were are needed to develop a room temperature cured one component
obtained with density ranged between 1920 and 2100 kg/m3, geopolymer system using solid activators instead of alkaline solu-
water absorption, 10–15% and dry compressive strength, 12– tions in view of its wider acceptance in the field.
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 Acknowledgements
produced on a machine using coarse aggregate (1180 kg/m3), fine
aggregate (296 kg/m3), flyash (494 kg/m3), activator (111 kg/m3) This paper forms part of a Supra Institutional Project of CSIR
and water (55.55 kg/m3). The properties of blocks are: density, R&D program (Govt. of India) and is published with the permission
2100 kg/m3; compressive strength, 9.87 MPa; water absorption of Director, CSIR-Central Building Research Institute, Roorkee.
(24 h), <10%; drying shrinkage, <1%. The hollow blocks of size Authors are also thankful to Ms. Sarika Sharma, Mr. Ankur Singh,
400  300  200 mm were also produced on a block making Mr. Rakesh Paswan and Mr. Md. Reyazur Rahman for their help
machine. The properties of blocks are: density, 1200 kg/m3; com- during the work.
pressive 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)). References

[1] http://www.worldcement.com/news/cement/articles/Cement_India_demand_
7. Conclusions 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;32:865–79.
Based on the discussions, it is concluded that geopolymer con- [3] Krivenko PV, Kovalachuk GYu. Directed synthesis of alkaline aluminosilicate
crete has considerable potential to be used as a construction minerals in a geocement matrix. J Mater Sci 2007;42:2944–53.
B. Singh et al. / Construction and Building Materials 85 (2015) 78–90 89

[4] Duxon P, Fernandez-Jiminez A, Provis JL, Luckey GC, Palomo A, Van Deventure [38] Brough AR, Atkinson A. Sodium silicate-based alkali-activated slag mortars Part
JSJ. Geopolymer technology: the current state of the art. J Mater Sci I. Strength, hydration and microstructure. Cem Concr Res 2002;32:865–79.
2007;42:2917–33. [39] Yang KH, Song JK, Lee KS, Ashour AF. Flow and compressive strength of alkali-
[5] Provis JL, VanDeventer JSJ, editors. Geopolymers, structure, processing, activated mortars. ACI Mater J 2009;106:50–8.
properties and application. UK: Woodhead Publishing Limited; 2009. [40] Yang KH, Song JK. Workability loss and compressive strength development of
[6] Li C, Sun H, Li L. A review: the comparison between alkali-activated slag (Si+Ca) cementless mortars activated by combination of sodium silicate and sodium
and metakaolin (Si+Al) cements. Cem Concr Res 2010;40:1341–9. hydroxide. J Mater Civ Eng ASCE 2009;21:119–27.
[7] Komnitsas KA. Potential of geopolymer technology towards green buildings [41] Hardjito D, Wallah SE, Sumajouw DMJ, Rangan BV. On the development of fly
and sustainable cities. In: International conference on green buildings and ash-based geopolymer concrete. ACI Mater J 2004;101:467–72.
sustainable cities, procedia engineering, vol. 21; 2011: p.1023–32. [42] Laskar AI, Bhattacharjee R. Rheology of fly ash-based geopolymer concrete. ACI
[8] Davidovits J. Geopolymers: inorganic polymeric new materials. J Therm Anal Mater J 2011;108:536–42.
1991;37:1633–56. [43] Montes C, Allouche EN. Rheological behaviour of fly ash-based geopolymers.
[9] Davidovits J, Buzzi L, Rocher P, Marini DGC, Tocco S. Geopolymeric cement STP 1566 on Geopolymer Binder Systems 2013, ASTM: p. 72–84.
based on low cost geologic materials – geocistem. In: second international [44] Palacios M, Banfill Phillip FG, Puertas F. Rheology and setting of alkali-
conference geopolymer 99, France; 1999. activated slag pastes and mortars: effect of organic admixture. ACI Mater J
[10] Van Jaarsveld J, Van Deventure J. Effect of alkali metal activators on the 2008;105:140–8.
properties of fly ash-based geopolymer. Ind Eng Chem Res 1999;38:3932. [45] Fernandez-Jiminez AM, Palomo A, Lopez-Hombrados C. Engineering properties
[11] Xu H, Van Deventure JSV. Geopolymerisation of multiple minerals. Min. Eng. of alkali-activated fly ash concrete. ACI Mater J 2006;103:106–12.
2002;15:1131–9. [46] Olivia M, Nikraz H. Properties of fly ash geopolymer concrete designed by
[12] Zhang B, MacKenzie KJD, Brown IWM. Crystalline phase formation in taguchi method. Mater Des 2012;36:191–8.
metakaolinite geopolymers activated with NaOH and sodium silicate. J [47] Singh A. Engineering properties of reinforced geopolymer concrete [B. Tech
Mater Sci 2009;44:4668–76. project report]. Roorkee, India. CSIR-Central Building Research Institute; 2012.
[13] Ishwarya G. Development of geopolymer concrete cured at ambient [48] Sofi M, Van Deventer JSJ, Mendis PA, Lukey GC. Engineering properties of
temperature [M. Tech thesis]. Roorkee, India: CSIR-Central Building Research inorganic polymer concretes. Cem Concr Res 2007;37:251–7.
Institute (AcSIR); 2013. [49] Diaz-Loya EI, Allouche EN, Vaidya S. Mechanical properties of fly ash-based
[14] Silva PD, Crenstil KS, Sirivivatnanon V. Kinetics of geopolymerization: role of geopolymer concrete. ACI Mater J 2011;108:300–6.
Al2O3 and SiO2. Cem Concr Res 2007;37:512–8. [50] Ryu GS, Lee YB, Koh KT, Chung YS. The mechanical properties of fly ash-based
[15] de Vargas AS, DalMolin DCC, Vilela ACV, da Silva FJ, Pavão B, Veit H. The effects geopolymer concrete with alkaline activators. Constr Build Mater
of Na2O/SiO2 molar ratio, curing temperature and age on compressive 2013;47:409–18.
strength, morphology and microstructure of alkali-activated fly ash-based [51] Bondar D, Lynsdale CY, Milestone NB, Hassani N, Ramezanianpour AA.
geopolymers. Cem Concr Comp 2011;33:653–60. Engineering properties of alkali-activated natural pozzolan concrete. ACI
[16] Yip CK, Luckey GC, Provis JL, van Deventure JSV. The coexistence of Mater J 2011;108:64–72.
geopolymeric gel and calcium silicate hydrate at the early stage of alkaline [52] Yost JR, Radlinska A, Ernst S, Salera M. Structural behavior of alkali activated
activation. Cem Concr Res 2005;35:1688–97. fly ash concrete. Part 1: mixture design, material properties and sample
[17] Somna K, Jaturapitakkul C, Kajitvichyanukul P, Chindaprasirt P. NaOH- fabrication. Mater Struct 2012;46:435–47.
activated ground fly ash geopolymer cured at ambient temperature. Fuel [53] Pan Z, Sanjayan JG, Rangan BV. Fracture properties of geopolymer paste and
2011;90:2118–24. concrete. Mag Concr Res 2011;63:763–77.
[18] Yip CK, Luckey GC, Provis JL, van Deventure JSV. Effect of calcium silicate [54] Sarker PK, Haque R, Ramgolam KV. Fracture properties of heat cured fly ash-
sources on geopolymerisation. Cem Concr Res 2008;38:554–64. based geopolymer concrete. Mater Des 2013;44:580–6.
[19] Kumar S, Kumar R, Melhotra SP. Influence of granulated blast furnace slag on [55] Mehta PK, Monteiro PJM. Concrete: microstructure, properties and
the reaction, structure and properties of fly ash-based geopolymer. J Mater Sci materials. USA: The McGraw Hill companies, Inc.; 2006.
2010;45:607–15. [56] Lee WKW, Van Deventure JSJ. The interface between natural siliceous
[20] Puligilla S, Mondal P. Role of slag in microstructural development and aggregates and geopolymers. Cem Concr Res 2004;34:195–206.
hardening of fly ash-slag geopolymer. Cem Concr Res 2013;43:70–80. [57] Lee WKW, Van Deventure JSJ. Chemical interactions between siliceous
[21] Chindaprasirt P, Silva PD, Crentsil KS, Hanjitsuwan S. Effect of SiO2 and Al2O3 aggregates and low-Ca alkali-activated cements. Cem Concr Res
on the setting and hardening of high calcium fly ash-based geopolymer 2007;37:844–55.
systems. J Mater Sci 2012;47:4876–83. [58] Zhang YS, Sun W, Li JZ. Hydration process of interfacial transition in potassium
[22] Guo X, Shi H, Dick WA. Compressive strength and microstructural polysialate (K-PSDS) geopolymer concrete. Mag Concr Res 2005;57:33–8.
characteristics of class C fly geopolymer. Cem Concr Comp 2010;32:142–7. [59] Demie S, Nuruddin MF, Shafiq N. Effects of micro-structure characteristics of
[23] Kusbiantoro A, Ibrahim MS, Muthusamy K, Alias A. Development of sucrose interfacial transition zone on the compressive strength of self-compacting
and citric acid as natural based admixture for fly ash based geopolymer. Proc geopolymer concrete. Constr Build Mater 2013;41:91–8.
Environ Sci 2013;17:596–602. [60] ACI 408R-03. Bond and development of straight reinforcing bars in tension.
[24] Nemotallahi B, Sanjayan JG. Effect of different superplasticizer and activator [61] Sarker PK. Bond strength of reinforcing steel embedded in fly ash-based
combinations on workability and strength of fly ash based geopolymer. Mater geopolymer concrete. Mater Struct 2011;44:1021–30.
Des 2014;57. 667-67. [62] Sofi M, Van Deventer JSJ, Mendis PA, Lukey GC. Bond performance of
[25] Jang JG, Lee NK, Lee HK. Fresh and hardened properties of alkali-activated reinforcing bars in inorganic polymer concrete (IPCs). J Mater Sci
fly ash/slag pastes with superplasticizers. Constr Build Mater 2007;42:3107–16.
2014;50:169–76. [63] Sumajouw DMJ, Hardjito D, Wallah SE, Rangan BV. Fly ash-based geopolymer
[26] Palomo A, Grutzeck MW, Blanco MT. Alkali-activated fly ashes: a cement for concrete: study of slender reinforced columns. J Mater Sci 2007;42:3124–30.
the future. Cem Concr Res 1999;29:1323–9. [64] Dattatreya JK, Rajamane NP, Ambily PS. Structural behaviour of reinforced
[27] Perera DS, Uchida O, Vance ER, Finnie KS. Influence of curing schedule on the geopolymer concrete beams and columns. CSIR-SERC Research Report RR-
integrity of geopolymer. J Mater Sci 2007;42:3099–106. 6;2009.
[28] Heah CY, Kamarudin H, Al Bakri AMM, Binhussain M, Luqman M, Nizar IK, et al. [65] Yost JR, Radlinska A, Ernst S, Salera M. Structural behaviour of alkali activated
Effect of curing profile on Kaolin-based geopolymers. Phys Procedia fly ash concrete. Part 2: structural testing and experimental findings. Mater
2011;22:305–11. Struct 2012;46:449–62.
[29] Rovnanik P. Effect of curing temperature on the development of hard structure [66] Ng TS, Amin A, Foster SJ. The behaviour of steel-fibre-reinforced geopolymer
of metakaolin-based geopolymer. Constr Build Mater 2010;24:1176–83. concrete beams in shear. Mag Concr Res 2013;65:308–18.
[30] Kani EN, Allahverdi A. Effects of curing time and temperature on strength [67] Neville AM. Properties of concrete. 4th ed. India: Dorling Kindersley
development of inorganic polymeric binder based on natural pozzolan. J Mater Publishing, Inc.; 1997.
Sci 2009;44:3088–97. [68] Kong DLK, Sanjayan JG, Crentsil KS. Comparative performance of geopolymers
[31] Van Jaarsveld JGS, Van Deventer JSJ, Lukey GC. The effect of composition and made with metakaolin and fly ash after exposure to elevated temperature.
temperature on the properties of fly ash and kaolinite based geopolymers. Cem Concr Res 2007;37:1583–9.
Chem Eng J 2002;89:63–73. [69] Kong DLK, Sanjayan JG, Crentsil KS. Factors affecting the performance of
[32] Chindaprasirt P, Chareerat T, Sirivivatnanon V. Workability and strength of metakaolin geopolymers exposed to elevated temperatures. J Mater Sci
coarse high calcium fly ash geopolymer. Cem Concr Comp 2007;29:224–9. 2008;43:824–31.
[33] Chindaprasirt P, Chareerat T, Hatanaka S, Cao T. High strength geopolymer [70] Kong DLY, Sanjayan JG. Effect of elevated temperatures on geopolymer paste,
using fine high-calcium fly ash. J Mater Civ Eng 2011;23:264–70. mortar and concrete. Cem Concr Res 2010;40:334–9.
[34] Temuujin J, van Riessen A, Mackenzie. Preparation and characterisation of fly [71] Zhao R, Sanjayan JG. Geopolymer and Portland cement concretes in simulated
ash-based geopolymer mortars. Constr Build Mater 2010;24:1906–10. fire. Mag Concr Res 2011;63:163–73.
[35] Khandelwal M, Ranjith PG, Pan Z, Sanjayan JG. Effect of strain rate on strength [72] Swamy RN, editor. The alkali-silica reaction in concrete. UK: Blackie and Sons
properties of low-calcium fly ash-based geopolymer mortar under dry Limited; 1992.
condition. Arabian J Geosci 2013;6:2383–9. [73] Patil KK, Allouche EN. Impact of alkali silica reaction on fly ash-based
[36] Hardjito D, Fung SS. Fly ash-based geopolymer mortar incorporating bottom geopolymer concrete. J Mater Civ Eng ASCE 2013;25:131–9.
ash. Mode Appl Sci 2010;4:44–52. [74] Fernandez-Jimenez A, Puertas F. The alkali-silica reaction in alkali-activated
[37] Sathonsaowaphak A, Chindaprasirt P, Pimraksa K. Workability and strength of granulated slag mortars with reactive aggregate. Cem Concr Res
lignite bottom ash geopolymer mortar. J Hazard Mater 2009;168:44–50. 2002;32:1019–24.
90 B. Singh et al. / Construction and Building Materials 85 (2015) 78–90

[75] Garcia-Lodeiro I, Palomo A, Fernandez-Jimenez A. The alkali-aggregate [85] Bhattacharyya SK, Singh B. High performance materials and construction
reaction in alkali activated fly ash mortars. Cem Concr Res 2007;37:175–83. technologies for sustainable built space. Supra Institutional Project Report (SIP
[76] Singh B, Ishwarya G, Gupta M, Bhattacharyya SK. Performance evaluation of 29), CSIR-Central Building Research Institute, Roorkee, India; 2012.
geopolymer concrete through alkali-silica reaction. In: Advances in chemically [86] Singh B, Sharma S, Gupta M, Bhattacharyya SK. Performance of fly ash-based
activated materials, Changsha, China; Jun 1–3, 2014. geopolymer pastes under chemical environment. In: International conference
[77] Temuujin J, Minjigmaa A, Lee M. Characterisation of class F fly ash geopolymer on advances in construction materials through science and engineering, Hong
pastes immersed in acid and alkaline solutions. Cem Concr Comp Kong; 5–7 September, 2011.
2011;33:1086–91. [87] Rajamane NP, Natraja MC, Dattatreya JK, Lakshmanan N, Sabitha D. Sulphate
[78] Bakharev T. Resistance of geopolymer materials to acid attack. Cem Concr Res resistance and eco-friendliness of geopolymer concretes. Ind Concr J
2005;35:658–70. 2012;86:13–21.
[79] Ariffin MAM, Bhutta MAR, Hussin MW, Mohd Tahir M, Aziah N. Sulfuric acid [88] Bernal SA, Gutierrez RM, Provis JL. Engineering and durability properties of
resistance of blended ash geopolymer concrete. Constr Build Mater concrete based on alkali-activated granulated blast furnace slag/metakaolin
2013;43:80–6. blends. Constr Build Mater 2012;33:99–108.
[80] Sathia R, Ganesh Babu K, Santhanam M. Durability study of low calcium fly ash [89] Reddy DV, Edouard JB, Sobhan K, Tipni A. Experimental evaluation of the
geopolymer concrete. In: 3rd ACF international conference, Ho chi minh city, durability of fly ash-based geopolymer concrete in the marine environment.
Vietnam; 2008. In: 9th Latin American and Caribbean conference on engineering for a smart
[81] Bakharev T, Sanjayan JG, Cheng Y-B. Resistance of alkali-activated slag planet, innovation, information technology and computational tools for
concrete to acid attack. Cem Concr Res 2003;33:1607–11. sustainable development Colombia, Australia; 2011.
[82] Fernandez-Jimenez A, Garcia-Lodeiro I, Palomo A. Durability of alkali-activated [90] Singh B, Gupta M, Chauhan M, Bhattacharyya SK. Lightweight geopolymer
fly ash cementitious material. J Mater Sci 2009;42:3055–65. concrete with EPS. In: CIB World Building Congress 2013, Brisbane, Australia;
[83] Bakharev T. Durability of geopolymer materials in sodium and magnesium 5–9 May, 2013.
sulfate solutions. Cem Concr Res 2005;35:1233–46.
[84] Ismail I, Bernal SA, Provis JL, Hamdan S, van Deventer JSJ. Microstructural
changes in alkali activated fly ash/slag geopolymers with sulphate exposure.
Mater Struct 2013;46:361–73.

Вам также может понравиться