You are on page 1of 16

Minerals Engineering 29 (2012) 89–104

Contents lists available at SciVerse ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Technical and commercial progress in the adoption of geopolymer cement


Jannie S.J. van Deventer a,b,⇑, John L. Provis a, Peter Duxson b
a
Department of Chemical & Biomolecular Engineering, University of Melbourne, Victoria 3010, Australia
b
Zeobond Pty. Ltd., P.O. Box 210, Somerton, Victoria 3062, Australia

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

Article history: If formulated optimally, geopolymer cement made from fly ash, metallurgical slags and natural pozzolans
Available online 12 October 2011 could reduce by 80% the CO2 emissions associated with the manufacturing of cement. However, almost all
standards and design codes governing the use of cementitious binders and concrete in construction are
Keywords: based on the use of Portland cement. The 100+ year track record of in-service application of Portland
Fine particle processing cement is inherently assumed to validate the protocols used for accelerated durability testing. Moreover,
Environmental the entire supply chain associated with cementitious materials is based on the production of Portland
Recycling
cement. The geopolymerisation of aluminosilicates constitutes a radical change in construction materials
Waste processing
Particle size
chemistry and synthesis pathways compared with the calcium silicate hydrate chemistry which under-
pins Portland cement. Consequently, there are regulatory, supply chain, product confidence and technical
barriers which must be overcome before geopolymer cement could be widely adopted. High profile dem-
onstration projects in Australia have highlighted the complex regulatory, asset management, liability and
industry stakeholder engagement process required to commercialise geopolymer cement. While the
scale-up from the laboratory to the real-world is technically challenging, the core challenge is the
scale-up of industry participation and acceptance of geopolymer cement. Demand pull by a carbon con-
scious market continues to be the key driver for the short term adoption of geopolymer cement. In the
absence of an in-service track record comparable in scale and longevity to Portland cement, research is
essential to validate durability testing methodology and improve geopolymer cement technology. Colloid
and interface science, gel chemistry, phase formation, reaction kinetics, transport phenomena, comminu-
tion, particle packing and rheology, which are familiar concepts to minerals engineers, are also key build-
ing blocks in the development of geopolymer knowledge. Analysis of the nanostructure of geopolymer
gels has enabled the tailored selection of geopolymer precursors and the design of alkali activator com-
position, aiding in establishing the relationship between geopolymer gel microstructure and durability.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction safe level (Richardson et al., 2009). In contrast, there remain many
sceptics, including in the minerals industry, who do not accept that
Under the 2009 Copenhagen Accord, more than 120 countries climate change is a result of human activity (Plimer, 2009). This pa-
have agreed to keep global average temperature increase below per does not aim to contribute to such debate, but instead assumes
2 °C. This maximum acceptable temperature increase is based on that there is economic and social benefit in reducing CO2 emissions
recommendations from numerous scientific studies, warning that and valorising waste materials at the same time, and develops dis-
increases in excess of 2 °C can trigger dangerous anthropogenic cussions from this viewpoint.
interference with the climate system, including climate-tipping Concrete made from Ordinary Portland Cement (OPC), including
points, with unmanageable consequences to water supply, agricul- its blends with mineral admixtures, is second only to water as the
tural productivity, sea-level rise, human habitability, and global commodity most used by mankind today. Global cement produc-
security (Rockström et al., 2009). This is an urgent goal, because tion in 2008 was around 2.6 billion tonnes (Freedonia Group,
many scientists estimate that the concentrations of CO2 and other 2009), contributing conservatively 5–8% of global anthropogenic
climate-forcing substances in the atmosphere already exceed the CO2 emissions (WWF-Lafarge Conservation Partnership, 2008;
Scrivener and Kirkpatrick, 2008), and the rapidly increasing de-
mand for advanced civil infrastructure in China, India, the Middle
⇑ Corresponding author at: Department of Chemical & Biomolecular Engineering,
East and the developing world is expected to expand the cement
University of Melbourne, Victoria 3010, Australia. Tel.: +61 3 9303 7777; fax: +61 3
and concrete industries significantly (Taylor et al., 2006). CO2 emis-
9303 7644.
E-mail address: jannie@zeobond.com (J.S.J. van Deventer). sions are due mainly to the decomposition of limestone and

0892-6875/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2011.09.009
90 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

combustion of fossil fuels during cement production; grinding and ing synchrotron-based and other nanostructural characterisation
transport are also significant contributors to the environmental of geopolymers to durability and engineering properties, identify-
footprint of the cement industry. ing the technical, commercial and regulatory barriers to industrial
The cement industry has started to acknowledge the role of adoption, and reviewing progress made in Australia along the path
alternative binders in a carbon constrained industry, given that to commercialisation. This will be addressed from a joint research-
there are significant reductions in CO2 emissions and also some commercialisation viewpoint, with a particular focus on the areas
advantages in performance only offered by these alternative bind- of research which are specifically beginning (or continuing) to im-
ing systems (Damtoft et al., 2008). Historically, the driver for com- pact developments in the commercial arena. The underlying
petition in the construction materials industry has been cost assumption in research papers and grant applications is that good
reduction, in which case alternative binders, starting from a low research will necessarily lead to adoption in industry, which is far
volume basis, could never compete against large-scale OPC produc- from the situation in reality. This paper will use an experiential and
tion. However, CO2 abatement and technical features are now play- tutorial style to show that the interplay between research problem
ing a major role in growth of alternative binder systems. Juenger identification, technical development work and commercial strat-
et al. (in press) recently presented a review of potential alterna- egy is as important as high-quality fundamental research, and that
tives to OPC technology, including calcium sulphoaluminate ce- both are required in the implementation of a new class of binders
ments, magnesium cements, the magnesium phosphate system on a large scale.
and alkali activated materials (AAM) or geopolymers.
Calcium sulphoaluminate cements are made from clinkers that
include ye’elimite as a primary phase, which requires a lower
2. The role of chemical research in the commercialisation of
amount of limestone and lower fuel consumption. Their commer-
geopolymers
cial use in expansive cements and ultra-high early strength ce-
ments has been pioneered in China. Magnesium-based cements
The commercial implementation of geopolymer technology in
and magnesium phosphate cements have been used in niche appli-
Australia is currently being driven by multiple teams operating
cations and can also give superior fire resistance, with much lower
in different parts of the country. Curtin University and the Centre
CO2 emissions than OPC. The use of magnesium silicate hydrates
for Sustainable Resource Processing, based in Perth, have made ad-
together with magnesium carbonate as a ‘carbon negative con-
vances in this area over the past decade and conducted a number
crete’ is also attracting commercial attention at present (Gartner
of trial pours in recent years. The Melbourne-based company, Zeo-
and Macphee, 2011). However, many of these alternative binders
bond, developed its own pilot-scale production facilities in 2007
require a new supply chain for raw materials, the development
and now supplies its concrete, E-Crete™, to major civil infrastruc-
of new chemical admixtures, regulatory approval, the development
ture projects including freeway expansion works and bridge con-
of new durability testing protocols, and an in-service track record
struction and repairs under licence. E-Crete™ utilises for its
before they would be adopted widely by industry.
binder a blend of fly ash and ground blast furnace slag, with com-
Geopolymer cement faces the same obstacles, but has a longer
binations of proprietary alkaline activating components which are
in-service track record (Shi et al., 2006; Van Deventer et al.,
tailored for specific raw materials and products.
2010a; Xu et al., 2008), supported by an expanding body of funda-
There are various technical and commercial factors driving the
mental research relating gel chemistry and nanostructure to dura-
commercial adoption of geopolymer technology, and demand pull,
bility. In geopolymer chemistry, the reactive aluminosilicate
led by a carbon-conscious end-user market, continues to be the key
phases present in precursors including fly ash from coal combus-
driver for the short term adoption of geopolymer concrete. Coun-
tion, metallurgical (including blast furnace) slag, calcined clays
ter-balancing this demand driver is the inherent resistance of the
and/or volcanic ash are reacted with alkaline reagents including al-
civil construction industry to the large-scale adoption of new prod-
kali metal silicates, hydroxides, carbonates, and/or aluminates
ucts, where time and demonstration on relevant industrial scales
(Provis and Van Deventer, 2009) to form aluminosilicate gel phases
are prerequisites for practical credibility, as well as the cost impli-
with varying (but generally low) degrees of crystalline zeolite for-
cations of the absence of economies of scale.
mation. Geopolymer concrete has been widely reported to display
A detailed chemical understanding of the properties of geopoly-
high resistance to fire and acids, and does not produce the high
mer binders, in particular in areas such as control of setting time,
evolution of reaction heat associated with OPC concrete, reducing
workability and durability, also plays an enabling role in the com-
cost and potential cracking issues when the material is placed in
mercialisation process. There is a growing volume of scientific lit-
large volumes. The benefits of geopolymerisation when compared
erature exploring the properties of geopolymeric materials on the
with OPC technology are largely based around the ability to valo-
laboratory scale. Unfortunately, much of this information has lim-
rise high-volume industrial waste streams into high-performance
ited direct value in commercial adoption; geopolymer concrete
concretes, with a highly significant reduction in CO2 emissions
that performs adequately according to all standards can be syn-
(Duxson et al., 2007a). Fly ash and slag appear at present to be
thesised easily in a laboratory, while it is much more difficult to
the most promising precursors for large-scale industrial produc-
reproduce such performance in a commercially and practically fea-
tion of geopolymer cement due to the more favourable rheological
sible form in real-world applications. Indeed, closing this apparent
properties and lower water demand achievable when compared to
disparity and gap between the laboratory and the real-world has
mixes based on calcined clays (Provis et al., 2010).
been the focus of much research conducted by the authors. Being
The history, chemical principles, reaction phenomena and engi-
able to achieve this challenging goal has thus unlocked the com-
neering properties of geopolymer concrete have been reviewed
mercial value of geopolymer technology.
extensively (Davidovits, 2009; Duxson et al., 2007b; Komnitsas
and Zaharaki, 2007; Pacheco-Torgal et al., 2008a,b; Provis and
Van Deventer, 2009; Shi et al., 2011). With the core focus of these
recent reviews being laboratory research on geopolymer binders 3. Developments in geopolymer gel phase chemistry
rather than the production of concrete, it is not the aim of this re-
view paper to duplicate such analysis or discussion. Instead, this This section summarises selected research advances in geopoly-
paper will fill a gap in the current literature by explaining the rel- mer gel chemistry and also identifies topics that require further
evance of particle technology in geopolymer concrete design, link- work, as depicted in Fig. 1.
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 91

Clearly, there is a need for a workable one-part (‘just add


water’) mix design if geopolymeric cements and concretes are to
achieve widespread market penetration, as this would greatly sim-
plify the process of handling and distributing the components of a
geopolymer binder compared to the use of two-part systems
involving large volumes of alkaline liquids. Various methodologies
for designing the necessary precursors have been proposed
(Koloušek et al., 2007; Duxson and Provis, 2008; O’Connor and
MacKenzie, 2010; Van Deventer et al., 2010b), but its successful
commercial implementation still remains to be demonstrated.

3.2. Binder phase chemistry

By studying the remnant fly ash and slag particles left embed-


ded in a hardened geopolymer binder using scanning electron
microscopy (SEM), Lloyd et al. (2009a,b) observed that calcium is
active during alkali activation, and that, if discrete high calcium
binder regions are formed (rather than the calcium being incorpo-
rated into the main aluminosilicate gel binder), it is on a nanome-
tre rather than a micron length scale. Yip et al. (2005, 2008)
suggested that both geopolymeric (alkali aluminosilicate) gel and
calcium silicate hydrate (C–S–H) gel co-exist at low alkalinities,
while geopolymeric gel appeared to be the dominant product at
high alkalinities, while Buchwald et al. (2007) also showed gel
coexistence at relatively high alkalinity. Recent work from
García-Lodeiro et al. (2011) shows different trends with respect
to alkalinity, with C–A–S–H gel proposed to be more stable than
N–A–S–H at high pH and further work is certainly required in this
area to resolve the remaining disagreements. Provis et al. (2009)
Fig. 1. Conceptual diagram showing geopolymer gel phase processes. did identify discrete Ca-rich regions within a hydroxide-activated
Class F fly ash binder using 80 nm-resolution X-ray fluorescence
microscopy; these regions were not observed in silicate-activated
3.1. Precursor design samples, and were proposed to be related to the precipitation of
Ca(OH)2 in poorly crystalline form during the very early stages of
Based on data obtained from the literature, Duxson and Provis reaction of the fly ash particles in highly alkaline environments,
(2008) proposed an ideal composition range for glassy aluminosil- in contrast to its wider distribution throughout the gel formed at
icate precursors containing network-modifying cations (particu- lower alkalinity.
larly calcium, magnesium, sodium and potassium) in order to Using the results of synchrotron X-ray diffraction, Oh et al.
give sufficiently high solubility to supply the necessary aluminium (2010) proposed that the geopolymeric gel contains zeolitic pre-
into the growing geopolymer gel. This control of aluminium avail- cursors related to disordered forms of the ABC-6 family of zeolites,
ability has been highlighted as being necessary to enable the depending on the pH environment of the paste. Despite a high cal-
growth of geopolymer gel particulates, leading to cross-linking, cium content, only very weak C–S–H(I) peaks and no Ca(OH)2
hardening and strength development (Fernández-Jiménez et al., phase were found in their Class C fly ash paste when activated with
2006; Provis and Van Deventer, 2007). The need for addition of a sodium hydroxide solution. This implies that the calcium in class C
separate alkali source may be either greatly reduced or even elim- fly ash did not dissolve in the activator solution as readily as the
inated if the correct glass, or combination of glasses, can be selec- calcium in the slag; slag-based systems yielded much higher levels
tively synthesised (Buchwald and Wiercx, 2010). Duxson and of crystalline C–S–H(I). The high calcium content in the Class C fly
Provis (2008) postulated that this may be achieved by addition of ash seemed to reduce the strength of the matrix, while the calcium
components into pulverised coal prior to combustion, or by the in the slag appeared to increase strength. Oh et al. (2010) postu-
selective manufacture of a highly reactive raw material (Van lated that the calcium in the slag is available to form C–S–H(I)
Deventer et al., 2010b) that can be blended with less-reactive while it is not available in the Class C fly ash ‘‘due to the different
raw materials to provide a geopolymer cement to a given chemical forms of calcium in these raw materials.’’ In summary,
specification. the precise role of calcium in geopolymer formation remains
Keyte (2008) determined that the aluminosilicate glass particles poorly defined, although it is clearly pivotal in determining the
present within Class F fly ashes predominantly consisted of two engineering properties and durability of the product.
intimately intermixed amorphous phases, one silica-rich and one Iron is also likely to be important; Lloyd et al. (2009a) observed
very alumina-rich, with composition close to Al6Si2O13 and resem- that a high iron content appears to render precursor particles rel-
bling mullite. By applying devitrification to geopolymers, Keyte atively unreactive, and noted that phase segregation between
(2008) showed that the poor performance of certain fly ashes in iron-rich and iron-poor glasses within particles can mean that
geopolymer synthesis is related to the low aluminium content of compositional information obtained even at an individual particle
the amorphous phases present in the precursor, resulting in a high level is not necessarily able to describe reactivity in geopolymer
silicon to aluminium ratio in the formed geopolymer. The coordi- formation. During alkali activation, iron does not appear to move
nation environment of aluminium in geopolymer precursors is a much from its original position within fly ash particles (Lloyd
complex issue, and plays a key role in controlling the availability et al., 2009b; Provis et al., 2009). However, the effect of iron on
of Al in geopolymerisation; this is an area of active research and the lability of precursor glassy phases and the role that iron may
debate (White et al., 2011). play during geopolymer reactions remain largely undescribed;
92 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

Mössbauer spectroscopy does appear to be giving some initial an- activated Class F fly ash system selective leaching of Al from the
swers in this area, at least for the case of very Fe-rich precursors fly ash produces in the first instance a loose, Al-rich, ‘primary’
(Gomes et al., 2010). gel, as proposed earlier by Fernández-Jiménez and Palomo
The mechanistic distinction between silicate-activated and (2005). The Al-deficient surface layer on the fly ash particles then
hydroxide-activated geopolymer formation can be attributed to dissolves, followed by later stoichiometric release of Al and Si spe-
some extent to the differences in the sites at which gel precipita- cies. During an induction period, the gel slowly comes to pseudo-
tion takes place, with hydroxide-activated gels forming predomi- equilibrium with the surrounding solution via depolymerisation/
nantly on fly ash particle surfaces rather than by polymerisation repolymerisation reactions (White et al., in press b). Gel nuclei
in the bulk region (Lloyd et al., 2009b). In silicate-activated fly (particles that are sufficiently stable to resist depolymerisation)
ash, the gel consisted of roughly similar colloidal-sized, globular begin to form, and the growth of a new gel phase begins. This
units closely bonded together at their surfaces. The gel appeared new gel is the phase predominantly responsible for strength devel-
to be homogenous throughout the sample, whether at the surface opment and durability in geopolymers.
of the ash particles or relatively far away in the interstitial space. It When nanoparticle seeds are added to the geopolymer system,
is, however, possible to influence this through nanoparticle seeding no induction period occurs, as the nanoparticles immediately cata-
of hydroxide-activated binders to manipulate the nucleation loca- lyse the formation of nuclei (Rees et al., 2008). In this case, when
tion and enhance gel growth (Rees et al., 2008). the first Al-rich species are released into the solution from the fly
The reason why this geopolymer gel is formed instead of large ash particles, their immediate addition to the nuclei forms an Al-
zeolitic crystallites is related to the degree of polymerisation of sil- rich gel. Structural reorganisation of this gel will later lead to the
icate species, as well as the influence of temperature and water formation of zeolites, which can differ in crystal structure from
content (Provis et al., 2005; Provis and Van Deventer, 2007, the zeolites which develop in unseeded systems (Rees et al.,
2009; Fernández-Jiménez et al., 2008). White et al. (in press a) used 2008). Dissolution of the remnant siliceous layer on the fly ash par-
neutron pair distribution function (PDF) analysis to show that, ticles releases Si-rich species, which also add rapidly to the grow-
with increasing reaction time, the geopolymer gel derived from ing nuclei, forming a high-silica gel region. Congruent ash
metakaolin transitions to a more ordered state via an increase in dissolution then releases Al and Si species which add to the grow-
cross-linking. Following the initial nuclear magnetic resonance ing nuclei in a similar way to the unseeded system, creating a bulk
(NMR) and thermal analysis data of Duxson et al. (2005a) and geopolymer gel with similar composition and structure to that ob-
Duxson et al. (2007c), a neutron PDF study of heated geopolymer served in the absence of seeds. Hajimohammadi et al. (2010,
gels (White et al., 2010) confirmed that the water in metakaolin 2011a,b) then followed this work by applying synchrotron radia-
geopolymers was present mainly as free water in large pores, with tion-based FTIR microscopy to the spatially-resolved analysis of
only a small percentage (approximately <5%) of water either phys- both one-part and two-part geopolymer systems, showing that
ically bound in small pores or chemically bound as hydroxyl the release rates of both Si and Al are critical in determining
groups attached to the aluminosilicate framework structure. strength development and microstructural evolution in growing
Coarse-grained Monte Carlo simulations, based on interaction geopolymer gels. The enhancement of the nucleation process
energies derived from density functional modelling, showed that, through seeding led to the formation of an additional silica-rich
as the activator silica content of a metakaolin geopolymer system phase in the early stages of the reaction, which left more Al avail-
is increased, more species (both aluminate and silicate) participate able to contribute to bulk gel formation and improved the early
in geopolymerisation, leading to a denser nanostructure and less strength development of geopolymer binders. The later release of
monomeric species existing in the pore solution (White et al., in more silica from the Si-rich phase then also enhanced final
press b). The precipitates formed become larger with increasing sil- strength.
ica content, which indicates that different structural transforma-
tion mechanisms occur depending on the type of activator used. 4. Role of particle technology in the optimisation of geopolymer
The results of these simulations have confirmed a previously pro- paste and concrete
posed hypothesis (Provis et al., 2005) regarding nucleation of pre-
cipitates in metakaolin-based systems, where in the presence of Fig. 2 depicts a conceptual model for the interrelationship be-
silicate activating solutions the early release of aluminium from tween geopolymer concrete mix design, the behaviour of wet con-
metakaolin was suggested to lead to localised nucleation close to crete, and the performance of in-service concrete regarding
the surface of the partially dissolved metakaolin particles; this engineering properties and durability. In this section the principles
behaviour was reflected in the simulation results. On the other of particle technology will be used to analyse the behaviour of
hand, in hydroxide-activated systems there is no localised nucle- fresh geopolymer cement paste and wet concrete, while in Section
ation taking place, and therefore the precipitates form throughout 5 recent progress on the interrelationship between binder micro-
the system. This differs from the nucleation behaviour observed in structure and durability will be reviewed.
fly ash systems, as outlined above; the aluminium release rates and
dissolution mechanisms differ significantly between the two 4.1. Particle shape effects in fresh pastes
systems, and this leads to differences in the influence of dissolved
silica in the nucleation processes. The simulation results for The grinding of clinker and blast furnace slag into cementitious
silicate-activated systems also provide direct evidence of Ostwald powders results in non-spherical particle morphologies. It has been
ripening of geopolymer gel particles, which has never before been hypothesised that the spherical particle shape of fly ash reduces vis-
explicitly shown to occur in geopolymer systems (White et al., in cosity and yield stress of fresh paste when fly ash is added to OPC
press b). (Laskar and Talukdar, 2008). Provis et al. (2010) used a packing
These differences in particle dissolution mechanisms, and their model to demonstrate the ‘ball-bearing’ effect of spherical ash par-
influence on geopolymer gel development, are essential in deter- ticles in a paste, by the reduction of particle interlocking when the
mining the processes which control the rate and location of geo- paste is sheared. Palomo et al. (2005) observed that ash chemistry
polymer gel formation, and thus the microstructure and and variability largely affected the rheology of fly ashes activated
performance of the final geopolymer binder. By using Attenuated by sodium hydroxide solutions. Various investigations have also
Total Reflectance Fourier Transform Infrared (ATR-FTIR) spectros- studied the effect of ash fineness on early-age properties of OPC
copy, Rees et al. (2007a,b) observed that in a sodium hydroxide and geopolymer pastes. However, many of these investigations
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 93

Fig. 2. Conceptual diagram showing interrelationships between mix design, gel chemistry, matrix characterisation, engineering properties and durability of geopolymer
concrete.

used milling to obtain different particle size distributions (Bou- 4.2. Water–binder ratio and rheology of geopolymer pastes
zoubaâ et al., 1997, 1999; Kumar et al., 2007), which adds the com-
plication of a change in particle shape as the glassy spherical ash The pore volume of calcium silicate-based (Portland cement),
particles are shattered into fragments during the milling process. aluminate-based (high-alumina cement) or sulphate-based (super-
Kumar et al. (2007) showed that slightly higher geopolymer sulphated or calcium sulphoaluminate cement) binders decreases
strengths were achievable by use of the finer particle size fractions significantly as they hydrate over the first few days to weeks after
obtained by classification or attrition milling, when compared to mixing, with this reduction in pore volume attributable to the
the use of raw fly ash, but that vibratory milling was able to give effective consumption of water as it participates in hydration reac-
an improvement in strength of as much as 50%. These effects could tions. A water to binder (w/b) ratio of approximately 0.2, where the
not be attributed solely to particle size reduction, given that the exact value depends on the cement composition (Brouwers, 2004),
vibratory-milled fly ash was not the finest of the ash samples stud- is converted during the hydration of OPC to ‘non-evaporable’ water
ied, and a mechanochemical activation process was postulated to in hydration products, in addition to the considerable amount of
enhance reactivity. In contrast, Keyte (2008) did not show a signif- water bound in ‘gel pores’ (less than 2.7 nm diameter), which is
icant effect of either classification or ring milling in the compres- also not readily removed from the gel. Thus, hydrated OPC with a
sive strength of geopolymers. nominal water/binder ratio of 0.5 has a final pore volume which
Clearly, the effect of particle geometry on the behaviour of fresh is markedly less than the volume of the water which was initially
geopolymer pastes is complex, so there is a need to deconvolute the added into the mix (Lothenbach et al., 2008).
effects of particle shape, particle size and mechanochemical phe- In contrast, geopolymeric binders are primarily aluminosilicate-
nomena and fly ash chemistry. Some detailed studies in this area based and do not form hydrate products, as discussed in Section
have used synthetic aluminosilicate glasses (Hos et al., 2002; Keyte, 3.2. Hence, geopolymers do not have the same pore volume reduc-
2008). However, to fully simulate the geopolymerisation of waste tion benefit through the conversion of water into a solid via its
materials, it is also necessary to incorporate non-framework cations incorporation into reaction products. This distinction has signifi-
into these synthetic glasses to better reflect the main reactive cant implications for the development of geopolymer concrete.
phases participating in alkaline activation (Duxson and Provis, While in geopolymer binders the pore size distribution is much
2008). Some of the knowledge of mixed alkali/alkaline earth alumi- finer, and is greatly refined throughout the hardening process
nosilicate glasses developed recently through the study of synthetic (Duxson et al., 2005b; Sindhunata et al., 2008; Lloyd et al.,
slags (Rajaokarivony-Andriambololona et al., 1990; Shimoda et al., 2009c), the absolute pore volume is not affected by the formation
2008) should also be relevant in future developments in this area. of hydration products and there is no significant amount of water
94 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

bound into the solidified geopolymer gel (Provis et al., submitted Weather-induced mechanisms of attack on the binder and concrete
for publication). Consequently, the minimisation of w/b ratio in (including freeze–thaw damage and/or salt scaling), and alkali-
geopolymer concrete is a particularly significant operational factor aggregate reactions involving some siliceous or carbonate aggre-
in ensuring quality low permeability concrete. gates, can also be problematic under given circumstances and in
some climatic conditions, but the focus of the review presented
here will be on mechanisms of degradation which are explicitly re-
4.3. Particle packing and mix design in geopolymer concretes
lated to ionic transport. The loss of alkalinity (via leaching, carbon-
ation or other mechanisms) and the ingress of chloride are the
Besides particle shape, particle size distribution also has a
primary causes of steel corrosion within concrete, and the resis-
marked effect on the mixing, workability and rheology of wet con-
tance of the material to degradation by these mechanisms will
crete. This is an active area of research in which many fundamental
obviously depend intrinsically on its mass transport properties. In
questions remain unanswered (Hunger and Brouwers, 2009), and
particular, the width, connectivity and degree of water saturation
the development of advanced analytical and simulation techniques
of the pores and cracks which are present throughout the binder
is continuing to provide important advances. Because this is based
will play a highly significant role in determining the resistance to
around the analysis of the locations and interactions of particles on
ingress by aggressive agents. Cracks and pores with relevance to
every length scale from a few nanometres (the basic building units
binder structure and durability performance are present on length
in C–S–H or geopolymer gel) to more than a centimetre (coarse
scales from nanometres to millimetres in most concrete structures,
aggregate), it is not a problem which has an easy or straightfor-
making this another true multiscale problem, similar to the prob-
ward answer. It is very difficult to develop a unified discussion of
lems of mix design and particle packing as discussed above. The
particle–fluid interactions across such a range of length scales, as
application of both traditional and advanced techniques in the anal-
the effects, relative importance and interactions of parameters
ysis of the microstructure of the binder therefore again becomes
such as surface tension, chemical admixtures and solution ionic
imperative to understanding and designing for durability, particu-
strength will differ across length scales (Provis et al., 2010). There-
larly when introducing a new binder system such as the alkali-
fore, most studies consider in detail either the packing of larger
activated family of binders.
particles (particularly the ‘fine’ and ‘coarse’ aggregates; usually
Fig. 3 shows a schematic illustration of some of the physico-
sand and crushed rock) (Amirjanov and Sobolev, 2008; Goltermann
chemical factors which can determine the service life of a concrete;
et al., 1997) or the packing of binder components (Wong and Kwan,
the complexity of this system is immediately apparent from this
2008), with the combination of these two types of components
diagram – which by no means provides an exhaustive listing of
presenting a much more challenging set of experimental and the-
the mechanisms by which concrete can be physicochemically dam-
oretical problems (Jones et al., 2002, 2003).
aged in-service. The discussion in the subsequent sections of this
The term ‘mix design’ in concrete is used mainly to describe the
paper will present a brief overview of selected aspects of the sce-
process of proportioning of binder constituents, chemical and min-
nario depicted in Fig. 3, with a specific focus on the design and
eral admixtures, rock and sand, with the aim of optimising some
characterisation of durable alkali-activated (including geopolymer)
specified combination of technical (mechanical and durability)
binders and concretes.
properties, placement and finishing, and cost (Ji et al., 2006). In
geopolymer concrete mix design, it is important to tailor the con-
tent and composition of the alkali activator and aluminosilicate 5.2. Microcracking phenomena
precursor materials, as well as the aggregate components. This
does not necessarily mean that the change from Portland cement In broad terms, cracking of concretes can be caused by chemo-
to a geopolymer binder requires wholesale changes to overall mechanical (shrinkage or expansion of gel phases, autogenous
mix designs, as the optimisation of the aggregate blends and aggre- heating during curing) or physicomechanical (applied load,
gate-binder ratio is driven by the order of magnitude cost differen- freeze–thaw) processes. Concrete is well known to be strong in
tials between cementitious materials and aggregate. However, compression but weak in flexion and tension; the use of steel
with a fundamental understanding of aluminosilicate chemistry, reinforcing, often in combination with techniques such as preten-
significant improvements in geopolymer concrete quality and
economy can be achieved, particularly with regard to control of Environment Embedded steel
curing conditions, activator dosage and efflorescence. Successful reinforcing
geopolymer concrete mix designs have been published by various Mg2+(aq) Gel binder
researchers; some of these mix designs incorporate chemical Cl-(aq)
admixtures (particularly superplasticisers), although analysis of
the effectiveness of many of the admixtures which are commonly SO42-(aq)
used in Portland cement mixes under the highly alkaline condi- CO2(g)
tions of geopolymerisation has shown mixed outcomes to date H+(aq)
(Puertas et al., 2003; Criado et al., 2009). OH-(aq)
Na+(aq)
5. Linking geopolymer binder structure and durability Ca2+ & Na+ in gel
Aggregate
5.1. Factors affecting the service life of reinforced concrete

The fundamental basis of the durability of reinforced concrete is


the ability to develop and maintain a dense, impermeable binder
gel which creates and stabilises a highly alkaline environment, with
appropriate chemical (in particular electrochemical) conditions to
enable the stabilisation of embedded steel in a passive state. Steel Fig. 3. Schematic diagram of the factors influencing the service life of concrete
corrosion is, on a worldwide basis, the predominant cause of pre- element (based on either alkali-activated or OPC binders) with embedded steel
mature failure of reinforced concrete elements (Hobbs, 2001). reinforcing.
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 95

sioning, and careful structural design, will often be intended spe- pared with ‘bulk’ binder regions, meaning that these specific re-
cifically to compensate for this weakness by ensuring that the con- gions can exert a disproportionately high (and usually negative)
crete itself bears minimal tensile load. However, an advantage of influence on the mass transport, tensile and flexural properties of
geopolymer binders is that they tend to demonstrate a flexural the geopolymer concrete. This is an area in which geopolymer
(and presumably also tensile, although this is rarely tested di- binders are believed to provide significant advantages over OPC.
rectly) strength significantly higher than is specified by the stan- The chemistry of the high-calcium OPC systems tends to lead to
dard relationships which apply for OPC concretes of similar the formation of a porous zone containing large, mechanically
compressive strength (Sofi et al., 2007). Mortar compressive weak crystals surrounding aggregate particles (Ollivier et al.,
strengths in excess of 100 MPa, and concrete strengths higher than 1995; Mehta and Monteiro, 2006), and this is a key pathway for
70 MPa, have been relatively widely reported for laboratory both mechanical failure and mass transport in a concrete. The
samples. Achieving such strengths consistently in large-scale pro- interfacial transition zone in geopolymer concretes has been iden-
duction is more complex, but the large-scale production of ‘high- tified as being dense and much less microstructurally distinct from
performance’ concretes by alkaline activation of suitably chosen the bulk of the binder region (Shi and Xie, 1998; Lee and Van
precursor blends is certainly possible, and is beginning to be dem- Deventer, 2004, 2007; Provis et al., 2007; Zhang et al., 2009), which
onstrated on a commercial scale. provides much higher tensile and flexural performance, as well as
It is also important to distinguish load-induced macroscopic removing the possibility for these regions to lead to the formation
cracking of the concrete from microcracking of the binder; both of a percolated porous pathway for mass transport through the
are likely to be deleterious in terms of durability performance, binder, thus enhancing the durability of the geopolymer material.
but the length scales on which the cracks form are very different. Fig. 4 shows an example, presented in detail by Lee and Van
In an early investigation, Byfors et al. (1989) observed significant Deventer (2007) and reproduced in simplified form here, of the ele-
microcracking in samples of ‘F-concrete’ (superplasticised NaOH/ mental compositions observed across the interface between a (K,
Na2CO3-activated slag). Häkkinen (1992a,b) observed a tendency Na) silicate-activated fly ash geopolymer binder and a quartz par-
towards microcracking in alkali-activated slag concretes, with a ticle within a section of siltstone. The elemental compositions were
corresponding increase in the rate of carbonation and reduction determined by scanning electron microscopy–energy dispersive
in strength, and found that a high activator content tended to cor- spectroscopy (SEM–EDS), and show a uniform binder region span-
relate with a high extent of microcracking. Collins and Sanjayan ning the distance from the bulk binder to the aggregate particle
(2001) conducted a detailed investigation in this area and con- surface. Similar results were also obtained for a range of phases
cluded that drying effects during curing were a primary cause of present within a basalt specimen; an iron-rich augite region in
microcracking of alkali-activated slags. the basalt did generate a distinguishable region less than 2 lm in
More recently, Bernal et al. (2011) proposed, via capillary suc- thickness and rich in Fe and Si at the interface, but there was not
tion measurements, that the extent of microcracking in silicate- a distinct and chemically/crystallographically different region at
activated slag-based concretes depends significantly on the paste the interface in any case when comparing to the bulk binder struc-
content of the concrete mix design; excessive binder content was ture. The SEM–EDS data in Fig. 4 obviously do not provide detailed
suggested to lead to heat generation during curing of concrete microstructural information; however, imaging of the same speci-
specimens, leading to thermally induced microcracking of the con- mens did not show any differences in microstructure or any appar-
crete at early ages. However, in this case, microcracking was not ent additional porosity in these regions (Lee and Van Deventer,
observed to have any significant influence on the rate of carbon- 2007), consistent with the general consensus in the literature
ation of the concretes, indicating that the higher binder content regarding alkali-activated binders.
was able to provide sufficient densification of the matrix to com-
pensate for additional carbonation taking place via transport along
5.4. Microporosity in the bulk of the geopolymer binder
cracks. The fundamental cause of microcracking is the partially re-
strained chemical shrinkage of binder phases after hardening,
There are a variety of techniques available for the analysis of
which has been proposed from a thermodynamic basis to be intrin-
micron-scale pores within a geopolymer (or any cement-type
sic to the chemistry of alkali-activated binders and related (pozzo-
binder), which can be classified roughly into the categories of
lanic) systems (Justnes et al., 1998; Chen and Brouwers, 2007).
one-dimensional, two-dimensional and three-dimensional analy-
Almost all cement-like binders show either shrinkage or expansion
sis. Among these categories, there is no universally applicable tech-
during or after hardening as a result of the process of crystallo-
nique which can provide a full multiscale characterisation of a
graphic (including gel) phase evolution which results in strength
complex material such as a geopolymer binder; a fuller toolkit of
generation and development (Brouwers, 2004). Maintaining
techniques is required to obtain detail across the length scales of
dimensional stability is thus a key challenge across the field of ce-
interest.
ment and concrete technology.
The observed trends in microcracking intensity in alkali-
activated concretes as a function of paste content and curing
regime thus highlight the value of understanding interactions be-
tween the binder and aggregate, and effects related to heat gener-
ation and heat and moisture transport during curing, in mitigating
the effects of microcracking on concrete performance and durabil-
ity. The key interactions take place in the region known as the
‘interfacial transition zone’, and the microstructural characteristics
of this region of the concrete are critical in terms of both strength
and durability performance.

5.3. Interfacial transition zone effects

In the area where the binder comes into contact with aggregate Fig. 4. Elemental concentrations (from scanning electron microscopy–energy
particles, there are often microstructural differences when com- dispersive spectroscopy); data from Lee and Van Deventer (2007).
96 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

One-dimensional analysis of porosity in the range of nanome- tion of Wood’s metal intrusion prior to the collection and analysis
tres to microns is usually conducted by gas sorption (most com- of SEM images has proven to be of value (Lloyd et al., 2009c). In this
monly the Barrett–Joyner–Halenda, BJH, technique), direct technique, the pores of the sample are filled with a high elemental-
penetration measurements (air, water and chloride penetration number, low melting point alloy, which is intruded into the pore
being the most commonly applied), or mercury intrusion porosi- space of the sample under moderate pressure (lower than the pres-
metry (MIP). The limitations and inaccuracies of MIP in application sures used in MIP, to prevent damage to delicate parts of the
to cements are well known (Diamond, 2000), but it is widely used microstructure) and at temperatures below 100 °C but still above
due to its apparent simplicity and the ready availability of instru- the melting point of the alloy. The sample is then cooled, and the
mentation. MIP largely fails in the case of samples with complex alloy solidifies within the pore network, to provide a high degree
pore geometries (the ‘ink-bottle’ effect), as the full volume of a of elemental contrast and make the pores visible against the low-
complex-shaped pore is registered as having an effective pore size elemental number binder regions (Willis et al., 1998; Nemati,
equivalent to that of the narrowest part of its entry. This results in 2000). In the recent application of this technique to alkali-activated
an under-estimation of mean pore size. The high pressures used to binders (Lloyd et al., 2009c), it was calculated that pores as small as
intrude mercury into the smallest pores in a sample are also prob- 11 nm – well below the spatial resolution achievable by standard
lematic, as the applied pressure will often exceed the strength SEM techniques – are able to be filled with the molten alloy, while
(compressive or tensile) of the material to which it is being applied, using an intrusion pressure lower than the compressive strength of
which can result in significant crushing effects. Multi-cycle MIP the binder material to minimise microstructural damage.
has been proposed as a method by which the ink-bottle effect The primary three-dimensional characterisation technique
can be minimised or avoided (Kaufmann et al., 2009), and appears which is able to be applied to the analysis of cements (alkali-acti-
to provide some significant advances in this area with good agree- vated and traditional) is X-ray microtomography. There have been
ment with N2 sorption data. Gas sorption is believed to provide a number of studies of OPC-based materials by this technique over
good characterisation of very small pores, but cannot accurately the past decade or more (Gallucci et al., 2007). The systematic anal-
measure the larger pores which are essential to mass transport ysis of alkali-activated binders by X-ray microtomography and
in samples with pores spanning a very wide range of length scales. hard X-ray nanotomography has recently been presented for the
Direct water and air permeability measurements of geopolymer first time (Provis et al., submitted for publication, 2011a).
and other alkali-activated binders have shown a range of perfor- In the analysis of the alkali-activated binder sample set dis-
mance, depending mainly on the mix designs tested. Binders with cussed here, the samples were analysed using beamline 2-BM at
well-cured alkali-activated binders with low water/binder ratios the Advanced Photon Source synchrotron at Argonne National Lab-
perform acceptably in these tests (Shi, 1996; Olivia et al., 2008; oratory, using 22.5 keV X-ray radiation and achieving a 750 nm
Sagoe-Crentsil et al., 2010), but generally do not provide results voxel resolution. Porosity was calculated during segmentation of
which could be considered particularly outstanding, most likely the sample into ‘pore’ and ‘solid’ regions using slightly modified
due to the low levels of space-filling bound water associated with versions of freely available data processing scripts (Nakashima
these gels. On the other hand, chloride permeability testing of alka- and Kamiya, 2007) based on the full greyscale tomographic recon-
li-activated mortars and concretes has shown a wide range of perfor- structions, and tortuosity was computed by random walker simu-
mance, with the outcomes depending to a significant extent on the lations utilising the segmented pore network, again based on the
details of the testing methodology selected. The ASTM C1202 test algorithms of Nakashima and Kamiya (2007). An example showing
methodology involves the application of a current to the sample the greyscale and segmented images for a given region of a fly ash-
and uses measurement of the charge passed as a proxy for the move- Na silicate geopolymer binder is given in Fig. 5.
ment of chloride ions into the sample under the electrical field gra- Fig. 5 also demonstrates some of the challenges associated with
dient. The outcomes of this test are strongly dependent on pore microtomographic characterisation of fly ash geopolymer samples.
solution chemistry (Liu and Beaudoin, 2000; Shi, 2004), and so it The large dark regions represent the interior of hollow or partially
sometimes registers alkali-activated binders as showing very good hollow (cenosphere or plerosphere) fly ash particles; these regions
resistance to mass transport (Shi, 1996; Roy et al., 2000; Bernal are not connected to the pore volume within the gel, and so will
et al., 2011), and sometimes as performing rather poorly. Direct not contribute to the transport properties of the material. However,
measurement of alkali cation movements from the pore solutions in measurements or calculations of total porosity, these regions
of alkali-activated binders (Lloyd et al., 2010) provides insight into must be excluded via the application of algorithms which considers
the cause of this behaviour. The mobility of alkalis will certainly dis- ‘connected porosity’ only. As an additional complicating factor, the
play a strong influence on the total charge passed by the samples tomographic reconstruction process (if sample alignment was not
during the test. It has been proposed that alternative methodologies on a single vertical axis during rotation for scanning in the instru-
which provide a more direct measurement of the progress of chlo- ment) can introduce streaking artefacts that effectively break
ride migration into the binder – for example ponding tests, or the through the shell of a fly ash particle and connect its interior to the
NordTest Build 492 accelerated test, will provide a more valid com- binder pore network. Thus, the collection of high-resolution, high-
parison which is relatively independent of the pore solution chemis- quality tomography data is of particular importance in enabling
try of the binder, and work in this area is ongoing. the correct analysis of pore network transport parameters for sam-
Two-dimensional analysis of pore structures is predominantly ples containing fly ash. This consideration is less problematic for
conducted using microscopic methods, in particular scanning elec- samples based on slag (or Portland cement) because the precursor
tron microscopy (SEM). The primary challenge associated with the particles do not contain significant inaccessible pore volumes.
observation of porosity by SEM (designed to detect solid matter) is The data collected by Provis et al. (submitted for publication) as
that the aim of the measurement is to detect the regions which shown in Fig. 6, demonstrate that the porosity of the more cal-
contain no solid matter. This process of ‘trying to see the parts cium-rich geopolymer binders (the systems containing 50% or more
which are not there’ provides challenges in pore identification. It slag) decreases as a function of curing duration, indicating that there
can be overcome to some extent for calcium-rich systems, such is some chemical binding of water taking place in this system, con-
as OPC or alkali-activated slags, by the use of image analysis algo- sistent with the known formation and coexistence of sodium alumi-
rithms (Brough and Atkinson, 2000; Wong et al., 2006). However, nosilicate hydrate (‘N–A–S–(H)’) and calcium (alumino)silicate
for aluminosilicate geopolymer systems with lower electron hydrate (‘‘C–(A)–S–H’’) gels in mixed fly ash/slag geopolymer
density contrast between the solid and pore regions, the applica- binders.
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 97

(a) (b)

200 µm

Fig. 5. Binary thresholding of a microtomographic image: (a) original greyscale image, (b) binary segmented image. Scalebar represents 200 lm.

development, an extended period of curing will provide marked


advantages in-service life and overall durability performance once
the binder is placed in-service and exposed to aggressive
environments.
The substitution of fly ash for slag is seen in Fig. 6 to lead to
higher porosity. This is consistent with the results of nitrogen sorp-
tion analysis of a range of similar binders (Lloyd et al., 2009c), and
the porosities obtained from tomography are also within 2 vol.% of
the values obtained in that study for samples of similar mix design
and curing duration. The slag-rich systems predominantly form a
C–(A)–S–H gel, which appears to bind a larger amount of water
into its nanostructure, while the fly ash-rich systems form N–A–
S–(H) gels with a low bound water content (Juenger et al., in press).
The discussion presented here therefore highlights the impor-
tance of understanding and controlling the porosity of geopolymer
binders, particularly in the case of fly ash-rich systems. If the high-
Fig. 6. Relationship between segmented porosity and curing duration for a range of
er porosity of the geopolymer-type (sodium aluminosilicate) gel
sodium silicate-activated fly ash/slag systems; data selected from Provis et al. when compared with calcium silicate gels such as those produced
(submitted for publication). by OPC hydration were to lead to high diffusivity (and thus rapid
mass transport) of aggressive ions through the binder to reach
the embedded steel reinforcing, this would mean that the durabil-
The results of drying tests conducted on a series of mortar sam-
ity of these materials would in all likelihood be unacceptably poor.
ples synthesised from the same binders as were studied by tomog-
Evidence from the in-service performance of geopolymer and other
raphy (Ismail et al., 2011) show further that the C–(A)–S–H phase is
alkali-activated binders (Deja, 2002; Shi et al., 2006; Xu et al.,
binding some water – but to a lesser extent than in the case of OPC
2008), as well as from laboratory chloride penetration testing as
hydrate products. As the gel evolves, and its porosity decreases over
discussed above, shows that the observed performance is signifi-
time, the tortuosity of the pore network also increases. The sample
cantly better than would be expected from raw permeability or
set presented in Fig. 6 shows close agreement with an inverse rela-
carbonation rate data (Shi et al., 2006; Provis et al., 2007; Rodrí-
tionship between connected porosity and diffusion tortuosity
guez et al., 2008; Adam, 2009; Bernal et al., 2011). Carbonation
(Provis et al., submitted for publication). The diffusion tortuosity
of alkali-activated binders is an open and active area of research,
describes the effect on mass transport of the constriction due to
and much remains to be explained in this area, particularly with
the presence of the pore network. An increase in this parameter
regard to the relationship between carbonation and steel degrada-
by almost a factor of two between early age (<7 days) and 45 days
tion, which may or may not be similar to the corresponding rela-
of age indicates that the rate of diffusion through a well-cured bin-
tionship in Portland cement-based concretes. This suggests that
der would be halved when compared to a poorly-cured binder. This
there are additional effects which compound with, or mitigate,
further highlights the importance of adequate curing in achieving
the direct influence of porosity on permeability (particularly ionic
high performance and durability in geopolymer concretes. A criti-
permeability, which also relates to gel chemistry-specific effects
cism of alkali-activation technology is that when mixes are
and interactions) and durability.
poorly-designed, thermal curing is often required for adequate
strength development. This is not the case for a well-designed
mix with sufficiently well-controlled activation conditions, as has 6. Technical challenges
been demonstrated for systems based on fly ash, on slag, and on
mixtures of these two precursors. However, it is clear from the In developing the technology of novel cements and concretes, it
tomography results that, regardless of the rate of early strength is often important to apply expertise obtained from the general
98 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

field of mineral processing, and in particular from the knowledge the case with OPC. To a large extent this addresses the supply chain
of fluid-particle interactions which has resulted from intensive glo- challenges and difficulties in price competition which are inherent
bal research efforts in the optimisation of hydrometallurgical pro- when sourcing materials from OPC-related suppliers. However, the
cesses. For example, the phase chemistry of cement and other process of establishing such dry binder processing facilities is cap-
mineral components (particularly fly ash and slag) is critical in ital intensive and requires significant market drivers for geopoly-
controlling reactivity. So, as was discussed in detail in the preced- mer concretes (synthesised via the existing process) to justify
ing sections of this paper, the accurate understanding of the differ- investment.
ent glass phases present in geopolymer precursor materials is
required to determine the optimal processing method and compo- 7. Reduction in carbon emissions
sition of binder components to be applied to manipulate concrete
properties. Classified fly ash and ground slag available through CO2 emissions from cement production are incurred through
the cement supply chain are not always able to be used straightfor- the consumption of fossil fuels, the use of electricity and the chem-
wardly with existing commercially available alkaline activators to ical decomposition of limestone during clinkerisation, which can
prepare optimal geopolymer binders, so special preparation of take place at around 1400 °C. The decarbonation of limestone to
alkaline activators may also be warranted. In addition, fly ash give the calcium required to form silicates and aluminates in clin-
and slag (being waste materials) are subject to far more variability, ker releases roughly 0.53 tonnes CO2 per tonne of clinker (Damtoft
especially when they are obtained from a range of sources, than et al., 2008). In 2005, cement production (total cementitious sales
OPC as a quality controlled product. Therefore, online monitoring including OPC and OPC blends) had an average emission intensity
and optimisation of activator types and binder components is nec- of 0.89 with a range of 0.65–0.92 tonnes CO2 per tonne of cement
essary in the production of consistent, high-quality geopolymer binder (International Energy Agency, 2007). Therefore, the decar-
concretes. bonation of limestone contributes about 60% of the carbon emis-
As was mentioned briefly above, it is important to control the sions of OPC, with the remaining 40% attributed to energy
rheology of fresh concrete to enable it to be placed and finished, consumption, most of which is related to clinker kiln operations;
and the ability to do this without adversely affecting the final prop- the WWF-Lafarge Conservation Partnership (2008) estimated that
erties of the hardened concrete depends primarily on the manipu- the production of clinker is responsible for over 90% of total ce-
lation of colloidal interactions by the use of chemical admixtures ment production emissions.
such as superplasticisers (Puertas et al., 2003; Criado et al., In view of the fact that the requirement for decarbonation of
2009). The existing range of commercially available superplasticis- limestone presents a lower limit on CO2 emissions in clinker pro-
ers have been developed specifically to suit the complex series of duction, and that there exist technical issues associated with the
chemical reactions which take place in the OPC system, and are addition of supplementary cementitious material (SCM, including
usually not effective in the geopolymer system. Moreover, most fly ash and ground granulated blast furnace slag), which restrict
of the various admixtures used to control slump, air dispersion, the viability of direct Portland cement supplementation by SCM
water retention, and other properties of the OPC system are less above certain limits, the possibility to reduce CO2 emissions using
effective in the geopolymer system. Consequently, there is a need OPC chemistry is limited. The WWF-Lafarge Conservation Partner-
to develop a whole set of new admixtures for the geopolymer sys- ship (2008) expects that the emissions intensity of cement, includ-
tem, which presents a significant challenge for an emerging indus- ing SCM, could be reduced to 0.70 tonnes CO2 per tonne of cement
try with a lack of scale, but which is still required to compete with by 2030 – which still amounts to around 2 billion tonnes of CO2 per
the well-established OPC industry. annum worldwide, even if cement production does not increase
Alkali activation also provides the potential for the utilisation of from its current level.
non-blast furnace slags; alternative materials such as ferronickel, Fig. 7 shows the CO2 emissions of various binder designs as a
steel and phosphorus slags have also been alkali-activated to form function of OPC content. There have been a limited number of life
usable binders and concretes, some on a laboratory scale and some cycle analyses of alkali activation technology. One reasonably
in larger-scale applications in China and the former Soviet Union extensive research program carried out in Germany (Weil et al.,
(Shi et al., 2006). The leaching of toxic metal components from 2009) has provided information regarding the selection of precur-
some of these slags during activation may prove to be a cause for sors and mix designs for a range of geopolymer-based materials.
some concern, but – as is the case for fly ashes – the selection of However, geographic specificity plays a significant role in a full life
appropriate waste materials for use as geopolymer precursors is
both important and possible.
At present, geopolymer concrete is most commonly produced in
a ‘two-part’ mix format, by blending coal fly ash, slag and alkali
activators together in a concrete batching plant. This requires a Business as usual
high level of skill from the operators of the batching plant, which
is rarely available in the premix concrete industry. Ash and slag Best Practice 2011
are waste materials of variable composition, so that the concrete
mix designs and the added alkalis need to be varied to compensate
for these variations in the ash and slag. Such a technology, where
the quality control is mainly at the batching plant level, is not scal-
Stretch/Aspirational
able on an industry-wide level and has limited appeal in the mar-
ket, despite the growing market pull for ‘green’ construction
materials.
E-CreteTM since 2008
In contrast with this situation, Zeobond has developed a process
whereby the various solid materials and proprietary activators are
processed together to produce a dry cement binder that behaves in
a similar way to OPC (Van Deventer et al., 2010b). The quality con-
trol is hence centralised in the cement binder plant, so that the dry
powder can be distributed to various concrete batching plants, as is Fig. 7. CO2 emissions of various cement binders as a function of OPC content.
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 99

cycle analysis, so there is the need for further studies considering drance to the acceptance of geopolymer technology. However, by
different locations in addition to a wider range of mix designs working with all stakeholders (Fig. 8), these barriers can be over-
spanning the broader spectrum of alkali activated materials. The come, provided that the intent of regulatory standards is met.
most carbon-intensive and also the most expensive ingredient in VicRoads, the state roads authority in Victoria, Australia, re-
geopolymer cement is the alkali activator, which should be mini- cently recognised geopolymer concrete as being equivalent to
mised in mix design. Two more recent studies (McLellan et al., OPC for non-structural applications in a 2010 update to their de-
2011; Habert et al., 2011) have also provided additional insight sign specification Section 703 (VicRoads, 2010); this specification
in this area, and further highlighted the importance of the activator requires acceptable binder content, strength development, water
in determining the environmental impact of geopolymer concretes. content and quality, aggregate properties, consistency, mixing
A commercial life cycle analysis was conducted by the NetBal- and placement methodologies, curing, finishing, and material sup-
ance Foundation, Australia, on Zeobond’s E-Crete™ geopolymer plier competency guarantees. Zeobond is working with VicRoads to
cement, as reported in the ‘Factor Five’ report released recently also recognise geopolymer concrete for structural applications in
by the Club of Rome (Von Weizsäcker et al., 2009). This life cycle Section 620. Zeobond’s E-Crete is already used in VicRoads struc-
analysis compared the geopolymer binder to the standard OPC tural projects, and by local councils and housing developers in
blends available in Australia in 2007 on the basis of both binder- sub-divisional works and slabs, applications which represent
to-binder comparison and concrete-to-concrete comparison. The approximately 70% of all concrete usage. These large scale applica-
binder-to-binder comparison showed an 80% reduction in CO2 tions are pivotal to the process of gradually convincing standards
emissions, whereas the comparison on a concrete-to-concrete ba- authorities to accept geopolymer concrete.
sis showed slightly greater than 60% savings, as the energy cost
of aggregate production and transport was identical for the two
materials. However, this study was again specific to a single loca- 9. Testing for durability
tion and a specific product, and it will be necessary to conduct fur-
ther analyses of new products as they reach development and The question of whether geopolymer concretes are durable re-
marketing stages internationally. Fig. 7 shows a comparison of mains the major obstacle to recognition in standards for structural
the CO2 emissions of four different E-Crete™ products against the concrete, and hence to their commercial adoption. A material such
‘Business as usual’, ‘Best practice 2011’, and a ‘Stretch/Aspirational’ as geopolymer, which has been subjected to detailed investigation
target for OPC blends. It is noted that in some parts of the world only recently, cannot possibly have the availability of decades of
(particularly Europe), some of the blends shown here in the in-service testing and durability data to prove its long-term stabil-
‘Stretch/Aspirational’ category are in relatively common use for ity. Most standard methods of testing cement and concrete dura-
specific applications, particularly CEM III-type OPC/slag blends, bility involve exposing small samples to very extreme conditions
but this is neither achievable on a routine scale worldwide at pres- – in particular highly concentrated acid or salt solutions, with or
ent, nor across the full range of applications in which Portland ce- without the application of electrical field gradients – for short peri-
ment concretes are used in large volumes. ods of time. The data obtained from these tests are then used to
predict how the material will perform under normal environmen-
tal conditions over a period of decades or more. In some of these
8. Standards framework predictive models, engineering concepts including mass transport
through porous media, reaction kinetics and particle packing are
Fig. 8 shows that the strategic development of standards is piv- used, although usually in a simplified and semi-empirical form to
otal to the commercialisation of geopolymer cement. The regula- enable utilisation of the derived equations by non-specialists in
tory framework governing concrete to be utilised in various the field. However, the key shortcoming of this approach to ‘prov-
applications relies on a typical cascade of standards, with applica- ing’ durability is that it can only provide indications of the ex-
tion standards referring to concrete standards, and concrete pected performance, rather than definitive proof. Therefore, there
standards referring to standards covering cement and other raw has been a very slow process of adoption of new materials, as it
materials. Hence, when considering the regulatory framework for is considered necessary to wait up to 20–30 years for ‘real-world’
a concrete binder system such as geopolymer cement or concrete, verification. Adoption of fly ash and slag in OPC concretes is the
most of the attention to regulatory aspects should be focused on prime example of this, where the use of these SCMs was resisted
the cement standards, although some aspects of concrete stan- for decades. It is the authors’ experience that asset owners and
dards also need to be considered. In general, all of the world’s con- their insurance companies are willing to use geopolymer concrete
crete application, concrete and cement standards are based on two in low risk applications based on accelerated durability testing.
‘super’ standards, i.e., European Union EN 197 and United States Higher-risk applications such as high rise buildings, which consti-
ASTM C150/C595/C1157. For instance, Chinese cement and con- tute a smaller fraction of the total concrete market, will follow only
crete standards are based largely on European Union Standards, when the market is comfortable with the real-world track record of
while Australian standards are based mainly on the American the material in low risk applications. Therefore, a staged approach
standards. towards the development of standards and commercial adoption
These standards have been developed over many years, and in needs to be followed, as outlined in Fig. 8.
collaboration with input from OPC manufacturing companies, with Zeobond also works closely with various roads authorities in
the chemistry and behaviour of OPC-based concretes intrinsically in monitoring the performance of in situ geopolymer concrete struc-
mind. However, prescriptive standards containing constraints such tures and also with researchers in developing appropriate test
as ‘minimum cement content’ are increasingly being viewed as methods for durability (Fig. 8). The work conducted by RILEM Tech-
excessively prohibitive, even for OPC-based systems. Products such nical Committee 224 on Alkali Activated Materials (AAM) provides
as geopolymer concrete may not simply be an evolution of existing essential advice to standards authorities about the structure of per-
OPC technology, but instead may require an entirely different chem- formance based standards for AAM and the associated testing
ical paradigm to understand their behaviour, and may perform en- methods for durability (http://www.rilem.net/tcDetails.php?tc=
tirely acceptably, but without conforming exactly to the established 224-AAM).
regulatory standards – particularly with regard to rheology and As outlined by Provis et al. (2011b), there have been a number
chemical composition (Hooton, 2008). This is a significant hin- of studies of chloride diffusion in alkali-activated slag concretes,
100 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

Fig. 8. Flow diagram for commercialisation of geopolymer cement and concrete.

with the performance of these materials in accelerated chloride production of GBFS. The further development and industrial adop-
penetration tests generally observed to be at least comparable to tion of the new CSIRO air-cooled slag granulation method (Jahan-
that of OPC. Zeobond’s E-Crete has been shown repeatedly to have shahi and Xie, 2010) is important in regions with a shortage of
significantly lower chloride diffusion and acceptable freeze–thaw clean water, which is the standard chilling medium used at pres-
performance compared with OPC concrete. As outlined in Section ent. The main risk facing GBFS supply for geopolymer use is pref-
5.4, the excellent performance of geopolymer concrete is related erential utilisation of GBFS in OPC blends rather than in
to the highly refined pore network forming a dense low-calcium geopolymers.
C–A–S–H phase (Provis et al., 2011b). It is envisaged that advanced There is a risk that a substantial reduction in world-wide carbon
techniques such as synchrotron-based nanotomography will ulti- emissions will result in reduced coal fired energy production, and
mately be used to compute transport properties of the geopolymer hence reduced availability of coal ash. While there is a global
gel generated from various mix designs (such as water/binder ra- determination to reduce energy dependency on thermal coal, it is
tio), and hence changes in durability performance. more likely that coal combustion will continue to increase in the
coming decades, but with some level of carbon capture through
sequestration or mineralisation. In any event, only a very small
10. Supply chain risks fraction of coal ash is currently used in concrete. A more substan-
tial risk is the well-intended control of coal ash as a hazardous
The presence of a competitive and efficient supply chain is material by, for example the US Environment Protection Agency,
essential for the successful scale-up of geopolymer technology. which leads to uncertainty in investment in the supply chain of
For instance, granulated blast furnace slag (GBFS) is produced dur- all SCMs.
ing the production of pig-iron. Although there is a risk that produc- In economically developed and structurally segmented markets
tion methods for pig-iron will change and reduce the availability of for cementitious materials, distribution networks for slag and fly
GBFS, this is not currently viewed as a substantial risk. As demand ash are established, making it uncompetitive or practically impos-
for GBFS increases, blast furnace operators will increasingly invest sible to invest in new channels to market. In some cases, even volu-
in rapid cooling slag handling equipment in order to increase the minous by-product streams such as slag and fly ash are largely
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 101

Fig. 9. E-Crete™ in Port Melbourne, Australia: (a) 25 MPa footpath along Westgate Freeway extension. (b) 55 MPa precast panels across Salmon Street bridge.

utilised or earmarked via long-term option contracts. Due to the formation, reaction kinetics and transport phenomena underlying
rapid increase in demand and production in markets such as China geopolymerisation of aluminosilicates, including by the applica-
and India, new opportunities for both OPC and geopolymer supply tion of advanced experimental and computational techniques.
channels are foreseeable. However, in established markets such as However, despite much research, the role of calcium in geopolymer
the UK, where fly ash production does not exceed its consumption gel processes and phase formation remains poorly understood.
in cement, the existing market for OPC utilises all of the available Analysis of the nanostructure of geopolymer gels has enabled the
fly ash. Despite the increasing number of research papers on geo- tailored selection of geopolymer precursors and the design of alkali
polymers, there has been little development of the supply chain activator composition; such studies have established the relation-
channels necessary for scale production, which will continue to ship between geopolymer gel microstructure and durability. Geo-
limit the wider adoption of geopolymer technology until this bot- polymer concrete has performed very well in accelerated
tleneck is resolved. durability tests, which is supported by its tortuous pore structure
and the specific details of its gel chemistry, although detailed ana-
lytical work in this area is still ongoing. In the absence of a long in-
11. The path of commercialisation
service track record, chemical research is essential to validate
durability testing methodology and improve geopolymer cement
Fig. 9a shows a small section of E-Crete™ paving completed as
technology. Although the principles of particle technology have
part of the Victorian Government upgrade of the Westgate Freeway
been applied to the mix design and rheology of fresh geopolymer
in Port Melbourne. While small in volume, this project required
concrete, this area is in its infancy and could benefit from the
that the concrete meet all of the technical specifications of the local
knowledge base of mineral processing engineers. For example, no
road authority, VicRoads, and the ultimate asset owner, the Port
suitable superplasticiser exists which can reliably enhance the rhe-
Melbourne City Council, and was approved for use by a construc-
ology and reduce the required mix water content of fresh geopoly-
tion consortium which included numerous national and multi-
mer concrete.
national construction companies and engineering firms. This small
Demand pull applied by a carbon conscious market continues to
project demonstrates the entire process of commercialisation of
be the key driver for the short term adoption of geopolymer ce-
geopolymer concrete (Fig. 8).
ment. While the scale-up from the laboratory to the real-world is
Fig. 9b shows the installation of 55 MPa E-Crete precast panels
technically challenging, the core challenge is the scale-up of indus-
for VicRoads. This concrete was required to meet the structural con-
try participation and acceptance of geopolymer cement. High pro-
crete code (VicRoads Section 610), which is far more stringent than
file demonstration projects in Australia have highlighted the vast
the requirements faced by the non-structural grade concrete in
regulatory, asset management, liability and industry stakeholder
Fig. 9a. Here the level of concrete specification and scrutiny was
engagement process required to commercialise geopolymer
consistent with bridge design, with the aim of using the highest
cement.
grade concrete specified by VicRoads as a ‘stretch target’ for trial
The fact that all premixed and precast concrete standards are
purposes.
based on an assumption of the use of OPC remains a major obstacle
As opposed to focusing on the technical detail of these projects,
to the commercial adoption of geopolymer cement. Even when as-
attention is drawn to the fact that these case studies demonstrate
set owners and specifiers such as government, architects and de-
something that is rarely seen in geopolymer technology, i.e., the
sign engineers accept the results of durability testing of
vast regulatory, asset management, liability and industry stake-
geopolymer concrete, the main barrier to entry of geopolymers
holder engagement process that has been undertaken by Zeobond
into an established market is access to a suitable supply of source
to commercialise its E-Crete™ (Fig. 8). The shift from the labora-
materials including fly ash, granulated blast furnace slag and alka-
tory to the real-world is not a process of scale-up of technology
line activators. Significant development of supply chains to accom-
alone. While such technical challenges are sufficient to be insur-
modate geopolymer cement, predominantly in developing
mountable to many, it is the ability to manage the scale-up of
markets, is essential in overcoming this barrier. This crucial aspect
industry participation and acceptance of geopolymer concrete that
of commercialisation is seldom appreciated by the research com-
provides the core challenge to the future world of a geopolymer ce-
munity, governments or the ultimate users of concrete. It is impor-
ment/concrete industry.
tant for commercial geopolymer concrete producers to work
closely with research partners to develop testing methods for
12. Conclusions accelerated durability, especially as longer term in-service testing
data become available. Substantial progress has been made in Aus-
Significant progress has been made in developing an under- tralia, where the local road authority has recognised geopolymer
standing of the colloid and interface science, gel chemistry, phase concrete for non-structural applications.
102 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

Acknowledgements Freedonia Group, 2009. World Cement to 2012. <http://www.freedoniagroup.com/


World-Cement.html>.
Gallucci, E., Scrivener, K., Groso, A., Stampanoni, M., Margaritondo, G., 2007. 3D
This work has been funded by Zeobond Pty. Ltd. and the Austra- experimental investigation of the microstructure of cement pastes using
lian Research Council (ARC) through an ARC Linkage Project grant synchrotron X-ray microtomography (lCT). Cement and Concrete Research 37
(3), 360–368.
awarded jointly to JLP and PD. Authors JSJvD and PD hold financial
García-Lodeiro, I., Palomo, A., Fernández-Jiménez, A., Macphee, D.E., 2011.
interest in Zeobond Pty. Ltd., a producer of geopolymer cements Compatibility studies between N–A–S–H and C–A–S–H gels. Study in the
and concretes. ternary diagram Na2O–CaO–Al2O3–SiO2–H2O. Cement and Concrete Research
41 (9), 923–931.
Gartner, E.M., Macphee, D.E., 2011. A physico-chemical basis for novel cementitious
References binders. Cement and Concrete Research 41 (7), 736–749.
Goltermann, P., Johansen, V., Palbol, L., 1997. Packing of aggregates: an alternative tool
Adam, A.A., 2009. Strength and Durability Properties of Alkali Activated Slag and Fly to determine the optimal aggregate mix. ACI Materials Journal 94 (5), 435–443.
Ash-based Geopolymer Concrete. Ph. D. Thesis, RMIT University, Melbourne, Gomes, K.C., Lima, G.S.T., Torres, S.M., De Barros, S., Vasconcelos, I.F., Barbosa, N.P.,
Australia. 2010. Iron distribution in geopolymer with ferromagnetic rich precursor.
Amirjanov, A., Sobolev, K., 2008. Optimization of a computer simulation model for Materials Science Forum 643, 131–138.
packing of concrete aggregates. Particulate Science and Technology 26 (4), 380– Habert, G., d’Espinose de Lacaillerie, J.B., Roussel, N., 2011. An environmental
395. evaluation of geopolymer based concrete production: reviewing current
Bernal, S.A., Mejía de Gutierrez, R., Pedraza, A.L., Provis, J.L., Rodríguez, E.D., research trends. Journal of Cleaner Production 19 (11), 1229–1238.
Delvasto, S., 2011. Effect of binder content on the performance of alkali- Hajimohammadi, A., Provis, J.L., Van Deventer, J.S.J., 2010. The effect of alumina
activated slag concretes. Cement and Concrete Research 41 (1), 1–8. release rate on the mechanism of geopolymer gel formation. Chemistry of
Bouzoubaâ, N., Zhang, M.H., Bilodeau, A., Malhotra, V.M., 1997. The effect of Materials 22 (18), 5199–5208.
grinding on the physical properties of fly ashes and a Portland cement clinker. Hajimohammadi, A., Provis, J.L., Van Deventer, J.S.J., 2011a. The effect of silica
Cement and Concrete Research 27 (12), 1861–1874. availability on the mechanism of geopolymerisation. Cement and Concrete
Bouzoubaâ, N., Zhang, M.H., Malhotra, V.M., Golden, D.M., 1999. Blended fly ash Research 41 (3), 210–216.
cements – a review. ACI Materials Journal 96 (6), 641–650. Hajimohammadi, A., Provis, J.L., Van Deventer, J.S.J., 2011b. Time-resolved and
Brough, A.R., Atkinson, A., 2000. Automated identification of the aggregate-paste spatially-resolved infrared spectroscopic observation of seeded nucleation
interfacial transition zone in mortars of silica sand with Portland or alkali- controlling geopolymer gel formation. Journal of Colloid and Interface Science
activated slag cement paste. Cement and Concrete Research 30 (6), 849–854. 357 (2), 384–392.
Brouwers, H., 2004. The work of powers and Brownyard revisited: Part 1. Cement Häkkinen, T., 1992a. The permeability of high strength blast furnace slag concrete.
and Concrete Research 34 (9), 1697–1716. Nordic Concrete Research 11 (1), 55–66.
Buchwald, A., Wiercx, J., 2010. ASCEM cement technology – alkali-activated cement Häkkinen, T., 1992b. The microstructure of high strength blast furnace slag
based on synthetic slag made from fly ash. In: Shi, C., Shen, X. (Eds.), Advances concrete. Nordic Concrete Research 11 (1), 67–82.
in Chemically Activated Materials CAM’2010. RILEM Publications, Jinan, China, Hobbs, D.W., 2001. Concrete deterioration: causes, diagnosis, and minimising risk.
pp. 15–21. International Materials Reviews 46 (3), 117–144.
Buchwald, A., Hilbig, H., Kapps, C., 2007. Alkali-activated metakaolin-slag blends – Hooton, R.D., 2008. Bridging the gap between research and standards. Cement and
performance and structure in dependence on their composition. Journal of Concrete Research 38 (2), 247–258.
Materials Science 42 (9), 3024–3032. Hos, J.P., McCormick, P.G., Byrne, L.T., 2002. Investigation of a synthetic
Byfors, K., Klingstedt, G., Lehtonen, H.P., Pyy, H., Romben, L., 1989. Durability of aluminosilicate inorganic polymer. Journal of Materials Science 37 (11),
concrete made with alkali-activated slag. In: Malhotra, V.M. (Ed.), 3rd 2311–2316.
International Conference on Fly Ash, Silica Fume, Slag and Natural Pozzolans Hunger, M., Brouwers, H.J.H., 2009. Flow analysis of water–powder mixtures:
in Concrete. ACI SP114, pp. 1429–1444. application to specific surface area and shape factor. Cement and Concrete
Chen, W., Brouwers, H., 2007. The hydration of slag, part 1: reaction models for Composites 31 (1), 39–59.
alkali-activated slag. Journal of Materials Science 42 (2), 428–443. International Energy Agency, 2007. Tracking Industrial Energy Efficiency and CO2
Collins, F., Sanjayan, J.G., 2001. Microcracking and strength development of alkali Emissions – Executive Summary. <http://www.iea.org/Textbase/npsum/
activated slag concrete. Cement and Concrete Composites 23 (4–5), 345–352. tracking2007SUM.pdf>.
Criado, M., Palomo, A., Fernández-Jiménez, A., Banfill, P.F.G., 2009. Alkali activated Ismail, I., Provis, J.L., Van Deventer, J.S.J., Hamdan, S., 2011. The effect of water
fly ash: effect of admixtures on paste rheology. Rheologica Acta 48 (4), 447–455. content on compressive strength of geopolymer mortars. In: CD Proceedings of
Damtoft, J.S., Lukasik, J., Herfort, D., Sorrentino, D., Gartner, E., 2008. Sustainable 7th AES-ATEMA 2011 International Conference on Advances and Trends in
development and climate change initiatives. Cement and Concrete Research 38 Engineering Materials and their Applications, Milan, Italy, 4–8 July 2011.
(2), 115–127. Jahanshahi, S., Xie, D., 2010. A new integrated dry slag granulation and heat
Davidovits, J., 2009. Geopolymer Chemistry & Applications. second ed. Institut recovery process. In: McCaffrey, R., Edwards, P. (Eds.), CD Proceedings of 6th
Géopolymère, Saint-Quentin, France, 590pp. Global Slag Conference. PRo Publications Int. Ltd., Sydney, Australia.
Deja, J., 2002. Carbonation aspects of alkali activated slag mortars and concretes. Ji, T., Lin, T.W., Lin, X.J., 2006. A concrete mix proportion design algorithm based on
Silicates Industries 67 (1), 37–42. artificial neural networks. Cement and Concrete Research 36 (7), 1399–1408.
Diamond, S., 2000. Mercury porosimetry. An inappropriate method for the Jones, M.R., Zheng, L., Newlands, M.D., 2002. Comparison of particle packing models
measurement of pore size distributions in cement-based materials. Cement for proportioning concrete constituents for minimum voids ratio. Materials and
and Concrete Research 30 (10), 1517–1525. Structures 35 (5), 301–309.
Duxson, P., Provis, J.L., 2008. Designing precursors for geopolymer cements. Journal Jones, M.R., Zheng, L., Newlands, M.D., 2003. Estimation of the filler content
of the American Ceramic Society 91 (12), 3864–3869. required to minimise voids ratio in concrete. Magazine of Concrete Research 55
Duxson, P., Lukey, G.C., Separovic, F., Van Deventer, J.S.J., 2005a. The effect of alkali (2), 193–202.
cations on aluminum incorporation in geopolymeric gels. Industrial & Juenger, M.C.G., Winnefeld, F., Provis, J.L., Ideker, J., in press. Advances in alternative
Engineering Chemistry Research 44 (4), 832–839. cementitious binders. Cement and Concrete Research. doi: 10.1016/
Duxson, P., Provis, J.L., Lukey, G.C., Mallicoat, S.W., Kriven, W.M., Van Deventer, J.S.J., j.cemconres.2010.11.012.
2005b. Understanding the relationship between geopolymer composition, Justnes, H., Ardoullie, B., Hendrix, E., Sellevold, E.J., Van Gemert, D., 1998. The
microstructure and mechanical properties. Colloids and Surfaces A 269 (1–3), chemical shrinkage of pozzolanic reaction products. In: Malhotra, V.M. (Ed.),
47–58. 6th CANMET Conference on Fly Ash, Silica Fume, Slag, and Natural Pozzolans in
Duxson, P., Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2007a. The role of inorganic Concrete. Bangkok, Thailand, pp. 191–205.
polymer technology in the development of ‘Green concrete’. Cement and Kaufmann, J., Loser, R., Leemann, A., 2009. Analysis of cement-bonded materials by
Concrete Research 37 (12), 1590–1597. multi-cycle mercury intrusion and nitrogen sorption. Journal of Colloid and
Duxson, P., Fernández-Jiménez, A., Provis, J.L., Lukey, G.C., Palomo, A., Van Deventer, Interface Science 336 (2), 730–737.
J.S.J., 2007b. Geopolymer technology: the current state of the art. Journal of Keyte, L.M., 2008. What’s Wrong With Tarong? The Importance of Coal Fly Ash Glass
Materials Science 42 (9), 2917–2933. Chemistry in Inorganic Polymer Synthesis, Ph.D. Thesis, University of
Duxson, P., Lukey, G.C., Van Deventer, J.S.J., 2007c. Physical evolution of Na- Melbourne, Australia.
geopolymer derived from metakaolin up to 1000 °C. Journal of Materials Science Koloušek, D., Brus, J., Urbanova, M., Andertova, J., Hulinsky, V., Vorel, J., 2007.
42 (9), 3044–3054. Preparation, structure and hydrothermal stability of alternative (sodium
Fernández-Jiménez, A., Palomo, A., 2005. Mid-infrared spectroscopic studies of silicate-free) geopolymers. Journal of Materials Science 42 (22), 9267–9275.
alkali-activated fly ash structure. Microporous and Mesoporous Materials 86 Komnitsas, K., Zaharaki, D., 2007. Geopolymerisation: a review and prospects for
(1–3), 207–214. the minerals industry. Minerals Engineering 20 (14), 1261–1277.
Fernández-Jiménez, A., Palomo, A., Sobrados, I., Sanz, J., 2006. The role played by the Kumar, S., Kumar, R., Alex, T.C., Bandopadhyay, A., Mehrotra, S.P., 2007. Influence of
reactive alumina content in the alkaline activation of fly ashes. Microporous and reactivity of fly ash on geopolymerisation. Advances in Applied Ceramics 106
Mesoporous Materials 91 (1–3), 111–119. (3), 120–127.
Fernández-Jiménez, A., Monzó, M., Vicent, M., Barba, A., Palomo, A., 2008. Alkaline Laskar, A.I., Talukdar, S., 2008. Rheological behavior of high performance concrete
activation of metakaolin–fly ash mixtures: obtain of zeoceramics and with mineral admixtures and their blending. Construction and Building
zeocements. Microporous and Mesoporous Materials 108 (1–3), 41–49. Materials 22 (12), 2345–2354.
J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104 103

Lee, W.K.W., Van Deventer, J.S.J., 2004. The interface between natural siliceous Rajaokarivony-Andriambololona, Z., Thomassin, J.H., Baillif, P., Touray, J.C., 1990.
aggregates and geopolymers. Cement and Concrete Research 34 (2), 195–206. Experimental hydration of two synthetic glassy blast furnace slags in water and
Lee, W.K.W., Van Deventer, J.S.J., 2007. Chemical interactions between siliceous alkaline solutions (NaOH and KOH 0.1 N) at 40 °C: structure, composition and
aggregates and low-Ca alkali-activated cements. Cement and Concrete Research origin of the hydrated layer. Journal of Materials Science 25 (5), 2399–2410.
37 (6), 844–855. Rees, C.A., Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2007a. Attenuated total
Liu, Z., Beaudoin, J.J., 2000. The permeability of cement systems to chloride ingress reflectance Fourier transform infrared analysis of fly ash geopolymer gel aging.
and related test methods. Cement, Concrete and Aggregates 22 (1), 16–23. Langmuir 23 (15), 8170–8179.
Lloyd, R.R., Provis, J.L., Van Deventer, J.S.J., 2009a. Microscopy and microanalysis of Rees, C.A., Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2007b. In situ ATR-FTIR study of
inorganic polymer cements. 1: remnant fly ash particles. Journal of Materials the early stages of fly ash geopolymer gel formation. Langmuir 23 (17), 9076–9082.
Science 44 (2), 608–619. Rees, C.A., Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2008. The mechanism of
Lloyd, R.R., Provis, J.L., Van Deventer, J.S.J., 2009b. Microscopy and microanalysis of geopolymer gel formation investigated through seeded nucleation. Colloids and
inorganic polymer cements. 2: the gel binder. Journal of Materials Science 44 Surfaces A 318 (1–3), 97–105.
(2), 620–631. Richardson, K., Steffen, W., Schellnhuber, H.J., Alcamo, J., Barker, T., Kammen, D.M.,
Lloyd, R.R., Provis, J.L., Smeaton, K.J., Van Deventer, J.S.J., 2009c. Spatial distribution Leemans, R., Liverman, D., Munasinghe, M., Osman-Elasha, B., Stern, N., Waever,
of pores in fly ash-based inorganic polymer gels visualised by Wood’s metal O., 2009. Synthesis Report from Climate Change: Global Risks, Challenges &
intrusion. Microporous and Mesoporous Materials 126 (1–2), 32–39. Decisions. University of Copenhagen, 10–12 March 2009, Copenhagen,
Lloyd, R.R., Provis, J.L., Van Deventer, J.S.J., 2010. Pore solution composition and Denmark <http://climatecongress.ku.dk/pdf/synthesisreport/>.
alkali diffusion in inorganic polymer cement. Cement and Concrete Research 40 Rockström, J., Steffen, W., Noone, K., Persson, Å., Chapin III, F.S., Lambin, E.F., Lenton,
(9), 1386–1392. T.M., Scheffer, M., Folke, C., Schellnhuber, H.J., Nykvist, B., De Wit, C.A., Hughes,
Lothenbach, B., Le Saout, G., Gallucci, E., Scrivener, K., 2008. Influence of limestone T., Van der Leeuw, S., Rodhe, H., Sörlin, S., Snyder, P.K., Costanza, R., Svedin, U.,
on the hydration of Portland cements. Cement and Concrete Research 38 (6), Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J., Walker, B.,
848–860. Liverman, D., Richardson, K., Crutzen, P., Foley, J.A., 2009. A safe operating space
McLellan, B.C., Williams, R.P., Lay, J., van Riessen, A., Corder, G.D., 2011. Costs and for humanity. Nature 461, 472–475.
carbon emissions for geopolymer pastes in comparison to ordinary Portland Rodríguez, E., Bernal, S., Mejía de Gutierrez, R., Puertas, F., 2008. Alternative
cement. Journal of Cleaner Production 19 (9–10), 1080–1090. concrete based on alkali-activated slag. Materiales de Construcción 58 (291),
Mehta, P.K., Monteiro, P.J.M., 2006. Concrete: Microstructure, Properties and 53–67.
Materials, third ed. McGraw-Hill, New York, 659pp. Roy, D.M., Jiang, W., Silsbee, M.R., 2000. Chloride diffusion in ordinary, blended, and
Nakashima, Y., Kamiya, S., 2007. Mathematica programs for the analysis of three- alkali-activated cement pastes and its relation to other properties. Cement and
dimensional pore connectivity and anisotropic tortuosity of porous rocks using Concrete Research 30 (12), 1879–1884.
X-ray computed tomography image data. Journal of Nuclear Science and Sagoe-Crentsil, K., Brown, T., Yan, S., 2010. Medium to long term engineering
Technology 44 (9), 1233–1247. properties and performance of high-strength geopolymers for structural
Nemati, K.M., 2000. Preserving microstructure of concrete under load using the applications. Advances in Science and Technology 69, 135–142.
Wood’s metal technique. International Journal of Rock Mechanics and Mining Scrivener, K.L., Kirkpatrick, R.J., 2008. Innovation in use and research on
Sciences 37 (1–2), 133–142. cementitious material. Cement and Concrete Research 38 (2), 128–136.
O’Connor, S.J., MacKenzie, K.J.D., 2010. Synthesis, characterisation and thermal Shi, C., 1996. Strength, pore structure and permeability of alkali-activated slag
behaviour of lithium aluminosilicate inorganic polymers. Journal of Materials mortars. Cement and Concrete Research 26 (12), 1789–1799.
Science 45 (14), 3707–3713. Shi, C., 2004. Effect of mixing proportions of concrete on its electrical conductivity
Oh, J.E., Monteiro, P.J.M., Jun, S.S., Choi, S., Clark, S.M., 2010. The evolution of and the rapid chloride permeability test (ASTM C1202 or ASSHTO T277) results.
strength and crystalline phases for alkali-activated ground blast furnace slag Cement and Concrete Research 34 (3), 537–545.
and fly ash-based geopolymers. Cement and Concrete Research 40 (2), 189–196. Shi, C., Xie, P., 1998. Interface between cement paste and quartz sand in alkali-
Olivia, M., Nikraz, H., Sarker, P., 2008. Improvements in the strength and water activated slag mortars. Cement and Concrete Research 28 (6), 887–896.
penetrability of low calcium fly ash based geopolymer concrete. In: Uomoto, T., Shi, C., Krivenko, P.V., Roy, D.M., 2006. Alkali-Activated Cements and Concretes.
Nga, T.V. (Eds.), 3rd ACF International Conference – ACF/VCA 2008. Ho Chi Minh Taylor & Francis, Abingdon, UK, 376pp.
City, Vietnam, pp. 384–391. Shi, C., Fernández-Jiménez, A., Palomo, A., 2011. New cements for the 21st century:
Ollivier, J.P., Maso, J.C., Bourdette, B., 1995. Interfacial transition zone in concrete. the pursuit of an alternative to Portland cement. Cement and Concrete Research
Advanced Cement Based Materials 2 (1), 30–38. 41 (7), 750–763.
Pacheco-Torgal, F., Castro-Gomes, J., Jalali, S., 2008a. Alkali-activated binders: a Shimoda, K., Tobu, Y., Kanehashi, K., Nemoto, T., Saito, K., 2008. Total understanding
review. Part 1. Historical background, terminology, reaction mechanisms and of the local structures of an amorphous slag: perspective from multi-nuclear
hydration products. Construction and Building Materials 22 (7), 1305–1314. (29Si, 27Al, 17O, 25Mg, and 43Ca) solid-state NMR. Journal of Non-Crystalline
Pacheco-Torgal, F., Castro-Gomes, J., Jalali, S., 2008b. Alkali-activated binders: a Solids 354 (10–11), 1036–1043.
review. Part 2. About materials and binders manufacture. Construction and Sindhunata, Provis, J.L., Lukey, G.C., Xu, H., Van Deventer, J.S.J., 2008. Structural
Building Materials 22 (7), 1315–1322. evolution of fly ash-based geopolymers in alkaline environments. Industrial &
Palomo, A., Banfill, P.F.G., Fernández-Jiménez, A., Swift, D.S., 2005. Properties of Engineering Chemistry Research 47 (9), 2991–2999.
alkali-activated fly ashes determined from rheological measurements. Sofi, M., Van Deventer, J.S.J., Mendis, P.A., Lukey, G.C., 2007. Engineering properties
Advances in Cement Research 17 (4), 143–151. of inorganic polymer concretes (IPCs). Cement and Concrete Research 37 (2),
Plimer, I., 2009. Heaven and Earth. Global Warming: The Missing Science. Connor 251–257.
Court Publishing, Ballan, Australia, 503pp. Taylor, M., Tam, C., Gielen, D., 2006. Energy Efficiency and CO2 Emissions from the
Provis, J.L., Van Deventer, J.S.J., 2007. Geopolymerisation kinetics. 2. Reaction kinetic Global Cement Industry. International Energy Agency.
modeling. Chemical Engineering Science 62 (9), 2318–2329. Van Deventer, J.S.J., Provis, J.L., Duxson, P., Brice, D.G., 2010a. Chemical research and
Provis, J.L., Van Deventer, J.S.J. (Eds.), 2009. Geopolymers: Structures, Processing, climate change as drivers in the commercial adoption of alkali activated
Properties and Industrial Applications. CRC Press, Cambridge UK, 54pp. materials. Waste and Biomass Valorization 1 (1), 145–155.
Provis, J.L., Lukey, G.C., Van Deventer, J.S.J., 2005. Do geopolymers actually contain Van Deventer, J.S.J., Feng, D., Duxson, P., 2010b. Dry Mix Cement Composition,
nanocrystalline zeolites? – a reexamination of existing results. Chemistry of Methods and Systems Involving Same. US Patent 7691,198 B2.
Materials 17 (12), 3075–3085. VicRoads, 2010. Standard Specification Section 703, General Concrete Paving.
Provis, J.L., Muntingh, Y., Lloyd, R.R., Xu, H., Keyte, L.M., Lorenzen, L., Krivenko, P.V., <http://www.vicroads.vic.gov.au/Home/Moreinfoandservices/TendersAndSuppliers/
Van Deventer, J.S.J., 2007. Will geopolymers stand the test of time? Ceramic StandardDocuments.htm>.
Engineering and Science Proceedings 28 (9), 235–248. Von Weizsäcker, E., Hargroves, K., Smith, M.H., Desha, C., Stasinopoulos, P., 2009.
Provis, J.L., Rose, V., Bernal, S.A., Van Deventer, J.S.J., 2009. High resolution nanoprobe Factor Five: Transforming the Global Economy Through 80% Improvements in
X-ray fluorescence characterization of heterogeneous calcium and heavy metal Resource Productivity. Earthscan, London, UK.
distributions in alkali activated fly ash. Langmuir 25 (19), 11897–11904. Weil, M., Dombrowski, K., Buchwald, A., 2009. Life-cycle analysis of geopolymers.
Provis, J.L., Duxson, P., Van Deventer, J.S.J., 2010. The role of particle technology in In: Provis, J.L., van Deventer, J.S.J. (Eds.), Geopolymers: Structures, Processing,
developing sustainable construction materials. Advanced Powder Technology Properties and Industrial Applications. Woodhead Publishing, CRC Press,
21 (1), 2–7. Cambridge, UK, pp. 194–212.
Provis, J.L., Rose, V., Winarski, R.P., Van Deventer, J.S.J., 2011a. Hard X-ray White, C.E., Provis, J.L., Proffen, T., Van Deventer, J.S.J., 2010. The effects of
nanotomography of amorphous aluminosilicate cements. Scripta Materialia temperature on the local structure of metakaolin-based geopolymer binder: a
65 (4), 316–319. neutron pair distribution function investigation. Journal of the American
Provis, J.L., Ismail, I., Myers, R.J., Rose, V., Van Deventer, J.S.J., 2011b. Characterising Ceramic Society 93 (10), 3486–3492.
the structure and permeability of alkali-activated binders. RILEM International White, C.E., Perander, L.M., Provis, J.L., Van Deventer, J.S.J., 2011. The use of XANES to
Conference on Advances in Construction Materials Through Science and clarify issues related to bonding environments in metakaolin: a discussion of
Engineering, Hong Kong, 4–8 September 2011. the paper S. Sperinck et al., ‘Dehydroxylation of kaolinite to metakaolin – a
Provis, J.L., Myers, R.J., White, C.E., Rose, V., Van Deventer, J.S.J., submitted for molecular dynamics study,’ J Mater Chem 21: 2118–2125; 2011. Journal of
publication. X-ray microtomography shows pore structure and tortuosity in Materials Chemistry 21 (19), 7007–7010.
alkali-activated binders. Cement and Concrete Research. White, C.E., Provis, J.L., Llobet, A., Proffen, T., Van Deventer, J.S.J., in pressa. Evolution
Puertas, F., Palomo, A., Fernández-Jiménez, A., Izquierdo, J.D., Granizo, M.L., 2003. of local structure in geopolymer gels: an in-situ neutron pair distribution
Effect of superplasticisers on the behaviour and properties of alkaline cements. function analysis. Journal of the American Ceramic Society. doi: 10.1111/j.1551-
Advances in Cement Research 15 (1), 23–28. 2916.2011.04515.x.
104 J.S.J. van Deventer et al. / Minerals Engineering 29 (2012) 89–104

White, C.E., Provis, J.L., Proffen, T., Van Deventer, J.S.J., in pressb. Molecular <http://assets.panda.org/downloads/cement_blueprint_climate_fullenglrep_lr.
mechanisms responsible for the structural changes occurring during pdf>.
geopolymerization: multiscale simulation. AIChE Journal. doi: 10.1002/ Xu, X., Provis, J.L., Van Deventer, J.S.J., Krivenko, P.V., 2008. Characterization of aged
aic.12743. slag concretes. ACI Materials Journal 105 (2), 131–139.
Willis, K.L., Abell, A.B., Lange, D.A., 1998. Image-based characterization of cement Yip, C.K., Lukey, G.C., Van Deventer, J.S.J., 2005. The coexistence of geopolymeric gel
pore structure using Wood’s metal intrusion. Cement and Concrete Research 28 and calcium silicate hydrate at the early stage of alkaline activation. Cement
(12), 1695–1705. and Concrete Research 35 (9), 1688–1697.
Wong, H.H.C., Kwan, A.K.H., 2008. Packing density of cementitious materials: Yip, C.K., Lukey, G.C., Provis, J.L., Van Deventer, J.S.J., 2008. Effect of calcium silicate
measurement and modelling. Magazine of Concrete Research 60 (3), 165–175. sources on geopolymerisation. Cement and Concrete Research 38 (4), 554–564.
Wong, H.S., Buenfeld, N.R., Head, M.K., 2006. Estimating transport properties of Zhang, J.X., Sun, H.H., Wan, J.H., Yi, Z.L., 2009. Study on microstructure and
mortars using image analysis on backscattered electron images. Cement and mechanical property of interfacial transition zone between limestone aggregate
Concrete Research 36 (8), 1556–1566. and sialite paste. Construction and Building Materials 23 (11), 3393–3397.
WWF-Lafarge Conservation Partnership, 2008. A Blueprint for a Climate Friendly
Cement Industry: How to Turn Around the Trend of Cement Related Emissions.

You might also like