Vous êtes sur la page 1sur 12

J. Am. Ceram. Soc.

, 97 [4] 997–1008 (2014)


DOI: 10.1111/jace.12831
© 2014 The American Ceramic Society

Journal
Durability of Alkali-Activated Materials: Progress and Perspectives
Susan A. Bernal and John L. Provis†
Department of Materials Science and Engineering, The University of Sheffield, Sheffield S1 3JD, South Yorkshire, UK

In the last decade, there has been rapid growth in interest in materials known as alkali-activated binders, or “geopoly-
alternative binders, as part of the toolkit of cement technolo- mers,” which have been seen to potentially offer an attractive
gies needed to mitigate the carbon footprint associated with the combination of good mechanical strength in the hardened
construction industry. Alkali-activated materials (AAMs), state, good durability in service, a reduced environmental
including geopolymer binders and other related systems, have footprint, and acceptable early-age rheology,4,5 if adequate
been identified as a key component of this move to lower CO2 formulation and curing conditions are adopted.
cements and concretes. These are clinker-free cements which These materials are produced by the reaction between an
can exhibit comparable performance to conventional portland/ alkali source, often (but not necessarily) supplied as a liquid,
blended cements, when they are adequately formulated and and a solid aluminosilicate precursor. The most commonly
cured. However, AAMs have a somewhat limited record of used precursors are also the materials which show good poz-
durability in service, and this is one of the main limitations fac- zolanic activity when blended with portland cement: blast-
ing their commercial adoption at present. To provide the best furnace slag, fly ash from pulverized coal combustion, and
possible answers to the question of long-term durability within calcined kaolinite-group clays. These precursors (which will
an experimentally accessible timeframe, standardized acceler- be simply referred to as slag, fly ash, and metakaolin, respec-
ated degradation testing methods have been widely adopted, in tively, through the remainder of this article) all contain reac-
an attempt to simulate natural processes. It has been identified tive Al and Si, and all dissolve under alkaline conditions to
that the interactions between material and environment, which release these components at a rate which is suitable for the
take place on microstructural and nanostructural levels, have a formation of a cement-like binder, and the components then
very significant influence on the outcomes of the durability rearrange through solution to form a strong dense binding
tests. Here, we present an overview of the results obtained gel.6
when AAMs are exposed to aggressive testing conditions such In a technical sense, the formation of a dense and strong
as elevated concentrations of CO2, sulfates or chlorides. The binding gel is what is needed in the production of mortars
key outcome of this article is a broader synthesis of the avail- and concretes, and there are many types of alternative bind-
able data regarding the interactions between these new materi- ers which undergo setting and hardening reaction processes
als and their surrounding environment, which is then available at comparable rates to portland cement, and provide a
to be used in the design, development, and implementation of broadly similar rate of strength development.3 However,
environmentally sustainable, high-performance cements and there are some limitations currently facing the adoption of
concretes for the 21st century. these alternative materials, such as the need for development
of a fuller characterization of their performance under differ-
ent physicomechanical and chemical conditions. This charac-
I. Introduction
terization is essential if practitioners are to ensure that

T HE development of lower CO2 materials which can be


used instead of pure portland cement in high-volume
applications is an area of intense research interest at present,
alternative materials can be fit-for-purpose in designated
applications, as strength development is not the only impor-
tant criterion when selecting a material for use in a specific
both with regard to blended cements containing a small application. Also, many non-portland binders are either
amount of portland cement (or clinker) and a high or major- much more expensive to produce, or require the use of
ity content of supplementary materials, and in the area of resources or components which are not widespread world-
non-portland cements.1–3 Among these various types of bind- wide. This is less of an issue for alkali-activated binder sys-
ers, one that has attracted particular attention is the class of tems, as the worldwide availability of reactive aluminosilicate
materials (slag, fly ash, clays, natural pozzolans, and some
others) is in general high, the chemistry of this class of mate-
D. J. Green—contributing editor rials is sufficiently flexible that an activator can be selected
and tailored to match the precursors available, and the acti-
vators themselves are generally derived from existing high-
volume industrial processes (e.g., the chlor-alkali process for
Manuscript No. 34049. Received October 30, 2013; approved December 29, 2013. NaOH, and the Solvay process or trona mining for Na2CO3,

Author to whom correspondence should be addressed. e-mail: j.provis@sheffield.
ac.uk which can either be used directly as an activator in some

Feature
998 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4

&
Sidebar: “A rose by any other name…”

One of the issues which is particularly troublesome in the research community investigating alkali-activated binders is the prolif-
eration of different names which are applied to essentially the same material. These names include, but are not limited to:
“alkali-bonded ceramic,” “alkali ash material,” “F-concrete,” “geocement,” “geopolymer,” “hydroceramic,” “inorganic polymer,”
“inorganic polymer glass,” “mineral polymer,” “SKJ binder,” “soil cement,” “soil silicate,” “zeocement,” “zeoceramic,” and a
variety of other names including various commercial trademarks. This scattering of terminology leads to confusion in the scien-
tific and technical communities, and means that it is unnecessarily difficult to identify relevant information in the scientific litera-
ture via simple searches. While there is no simple solution to this situation, this article will follow the recommendations of the
RILEM (Reunion Internationale des Laboratoires et Experts des Materiaux, systemes de construction et ouvrages) Technical
Committee 224-AAM (Alkali Activated Materials), in naming each system explicitly according to the components it contains.
This has the disadvantage of leading to some long names, particularly for materials such as sodium metasilicate-activated slag-fly
ash blends (which will be discussed in detail in the latter part of the article), but does offer clarity among the sometimes-confus-
ing selection of terminology which is on offer in this field. The term “geopolymer” will be used specifically to refer to the almost
fully cross-linked low-calcium alkali aluminosilicate gel which is present in many alkali-activated binders, as distinct from the
chain-type calcium silicate hydrate (C–S–H) type gel, which is formed in binders with higher calcium content.
So, to answer the question to which the title of this Sidebar alludes—a rose by any other name does smell just as sweet, but
in this instance, the smell is less essential than the ease of identification which comes from the use of more direct terminology.

systems, or thermally combined with silica to form sodium along with a low concentration of Q3 sites in chain-bridging
silicate).4 roles in some systems depending on the concentrations of Ca
Therefore, the family of alkali-activated binder systems and Al.8–11 Binders containing much less calcium than this
can be seen to generally meet the two main criteria required (usually <10% on a mass basis in the precursor), derived
for a high-volume portland cement alternative: (1) these from the alkali-activation of fly ashes and/or metakaolin, are
materials can be produced in many parts of the world, from dominated by a Q4-structured alkali aluminosilicate gel,12–16
a wide range of feedstocks, and (2) the technical properties often described as a geopolymer gel.14,17 Binders containing
of the materials, in both fresh and hardened states, are suffi- an intermediate level of calcium (alkali-activated slag/fly ash,
ciently comparable to those of portland cement that they can slag/metakaolin blends, and some sole precursor materials4)
be considered a like-for-like substitute in suitable applica- are much more complex in chemistry, with descriptions of
tions, enabling material placement and finishing using the the gel ranging from a single-phase C–N–A–S–H gel to a
existing skills of trained concrete workers.5 fully phase-separated assemblage of C–A–S–H and geopoly-
It is also noted that it is extremely unlikely that alkali-acti- mer-type gels, along with secondary phases.4,11,18–23 Fig. 1
vated binders will be a suitable replacement for portland shows an example of the gel compositions which have been
cement across the full range of applications in which this observed by scanning electron microscopy–energy diffractive
material is used worldwide—and the globally achievable pro- X-ray spectroscopy (SEM–EDX) analysis18 of a binder syn-
duction volumes of alkali-activated materials (AAMs) would thesized by sodium metasilicate activation of slag, fly ash
be far too low to meet such demand in any case. Instead, a (containing less than 0.5 wt% CaO by XRF analysis), and a
toolkit-type approach is needed, where each type of cement- 1:1 mixture of the two. Several distinct regions of gel compo-
ing material (including portland cement, alkali-activated sition are observable, both in the fly ash-only binder and in
binders, and many others) will be used in the applications to the blended system. The SEM–EDX spot analyses were
which it is best suited. conducted at points which were visually identifiable as corre-
With this in mind, it therefore becomes essential to under- sponding to the binder gel rather than unreacted precursor
stand and predict the gel chemistry and durability properties
of alkali-activated binders, because it is only with such
understanding that the optimal fit-for-purpose mixes and
applications can be identified, and the material put into use
appropriately. This article will, therefore, present a brief dis-
cussion of the key aspects which control the gel chemistry of
alkali-activated binders on both nanostructural and micro-
structural levels, leading to a discussion of the influence of
some of these parameters on the durability of the materials
when exposed to three commonly studied types of chemically
aggressive environment: carbonation, sulfate, and chloride. It
is noted that predominantly physical modes of degradation
are not covered in this review; the reader is referred instead
to the relevant chapter of the State-of-the-Art Report of
RILEM TC 224-AAM7 for detailed discussion in this area.

II. Gel Chemistry in Alkali-Activated Binders


(1) One, Two, or Many?
In general, alkali-activated binders can be divided into two
main classes, according to the calcium content of the binder
systems. Binders derived from the alkali-activation of slags
with a Ca/(Al + Si) ratio similar to one are generally domi- Fig. 1. Pseudo-ternary plot of alkali-activated binder gel
nated by a C–S–H type gel (normally described as C–A–S–H compositions, measured by SEM–EDX after 28 d of curing. Binders
to reflect its high Al concentration compared to the low-Al are synthesized by sodium metasilicate activation of fly ash (FA),
gels formed through portland cement hydration processes), slag (BFS), and a 1:1 mixture of these two precursors. Data from
with silicate chains comprising Q1 and Q2 environments, Ref. [18].
April 2014 Durability of Alkali‐Activated Materials 999

particles, although it is almost impossible to exclude the (metakaolin in Ref. [30] or fly ash in Ref. [34]) where the
possibility that some buried unreacted particles were included material changes from being predominantly C–A–S–H in
in the probe volume.24 character at higher slag content, with a pore volume that
The exact details of the chemistry in this compositional decreasing with increasing binder maturity, to being predomi-
region are still the subject of active research, enabled by nantly geopolymer-like at low Ca content. In such systems,
rapid developments in analytical capabilities, and are to a the pore structure becomes much more refined as the mate-
significant extent also controlled by the nature (and particu- rial becomes more mature, but the overall porosity does not
larly the pH) of the activator selected, where a very high pH change much as a function of time.
leads to precipitation of the calcium as Ca(OH)2, either as
discrete crystallites or as amorphous/nanoscale particu-
lates,25,26 rather than having it incorporated into gels as is (2) Heterogeneity–Gel Chemistry as a Function of Time
the case at a more moderate pH. and Space
This possibility for coexistence of multiple gels leads to The spatial aspect associated with gel chemistry in alkali-acti-
scope for tailoring of some potentially interesting chemical vated binder systems is clearly essential in determining the
and physical properties in binder systems produced from final binder structure and properties. These materials are
blended precursors. For example, Lecomte et al.27 found that highly spatially heterogeneous,35 particularly when precursors
the fine slag particles present in alkali-activated blends of such as fly ash, containing a distribution of particle size,
slag and metakaolin were the main source of the chemical chemistry, and mineralogy,24,26,36–39 are themselves contribut-
species involved in initial hardening and development of ing to the heterogeneity on a length scale of micrometers or
early strength, and identified both highly cross-linked more. The various rates of reactions taking place at particle
(Q4(nAl)) and undercoordinated chain-type (Q1 and Q2) Si surfaces, in the free solution volume, and in different loca-
sites in alkali-activated binders with a 2:1 ratio of slag to me- tions within the hardened binder, can leave signature features
takaolin. Zhang et al.28 identified an optimum in strength at in the hardened microstructure in terms of chemical composi-
a 1:1 ratio of slag to metakaolin, while Bernal et al.29 found tion and morphology. Thus, the ability to identify these fea-
that the optimal slag/metakaolin ratio for strength develop- tures in the microstructure of the binders using both in situ40
ment increased with increasing activator modulus. The resid- and ex situ13,24,41 microscopic analysis, as well as tomogra-
ual strength of blended alkali-activated slag-metakaolin phy on varying length scales,34,42 can provide insight into the
samples after heating to 1000°C is also better than that of reaction processes, which led to their formation.
solely slag-based materials,30,31 as the thermally induced The use of fly ash as a component of cementitious binders
shrinkage and cracking processes are less destructive in the is not new; fly ashes have been blended with portland cement
blended systems due to the influence of the additional Si and in concretes since the 1930s,43 including extensive use in cur-
Al supplied by the metakaolin. rent practice (either as blended cements or blended concretes)
The physicochemical influences of gel coexistence can be in some parts of the world. High-volume fly ash concretes,
observed in analysis of the transport properties of the mate- where portland cement is blended with more than 50% fly
rial as a whole, where the two types of gel will each preferen- ash, are finding more widespread application as their specific
tially interact with different dissolved species, which will be workability and curing requirements are beginning to be bet-
discussed in detail below in the context of analysis of chlo- ter understood,44 and “hybrid” alkali-portland-fly ash bind-
ride ingress into these materials. There are also important ers have also been described recently.45 However, the specific
differences in space-filling properties between the C–A–S–H case of alkaline activation in the absence of clinker necessi-
gel, which has high bound-water content, and the geopoly- tates a much fuller understanding of fly ash glass chemistry,
mer-type gel which does not chemically bind water. These as this precursor becomes the only source of reactive compo-
differences become evident when the materials are analyzed nents for binder formation.
by thermogravimetry30,32 (Fig. 2), or when porosity is deter- In understanding the formation and structure of an alkali-
mined either by laboratory testing33 or by X-ray microto- activated binder, it is thus essential to understand the reac-
mography.34 It is becoming increasingly clear that there is a tion mechanism by which the solid precursor is converted to
tipping point, close to a 1:1 ratio of slag to Ca-free precursor the final binder structure. In the case of fly ash, the reaction
mechanisms involved in its conversion into a monolithic
alkali-activated gel are complex and not yet well understood.
This issue is complicated further by the fact that fly ashes
vary dramatically between sources, and as a function of time
even when sourced from a single electricity generation facil-
ity. A brief discussion of fly ash chemistry, and its relation-
ship to the mechanical and durability performance of the
alkali-activated products, is thus important here.
A microstructural-scale conceptual model for NaOH acti-
vation of fly ash was presented by Fern andez-Jimenez et al.46
whereby the attack of the hydroxide on the glassy shell of a
partially hollow spherical fly ash particle leads to the forma-
tion of reaction products both outside and inside the ash
particle, leading to a final microstructure which contains
embedded fly ash particles with varying degrees of reaction.
This range of reactivities is evident in Fig. 3, where the dif-
ferent types of particles highlighted by colored circles have
visibly reacted to different extents, and contributed in differ-
ent ways to the development of the gel network, which has
filled the interparticle region. It is also notable from Fig. 3
Fig. 2. Differential thermograms for alkali silicate-activated blended
that there are no distinct “outer product” and “inner prod-
binders based on blends of granulated blast-furnace slag (GBFS) and
metakaolin, with the low-temperature region expanded (inset) to uct” regions in this material, and this is common to the
highlight the distinct signals which are attributable to the two types majority of well-cured silicate-activated fly ash binders.
of gel present in the materials as a function of slag/metakaolin ratio. Since the work of Fern andez-Jimenez et al.46 was pub-
Successive data sets are offset by 0.07%/°C for display. Data from lished, other workers24,41,47,48 have expanded the understand-
Ref. [30]. ing of various aspects of the reaction model, in particular
1000 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4

in enhancing the reaction of alkali-activated slag binders,


unlike the situation for fly ash-based materials, due to prob-
lems associated with drying. An increasing activator dose in
an alkali-activated slag binder generally leads to a more
refined pore network,57 similar to the situation for fly ash-
based binders,58 although the optimal silicate modulus for
generation of a dense structure in these binders is often
higher than the values which have been identified for binders
based on metakaolin or fly ash.29,59–62

III. Pore Solution Chemistry, and its Importance at Early


and Later Ages
One of the key developments in the chemistry of portland
cement-based concretes in the past decades has been the
enormous advancement which has been made in the design
and use of organic admixtures to influence concrete proper-
ties in both the fresh and hardened state. This enables the
Fig. 3. SEM image of a polished section of a fly ash geopolymer, use of the advanced construction techniques, which are criti-
with different fly ash particle types circled in different colors. Red: cal to the utilization of modern concretes in advanced infra-
iron-rich particles; yellow: reactive glass particles surrounded by a
reaction rim; blue: hollow particles whose shell was not broken
structure applications such as ultra-high-rise buildings.
during the reaction process; pink: relatively unreactive glass particles However, because the additives which are now commercially
with no visible reaction rim; green: glass particles containing visible available have been specifically tailored to be compatible
crystalline inclusions. Image courtesy E. D. Rodrıguez. with the chemistry of portland cement, their use in alkali-
activated binders tends to be much less effective.4,63–65 The
discussing the importance of gel nucleation, which can take surface charging behavior of the precursor particles in a fresh
place either on, or away from, the particle surfaces, depend- alkali-activated binder mix is very different from the surface
ing on the silica content of the activator23,40 and the provi- charge on portland cement particles, which means that dis-
sion of additional nucleation seeding sites in the form of persant-type molecules would require quite different chemis-
added nanoparticles.49,50 The question of when and where try to act effectively, and the aqueous solution environment
nucleation takes place, and particularly whether the nucle- in an alkali-activated binder is much higher in both pH and
ated gel particles grow outwards from the particle surfaces overall dissolved ion concentrations. The high degree of vari-
and thus block the possibility of further reaction, is critical ability in physical properties and chemical composition of
in determining the final extent of reaction which is achieved. the precursors used in producing alkali-activated binders rep-
Turning to the higher-calcium end of the spectrum of resents another important challenge in the development of
AAMs, the structural development of binders based on alkali- organic admixtures for these systems, as the interaction of
activated slag is also a significantly heterogeneous reaction these compounds and the particle surfaces can be strongly
process, even though the precursor itself is much more homo- influenced by the nature and source of a specific precursor.
geneous on both interparticle and intraparticle levels. In this This means that a good deal of further development will be
case, the heterogeneity is most important on the nanoscale, required to place alkali-activated concretes on a more equal
where the chemical constraints on C–A–S–H gel compositions footing with portland cement concretes, particularly for the
introduce the formation of a wide range of secondary prod- production of premixed concretes. In fact, it will probably be
ucts, often enriched in Al (layered double hydroxides and/or necessary to specifically design a new suite of organic addi-
zeolites),8,9,51,52 and also containing the majority of the Mg tives which are stable in the highly alkaline pore solution
which is supplied by the slag, as this is largely excluded from environment of AAMs, and which are compatible with the
the tobermorite-like C–A–S–H structure. The reaction process surface chemistry (charge and functionality) of the alumino-
is governed by a four-step mechanism: dissolution of the silicate precursors used in AAMs, to achieve the desired rhe-
glassy precursor particles, nucleation and growth of the initial ology modification in these binder systems.64–68 Initial steps
solid phases, interactions and mechanical binding at the in this direction have been made in recent years, but there is
boundaries of the phases formed (which leads to setting of the still much development work required to make this into a
paste), and ongoing reaction via dynamic chemical equilibria commercial-scale reality.
and diffusion of reactive species through the reaction products Considering now the material in its hardened state, the
formed at advanced times of curing. aqueous environment prevailing within the pore solution is
In these slag-rich systems, there are often microstructurally also critical in terms of durability. In any concrete which is
distinct inner product and outer product regions correspond- reinforced with mild steel, it is essential that the pore solu-
ing to the areas initially occupied by slag particles or by the tion pH is sufficiently high (above approximately pH 10–12)
solution, respectively. This is particularly evident in silicate- to passivate the embedded steel reinforcing bars,69 and that
activated binders,53 while it has been noted that in some this condition is maintained throughout the service life of the
hydroxide-activated systems, the inner product and outer material. Given the initially very high pH of all alkali
product regions of the gel tend to have similar composi- hydroxide-activated and silicate-activated binder systems,
tions.54 In a silicate-activated system, the outer product forms which can often exceed 13.5 and may be as high as 14.5 in
rapidly, driven by the high concentrations of silica in the ini- some hydroxide-activated binders, it may appear that this
tially fluid-filled regions, which thus become supersaturated should be straightforward to achieve. However, the fact that
with respect to C–S–H type phases at a relatively low concen- the phase assemblage of a hardened alkali-activated binder
tration of Ca.55,56 This means that the outer product regions does not include a pH-buffering solid hydroxide phase, such
can show a lower Ca/Si ratio than the inner product regions. as portlandite among portland cement hydrates, means that
The pore structure of alkali-silicate-activated slag binders the ability to maintain high alkalinity is contingent on
has been reported to be either very impermeable or rather preventing either the ingress of acidic components (particu-
open, depending on the mix design, curing conditions and larly H+ or CO2) from the external environment, or the
specimen maturity, where obviously an immature specimen leaching of mobile alkalis from the pore solution.70 It is also
has a more open pore network. This is controlled by the important to note that Huet et al.69 evaluating the transition
reaction mechanism. Thermal curing is not always effective of steel rebars from a passive to an active state in polluted
April 2014 Durability of Alkali‐Activated Materials 1001

simulated pore solutions, have identified buffering effects due This then leaves a variety of open questions regarding the
to carbonate species in the system, and, therefore, it is not correct application of standardized testing procedures in the
ideal to consider the pH as the sole criterion to study this analysis of AAMs, particularly where the testing standards
phenomenon. This means that the pore structure (porosity specify the curing and preconditioning regimes to be applied to
and tortuosity) of an alkali-activated binder, coupled with the material. It has been noted that in some instances the tests
key chemical considerations, is likely to be the key factor can be fairly straightforwardly modified (e.g., by changing the
which determines its durability in a reinforced concrete appli- details of curing environments or the temperature of a precon-
cation.5 These two factors are therefore the focus of the sub- ditioning step) to give results which are then more logical and
sequent sections. valuable in predicting in-service performance,4,33 and an
international technical committee (RILEM TC 247-DTA,
http://www.rilem.org/gene/main.php?base=8750&gp_id=290)
IV. Durability of Alkali-Activated Binders—Testing and is now examining this issue in detail across a range of different
Key Parameters durability testing methods. It is hoped that this will lead
(1) Curing, and Why it is Important to be Kind to Your toward an improved understanding of how exactly these mate-
Concrete rials should be tested in a laboratory context to enable accu-
Related to the issue of pore structure refinement and/or pore rate and reliable prediction of in-service performance.
filling, as mentioned above in the context of binder reaction From this background, three specific areas of durability
mechanisms, is the need to select appropriate curing condi- will be discussed in the following sections: resistance to car-
tions for an alkali-activated binder. It is often stated that bonation, resistance to external sulfate attack, and resistance
low-calcium geopolymer-type binders require elevated- to chloride ingress.
temperature curing for satisfactory strength or microstruc-
tural development, and this has been considered by some to
place restrictions on the industrial-scale applicability of this (2) Carbonation
technology. However, it has been demonstrated, both in the Durability of hydraulic cement products is broadly defined
laboratory and in the field,5 that ambient-temperature or as the ability of the material to resist weathering action,
near-ambient curing of these materials can provide very good chemical attack, abrasion or any process of deterioration.73
results in terms of strength development (as high as 60 MPa Specifically, weathering and chemical attack are often com-
after 28 d at 23°C, and up to 100 MPa at 180 d, for 50-mm bined in a degradation process referred to as carbonation. In
mortar cubes71). However, this does require control of three conventional portland cement, carbonation is understood as
key parameters: the chemical reaction taking place between the carbonic acid
formed through the dissociation, in the pore solution, of the
1. the mix designs must be formulated to enable a low
CO2 that diffuses into the material from the external atmo-
water/binder ratio, otherwise setting and strength
sphere, and the hydration products of the binder, producing
development will be delayed;
carbonate deposits within the binder itself.74 This mechanism
2. the activator concentration and composition must be
of deterioration leads to a significant decrease in the pH of
tailored to match the chemistry of the precursor, to
the system, sometimes accompanied by reduction in mechani-
prevent either slow reaction due to insufficient alkali
cal strength, along with an increase in the total porosity.
dosage, or efflorescence if there are too many mobile
These combined physicochemical changes in the cement
alkalis in the final product; and
mean that the metallic component in reinforced concretes
3. curing must be conducted under sealed conditions to
becomes very prone to suffer corrosion, which can lead to
prevent drying or microcracking.
the eventual failure of the material.75
These binders are also somewhat prone to suffer micro- Carbonation is controlled by both diffusional and chemi-
cracking under the aggressive drying conditions used in many cal mechanisms, and is influenced by the relative humidity in
standardized durability testing methods, meaning that their the material, the concentration of CO2 in the surrounding
potentially good in-service performance is not always well atmosphere, and the tortuosity of the pore network, along
represented by the results of standardized laboratory with the chemistry of the binding phases and the pore solu-
tests.4,18,32 A somewhat obvious, yet too-often neglected, tion environment within a specific binder. In AAMs, the
comment which must be made here is that in-service perfor- mechanism of carbonation is strongly dependent on the type
mance is far more important than accelerated laboratory test of precursor used in producing these binders (slag or fly
results in proving the actual value of a material. However, it ash)11 and the nature and concentration of the activator
is equally true that poor laboratory test outcomes will in all used,76,77 as these parameters control the type of the reaction
likelihood prevent any application of the material in service, products formed. It has also been identified that the MgO
and the most important parameter in ensuring that a mate- content of the slag precursor has a significant influence on
rial performs well in the laboratory is to develop a well- both the mechanism and rate of carbonation of AAMs.78
cured, good-quality specimen for analysis. This is associated with the formation of layered double
This then requires careful specification and control of a hydroxides with a hydrotalcite type structure as a secondary
laboratory curing environment to enable the material to reaction product, when using slags with higher MgO con-
develop sufficient maturity prior to service or testing. This is tents. Layered double hydroxides have the capacity to absorb
critical in the context of materials testing for performance CO2, and the formation of these Al-rich secondary phases in
validation, where the standard curing conditions which are silicate-activated slag binders also reduces the degree of Al
used for portland cement-based materials (particularly lime- incorporation in the C–A–S–H type gel. These combined
water curing) are unlikely to give adequate binder maturity effects seem to enhance the resistance to carbonation of
in alkali-activated systems due to the likelihood of leaching alkali-activated slag binders.78
of the alkalis from the pore solution if they are cured from It has also been noted that design parameters of the con-
early age immersed in any aqueous environment which is not crete, such as the content of binder,79 can influence the rate
highly enriched in alkalis. Internal curing has also been of carbonation in AAMs. This means that adequate formulation
shown to be useful in mitigating microcracking in alkali-acti- and selection of precursors, aiming to reduce the permeability,
vated slags.72 It may also be that the usual 28-d curing per- can strongly reduce the likelihood of performance loss associ-
iod, which is often defined as giving “final” strength is not ated with this degradation mechanism.
sufficient for alkali-activated binders based on less-reactive Carbonation is generally a relatively slow process under
precursors,4 as has also been widely identified for high- environmental conditions due to the low CO2 concentrations
volume blended portland cements. in the atmosphere (~0.04%).76 So, to predict the performance
1002 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4

of specific binders against this mode of degradation within an rates of less than 1 mm/yr are identified,61,83 which is compa-
experimentally accessible timescale, accelerated carbonation rable to what can be expected for reasonably good-quality
testing methods have been adopted, using environmental portland cement-based materials exposed under the same
chambers with high CO2 concentrations, where the materials conditions. Evaluation of 7-yr old naturally carbonated sili-
are exposed to controlled conditions and the ingress of the cate-activated slag concretes84 shows that the water content
carbonation front into the material is monitored over a speci- in the concrete mix has little influence on the carbonation
fied period of time. Accelerated carbonation testing methodol- rate at advanced age. Some of the concretes evaluated (for-
ogies have mostly not been validated in detail by comparison mulated with 500 kg/m3 of slag) showed carbonation depths
to natural carbonation of comparable samples, and there is of 1.7 mm after 7 yr, which is low compared with normal-
still ongoing discussion regarding the most applicable testing strength portland cement concretes, where carbonation
protocols and conditions,80 with no testing methodology or depths between 2 and 10 mm are reported after 1 yr of natu-
conditions being universally preferred. ral exposure.85 A backscattered electron image and corre-
Consequently, the relationship between accelerated car- sponding elemental maps of a 7-yr old naturally carbonated
bonation rates and in-service carbonation rates in alkali-acti- alkali-activated slag concrete are shown in Fig. 5. The areas
vated concretes, which is essential in any attempt to use the enriched in Si, Al, and Na, with little content of Ca, are a
results of accelerated carbonation testing to calculate cover zeolite type phase, most likely garronite/heulandite, as identi-
depths required for a given service life, is largely unknown. fied via X-ray diffraction.84 Regions enriched in Ca (light
Most calculations conducted in this area use a simple diffu- gray in the BSE image) are unreacted slag particles, and the
sion-controlled model (t1/2 functional form) to extrapolate gray areas enriched in Ca, Si, and Al correspond to the bind-
accelerated testing data to service conditions, without consid- ing gel. Darker gray areas in the BSE image enriched in C
eration of any possible chemical effects associated with and O are assigned to carbonated regions. These results pro-
changing the CO2 concentration by a factor of 100 or more. vide evidence that in naturally carbonated slag activated con-
Under accelerated carbonation testing conditions, a higher cretes, the carbonated and noncarbonated regions are highly
susceptibility to carbonation has been identified in alkali-acti- intermixed, and neither a distinctive carbonation front, nor
vated slag compared to portland cements,77,79,81,82 which is substantial destruction of the material, is taking place
attributed to the absence of portlandite as a reaction product (Fig. 5). This is very different from the situation for acceler-
in these systems, the low Ca/Si ratio of the binding gel, and ated carbonated specimens (Fig. 4).
the high alkali content of the pore solution. All of these fac- The discrepancies between natural and accelerated carbon-
tors are stated to couple to increase the risk of carbonation ation results in AAMs are associated with the differences in
through the decalcification of C–A–S–H type gels to form exposure conditions other than CO2 concentration, as for
calcium carbonate. Consistent with this, Fig. 4 shows the instance controlled accelerated carbonation cannot simulate
change in the microstructure of an alkali-activated slag paste the dry/wet cycles to which the materials are subjected in ser-
which has been subjected to accelerated carbonation. The vice. The effects of the growing maturity of the binding
carbonated material is evidently highly porous and less dense phases during an extended service life are also important.
compared with the initial conditions before CO2 exposure. Under service conditions, the alkali activation reaction will
However, it is important to note that these observations continue to take place as long as the pore solution alkalinity
differ from the trends reported in specimens evaluated after remains high, and gel densification and pore network tortu-
several decades of service life, where natural carbonation osity will increase over time. This cannot be captured by lab-
oratory testing after 28 d of curing when a material is
designed for a life span of 50 yr, as the microstructures of
the two materials will be completely different. A recent study
has also demonstrated that the pore solution chemistry in
alkali-activated binders, which is dominated by alkali
hydroxides,55,70,86,87 plays a significant role in how the car-
bonation of these binders proceeds depending on the expo-
sure conditions. The chemistry of the pore solution of a
carbonated alkali-activated binder can be approximated by
the system Na2CO3–NaHCO3–H2O, where even slight incre-
ments in the CO2 concentration above natural conditions (as
little as 0.2%) favor the formation of nahcolite (NaHCO3)
and trona (Na2CO3NaHCO32H2O), instead of natron
(Na2CO310H2O), which is the stable product at room tem-
(a) perature and ambient CO2 concentration.88,89 In alkali-acti-
vated binders, the formation of natron has been observed
when the material has been exposed to atmospheric CO2,
while in specimens exposed to 1% CO2 or higher, formation
of nahcolite prevails.11,89,90 These differences are observed in
binders derived from both fly ash and slag precursors.
Natron has a greater molecular volume than nahcolite, and
it is likely that the relative extent of formation of each of
these phases has an impact on the permeability of the system,
as the formation of more space-filling phases such as natron
can contribute positively to pore blockage and reduction in
CO2 ingress. The modification in the carbonate/bicarbonate
phase equilibrium under accelerated carbonation conditions
also leads to a substantial reduction in the pH in accelerated
carbonated samples, which is unlikely to occur under natural
(b)
carbonation conditions.89
The formation of sodium carbonate heptahydrate
Fig. 4. Scanning electron micrograph (a) and backscattered electron (Na2CO37H2O) has been observed in 1-d cured fly ash-based
image (b) of a silicate-activated blast-furnace slag binder before (a) binders carbonated at 5% CO2. This indicates that the initial
and after (b) exposure to 1% CO2 for 1000 h. carbonation in these systems leads to the formation of an
April 2014 Durability of Alkali‐Activated Materials 1003

Fig. 5. Backscattered electron image and elemental maps (environmental SEM, uncoated polished sample) of a near-surface section of aged
silicate-activated slag concrete which has been subjected to natural carbonation exposure for 7 yr (images courtesy of R. San Nicolas).

internal CO2 concentration gradient (i.e., mass transport of alkali-activated slag binders in sodium sulfate solutions
CO2 is more rapid than the chemical reaction which causes its according to protocols similar to those used to investigate
precipitation as a solid phase), which means that the carbon- portland cement-based materials does not promote expansion
ation front becomes broadened, before a high CO2 concentra- or cracking of the material, in contrast to the situation for
tion is later able to diffuse through the pore network to portland cement-based samples, where significant expansion
permeate the binder. The N–A–S–H type geopolymer gel and cracking associated with the formation of secondary
forming in these fly ash-based binders is more porous than ettringite and/or gypsum is observed.91,92 Similar results have
the C–A–S–H phases forming in alkali-activated binders, as been reported for fly ash and bottom ash geopolymers
identified by X-ray microtomography,34 and discussed in Sec- exposed to Na2SO4-rich environments. Conversely, several
tion III. Less chemical disruption of the N–A–S–H gel struc- studies93–95 report significant loss of compressive strength
ture has been observed in fly ash-based binders when exposed and formation of gypsum and ettringite as sulfate attack
to high CO2 concentrations than is observed in C–A–S–H- reaction products in low-calcium fly ash-based alkali-acti-
based binders11; however, it is sometimes noted that acceler- vated binders after exposure to MgSO4 solutions. This is
ated carbonated Ca-free AAMs can lose their mechanical consistent with observations in high calcium fly ash geopoly-
strength, and some can even be crumbled by hand after expo- mers,96 where formation of gypsum, brucite, and magnesium
sure to elevated CO2 concentrations. This indicates that car- silicate hydrate phases has been observed after immersion in
bonation in Ca-free activated binders is not limited to the MgSO4.
pore solution, as modification in the role of the alkalis which Recent results92 have revealed that the key factor control-
were initially associated with the binding gels to charge bal- ling the degradation mechanism of AAMs is not the sulfate
ance Al seems to be occurring in the presence of high levels of itself, as tends to be the case in portland cement systems, but
CO2. Further work in this area is required to determine the rather it is the nature of the cation accompanying the sulfate
details of the mechanisms taking place. anions. Exposure to Na2SO4 seems to favor the structural
However, all of these comparisons and unexpected behav- evolution of the binding phases and densification of the sys-
iors under accelerated carbonation, and particularly the fac- tem,92 which is consistent with the known role of Na2SO4 as
tors of binder maturity and exposure conditions, combine to an activator in some higher-Ca alkali-activated and hybrid
indicate that accelerated laboratory testing is severely overes- binders.45,61 Conversely, the presence of magnesium as the
timating the real impact that carbonation will have on the cation accompanying sulfate anions leads to decalcification
pore solution chemistry, binder phase evolution and electro- of the Ca-rich gel phases in AAMs, promoting the decay of
chemistry of AAMs. the main binding phases and leading to formation of low-
strength M–S–H type phases.92
Figure 6 shows a BSE image and corresponding elemental
(3) Sulfate Exposure maps of an alkali-activated slag/fly ash concrete after
There is as yet a relatively limited understanding of the effect 6 months of exposure to MgSO4 solution. Angular particles
of sulfate exposure on the durability of AAMs. Immersion of correspond to unreacted slag and spherical particles are
1004 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4

Fig. 6. BSE image and elemental maps of alkali-activated 75 wt% slag/25 wt% fly ash concrete with 28 d curing after 6 months of MgSO4
exposure (Courtesy of I. Ismail).

partially reacted fly ash. Areas highly rich in Si are assigned although, there is a general consensus that these materials
to the fine and coarse aggregates used for producing the con- exhibit a lower permeability to chlorides compared to port-
cretes. In the exposed concrete, the main binding phase is land/blended cements, when cured adequately and evaluated
enriched in Mg and Si, with minor traces of Al, which dem- under the same experimental conditions. However, there is
onstrate that Mg2+ is effectively inducing the decalcification not yet an extensive database of chloride diffusion coefficients
of the binding phase, and promoting the formation of a available for alkali-activated concretes, and there is particu-
M–(A)–S–H type gel instead. A vein of gypsum running ver- larly a lack of information describing their variation with
tically through the specimen is also observed in the Ca and S time, which is needed for service life modeling. This is due in
elemental maps. This suggests that during the decalcification part to the fact that AAMs are more often studied by mate-
of the C–A–S–H products resulting from alkali-activation of rials scientists than by concrete technologists with a focus on
the slag, Ca2+ is released into the pore solution, which in the concrete durability, and partly because there are still signifi-
presence of a high concentration of SO42 favors the precipi- cant open questions regarding how these coefficients can be
tation of gypsum. reliably determined for non-portland cements. This is there-
fore a critical question requiring attention from the scientific
community.
(4) Chloride Ingress The various electrically accelerated testing methods
The rate of chloride ingress into a reinforced concrete is adopted to evaluate chloride movement in concretes vary
often used as a key parameter in durability or service life greatly in terms of what is actually measured102; for exam-
analysis, because under many types of environmental condi- ple, the Rapid Chloride Permeability test described in the
tions experienced in the developed world, the corrosion of ASTM C1202 standard in fact measures the conductivity
embedded steel reinforcing elements by chlorides is the main of the material, rather than actual permeability.104 Consid-
cause of concrete degradation. Chlorides are either applied ering the highly ionic pore solution environment present in
as de-icing salts in cold climate regions in the winter, or enter alkali-activated binders,70,87 as discussed in Section III, it is
the material due to marine spray/aerosols in coastal regions. expected that the mobility of all ions present (not just
The degree of structural change induced by chloride in the chloride, and particularly alkali cations and hydroxide
actual cement binder is generally low—particularly in the anions in alkali-activated binders) will be registered in the
case of alkali-activated binders, which do not show signifi- test results. This indicates that results obtained from
cant conversion of layered calcium sulfoaluminate hydrates cementitious materials with different pore solutions, such as
to chloroaluminates such as Friedel’s salt—and so the main portland-blended cements and AAMs, cannot be used as a
influence of binder characteristics in resisting chloride pene- criterion to determine which material will present lower
tration is that the binder must provide resistance to chloride chloride permeability in service, as the results cannot be
transport. directly comparable. The adoption of other testing methods
There are to date a limited number of studies focused to assess chloride permeability, such as NordTest NT Build
specifically on the chloride permeability of AAMs29,33,97–103; 492, and ponding tests including ASTM C1543, offers
April 2014 Durability of Alkali‐Activated Materials 1005

scope for a better indication of the actual penetration of (1) Formulation and Materials Design Aspects
chlorides into AAMs in service, as these tests involve the 1. Availability of admixtures, organic or inorganic, which
use of a colorimetric approach (spraying of tested speci- will enable enhanced workability of AAMs at low
mens with silver nitrate) to determine the chloride penetra- water content.66
tion depth. This can then be used to calculate diffusion 2. Identification and development of alternative precur-
coefficients via consideration of the chemistry of the pore sors which are (a) not waste streams from other indus-
solution (particularly Cl /OH ratio) in the alkali-activated tries and thus subject to quality control issues, and (b)
binder system.33 not in such high demand from the portland-blended
Recent results33 have demonstrated that the nature of the bin- cement industry as slag and fly ash are in many parts
der gel strongly influences the chloride permeability in alkali- of the world at present.
activated slag/fly ash systems. This is a combined effect of the 3. Description of the corrosion chemistry of mild steel
space-filling characteristics of the gels, where C–A–S–H type reinforcing elements within the pore solution environ-
gels are less porous than N–A–S–H “geopolymer” type bind- ments of alkali-activated binders based on either fly
ers,34 and the capacity of each of the gels to bind chloride ions. ash or slag, considering the differences in ionic envi-
In high-Ca systems, chloride ions can not only be chemisorbed ronments and redox potential between these systems.
onto C–A–S–H products in the layers and interlayer spaces but
can also be chemically bound to the C–S–H.105 In portland
cement/slag binders, it has been observed that the layered dou-
ble hydroxides forming as secondary reaction products play a (2) Processing and Testing Aspects
significant role in determining the chloride permeability, as these 1. Adequate control of curing conditions to minimize
phases have the capacity to adsorb chlorides.106 The sorption of shrinkage and cracking, including particularly the area
chlorides on the secondary layered double hydroxide phases of shrinkage-controlling admixtures63,121—the cracking
forming in AAMs has not yet been studied in detail; however, tendency under applied load or restrained shrinkage in
hydrotalcite is the main secondary reaction product formed in alkali-activated slag concretes has been assessed as
most alkali-activated slag binders,11,107 and can ion exchange being “comparable” to the performance of portland
chlorides for hydroxides to some extent.108 Therefore, it is highly cement-based materials,122,123 and it could thus be
possible that this phase plays a comparable chloride-binding expected that similar design codes should be able to be
role in alkali-activated slags. used for both types of materials, but this requires fur-
In low-Ca alkali-activated systems, a higher degree of ther validation.
direct chloride binding than takes place in Ca-rich systems 2. Understanding of long-term properties such as chloride
has been observed.33 This is associated with the higher sur- diffusion coefficients, and particularly how they can be
face area of the N–A–S–H type gel when compared with predicted from laboratory tests, with the knowledge
C–A–S–H gel,32 which favors a higher degree of physical that the time-temperature-maturity relationships for
absorption of chlorides in the low-calcium gels, and there the microstructure of an alkali-activated material may
may also be chemical effects introduced by the differences in be very different from the relationships in effect for a
(alumino)silicate framework charge, but these have not yet portland cement-based material.
been explored in detail. In alkali-activated fly ash systems,
precipitation of halite (NaCl) has been identified in dried
paste specimens taken from regions where chloride has pene-
(3) Service Life Prediction Aspects
trated,33 indicating that the N–A–S–H gels are likely to be
hosting a chloride-rich pore solution. 1. The availability of validated laboratory testing proce-
Consideration of the nature of the binder gels forming in dures to enable the analysis of key degradation mecha-
AAMs is then a key factor in predicting chloride penetra- nisms on an accelerated timescale.
tion into reinforced alkali-activated concretes, as enhanced 2. The integration of these testing outcomes into a vali-
binding of chlorides to the reaction products in the system dated model for the service life of alkali-activated con-
will reduce the content of free chlorides that can migrate to cretes under various types of real-world service
promote corrosion processes in the steel rebar. This then conditions, particularly including the simultaneous
provides further scope for the design of highly durable influences of multiple modes of degradation.
binders through the analysis and tailoring of binder gel These are all areas of active ongoing research at this point
chemistry. in time, and are likely to remain so for many years to come
providing ample scope for major advancements to be made
V. Perspectives in the science and technology of these materials in coming
years.
In general, and particularly looking to extend the discussion
from mainly paste-related factors to also include some
VI. Conclusions
comments on the understanding of alkali-activated concretes:
the various influences on concrete strength and durability of The durability of AAMs is strongly dependent on the nano-
parameters such as paste content,76,79,109,110 precursor and microstructure of the reaction products forming in these
chemistry,61,111 precursor blending,29,112 water/binder systems, as a function of the type of precursor, the nature
76,110,113,114
ratio, activator chemistry (dose and composi- and concentration of the activator, and also the maturity of
tion),100,109,114–116 curing conditions113,117, and fiber rein- the material. High-Ca systems have a structure mainly domi-
forcement118,119 are now becoming better understood. It is nated by a C–A–S–H gel, which is less porous than the “geo-
also being more widely identified that drying-related micro- polymer” gel forming in low-Ca systems. This is one of the
cracking is a key issue which requires careful control,115,120 main factors controlling the transport properties of AAMs;
and that this is essential in achieving good overall perfor- however, the differences in chemistry of both pore solution
mance—particularly in terms of superficial appearance, and reaction products in AAMs produced with different pre-
which is often just as important as actual material perfor- cursors modify the chemical mechanisms that can lead to the
mance in achieving commercial acceptance for a material. decay of these binders when exposed to aggressive environ-
The key areas with need for future work related to the tech- ments.
nology of alkali-activated binders, and particularly the inter- Exposure to sulfate salts can sometimes negatively impact
relationships between microstructure and durability, include the structure of AAMs, depending on the cation accompany-
(but are by no means limited to) the following. ing the salt. Immersion of alkali-activated specimens in
1006 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4

Na2SO4 favors the structural evolution and increasing matu- the Nature of the Calcium Silicate Hydrate (C-S-H) Paste,” Cem. Concr. Res.,
rity of the binding phases, and does not have any apparent 24 [5] 813–29 (1994).
10
F. Puertas, M. Palacios, H. Manzano, J. S. Dolado, A. Rico, and J.
negative effect on the material. On the other hand, the expo- Rodrıguez, “A Model for the C-A-S-H Gel Formed in Alkali-Activated Slag
sure of AAMs to MgSO4 promotes the destruction of the Cements,” J. Eur. Ceram. Soc., 31 [12] 2043–56 (2011).
11
main reaction products via a cation-exchange mechanism, S. A. Bernal, J. L. Provis, B. Walkley, R. San Nicolas, J. D. Gehman, D.
which leads to the formation of M–S–H type gels. Special G. Brice, A. Kilcullen, P. Duxson, and J. S. J. van Deventer, “Gel Nanostruc-
ture in Alkali-Activated Binders Based on Slag and Fly Ash, and Effects of
attention should be given to the cations present in sulfate- Accelerated Carbonation,” Cem. Concr. Res., 53, 127–44 (2013).
rich environments when using alkali-activated binders. 12
H. Rahier, B. van Mele, M. Biesemans, J. Wastiels, and X. Wu, “Low-
Carbonation is also controlled by the nature of the bind- Temperature Synthesized Aluminosilicate Glasses. 1. Low-Temperature Reac-
ing phases, and more importantly by the chemistry of the tion Stoichiometry and Structure of a Model Compound,” J. Mater. Sci., 31
[1] 71–9 (1996).
pore solution. Under natural carbonation conditions, AAMs 13
J. L. Provis, G. C. Lukey, and J. S. J. van Deventer, “Do Geopolymers
present comparable carbonation rates to portland cements; Actually Contain Nanocrystalline Zeolites? - A Reexamination of Existing
however, accelerated carbonation testing conditions can Results,” Chem. Mater., 17 [12] 3075–85 (2005).
14
induce the rapid decay of an alkali-activated material. For P. Duxson, G. C. Lukey, F. Separovic, and J. S. J. van Deventer, “The
Effect of Alkali Cations on Aluminum Incorporation in Geopolymeric Gels,”
reliable testing of the susceptibility of AAMs to carbonation, Ind. Eng. Chem. Res., 44 [4] 832–9 (2005).
a better understanding of the effect of exposure conditions 15
A. Palomo, S. Alonso, A. Fernandez-Jimenez, I. Sobrados, and J. Sanz,
on the material needs to be developed, to enable selection “Alkaline Activation of Fly Ashes: NMR Study of the Reaction Products,” J.
of adequate testing conditions to properly simulate service Am. Ceram. Soc., 87 [6] 1141–5 (2004).
16
A. Palomo and F. P. Glasser, “Chemically-Bonded Cementitious Materials
conditions. Based on Metakaolin,” Br. Ceram. Trans. J., 91 [4] 107–12 (1992).
AAMs are well known for their reduced chloride penetra- 17
J. Davidovits, “Geopolymers - Inorganic Polymeric New Materials,” J.
tion rates, which have in some instances been attributed to Therm. Anal., 37 [8] 1633–56 (1991).
18
their reduced overall permeability compared with portland I. Ismail, S. A. Bernal, J. L. Provis, R. San Nicolas, S. Hamdan, and J. S.
J. van Deventer, “Modification of Phase Evolution in Alkali-Activated Blast
cements. However, it has been demonstrated that this is a Furnace Slag by the Incorporation of Fly Ash,” Cem. Concr. Compos., 45,
combined effect not only depending on the bulk transport 125–35 (2014).
19
properties of the system, as the chemistry of the binding C. K. Yip and J. S. J. van Deventer, “Microanalysis of Calcium Silicate
phases strongly influences the binding of chlorides, and, Hydrate Gel Formed within a Geopolymeric Binder,” J. Mater. Sci., 38 [18]
3851–60 (2003).
therefore, the time taken before corrosion of the steel rebars 20
A. Buchwald, H. Hilbig, and C. Kaps, “Alkali-Activated Metakaolin-Slag
in reinforced structures is initiated. Blends – Performance and Structure in Dependence on Their Composition,” J.
In general, although scientific knowledge in this area is Mater. Sci., 42 [9] 3024–32 (2007).
21
increasing rapidly, a deeper understanding of the mechanisms K. Dombrowski, A. Buchwald, and M. Weil, “The Influence of Calcium
Content on the Structure and Thermal Performance of Fly Ash Based Geo-
inducing structural changes in AAMs still needs to be devel- polymers,” J. Mater. Sci., 42 [9] 3033–43 (2007).
oped, based on enhanced microstructural and nanostructural 22
I. Garcıa-Lodeiro, A. Fernandez-Jimenez, A. Palomo, and D. E. Macphee,
descriptions of the binder and its interaction with different “Effect of Calcium Additions on N–A–S–H Cementitious Gels,” J. Am.
environments. This will enable selection of the correct testing Ceram. Soc., 93 [7] 1934–40 (2010).
23
I. Garcıa-Lodeiro, A. Palomo, A. Fernandez-Jimenez, and D. E. Macphee,
conditions to enable accurate evaluation of the performance “Compatibility Studies between N-A-S-H and C-A-S-H Gels. Study in the Ternary
of these materials under different aggressive conditions, and Diagram Na2O-CaO-Al2O3-SiO2-H2O,” Cem. Concr. Res., 41 [9] 923–31 (2011).
24
also to generate limit state criteria and service life models to R. R. Lloyd, J. L. Provis, and J. S. J. van Deventer, “Microscopy and
aid in the formulation of AAMs with higher resistance to Microanalysis of Inorganic Polymer Cements. 1: Remnant Fly Ash Particles,”
J. Mater. Sci., 44 [2] 608–19 (2009).
specific decay processes, depending on the application for 25
S. Alonso and A. Palomo, “Alkaline Activation of Metakaolin and Cal-
which the material is being designed. cium Hydroxide Mixtures: Influence of Temperature, Activator Concentration
and Solids Ratio,” Mater. Lett., 47 [1–2] 55–62 (2001).
26
J. L. Provis, V. Rose, S. A. Bernal, and J. S. J. van Deventer, “High Reso-
Acknowledgments lution Nanoprobe X-ray Fluorescence Characterization of Heterogeneous Cal-
cium and Heavy Metal Distributions in Alkali Activated Fly Ash,” Langmuir,
We are grateful to the colleagues who have generously provided graphics to be 25 [19] 11897–904 (2009).
27
used in this article: Dr. Rackel San Nicolas (U. Melbourne), Dr. Erich I. Lecomte, M. Liegeois, A. Rulmont, R. Cloots, and F. Maseri, “Synthe-
Rodrıguez (U. del Valle), and Dr. Idawati Ismail (U. Melbourne/U. Malaysia sis and Characterization of New Inorganic Polymeric Composites Based on
Sarawak). The authors are supported by funding from the European Research Kaolin or White Clay and on Ground-Granulated Blast Furnace Slag,” J.
Council, Starting grant no. #335928 (GeopolyConc), and from the University Mater. Res., 18 [11] 2571–9 (2003).
28
of Sheffield. Y. Zhang, W. Sun, Q. Chen, and L. Chen, “Synthesis and Heavy Metal
Immobilization Behaviors of Slag Based Geopolymer,” J. Hazard. Mater., 143
[1–2] 206–13 (2007).
29
References S. A. Bernal, R. Mejıa de Gutierrez, and J. L. Provis, “Engineering and
Durability Properties of Concretes Based on Alkali-Activated Granulated
1
B. Lothenbach, K. Scrivener, and R. D. Hooton, “Supplementary Cementi- Blast Furnace Slag/Metakaolin Blends,” Constr. Build. Mater., 33, 99–108
tious Materials,” Cem. Concr. Res., 41 [12] 1244–56 (2011). (2012).
2 30
R. Snellings, G. Mertens, and J. Elsen, “Supplementary Cementitious S. A. Bernal, E. D. Rodrıguez, R. Mejıa de Gutierrez, M. Gordillo, and J. L.
Materials,” Rev. Mineral. Geochem., 74, 211–78 (2012). Provis, “Mechanical and Thermal Characterisation of Geopolymers Based on Sili-
3
M. C. G. Juenger, F. Winnefeld, J. L. Provis, and J. Ideker, “Advances in cate-Activated Metakaolin/Slag Blends,” J. Mater. Sci., 46 [16] 5477–86 (2011).
31
Alternative Cementitious Binders,” Cem. Concr. Res., 41 [12] 1232–43 (2011). S. A. Bernal, R. Mejıa de Gutierrez, F. Ruiz, H. Qui~ nones, and J. L. Pro-
4
J. L. Provis and J. S. J. van Deventer (eds.), Alkali-Activated Materials: vis, “Desempe~ no a Temperaturas Altas de Morteros y Hormigones Basados
State-of-the-Art Report, Rilem TC 224-AAM. Springer/RILEM, Dordrecht, en Mezclas de Escoria/Metacaolın Activadas Alcalinamente (High-Tempera-
The Netherlands, 2014. ture Performance of Mortars and Concretes Based on Alkali-Activated Slag/
5
J. S. J. van Deventer, J. L. Provis, and P. Duxson, “Technical and Com- Metakaolin Blends),” Mater. Constr., 308 [62] 471–88 (2012).
32
mercial Progress in the Adoption of Geopolymer Cement,” Miner. Eng., 29, I. Ismail, S. A. Bernal, J. L. Provis, S. Hamdan, and J. S. J. van Deventer,
89–104 (2012). “Drying-Induced Changes in the Structure of Alkali-Activated Pastes,” J.
6
J. L. Provis, “Geopolymers and Other Alkali Activated Materials - Why, Mater. Sci., 48 [9] 3566–77 (2013).
33
How, and What?,” Mater. Struct. (2014). (in press). doi: 10.1617/s11527- I. Ismail, S. A. Bernal, J. L. Provis, R. San Nicolas, D. G. Brice, A. R.
013-0211-5 Kilcullen, S. Hamdan, and J. S. J. van Deventer, “Influence of Fly Ash on the
7
J. L. Provis, V. Bılek, A. Buchwald, K. Dombrowski-Daube, and B. Varel- Water and Chloride Permeability of Alkali-Activated Slag Mortars and Con-
a, “Durability and Testing - Physical Processes,” pp. 277–307 in Alkali-Acti- cretes,” Constr. Build. Mater., 48, 1187–201 (2013).
34
vated Materials: State-of-the-Art Report, Rilem TC 224-AAM, Edited by J. L. J. L. Provis, R. J. Myers, C. E. White, V. Rose, and J. S. J. van Deventer,
Provis and J. S. J. van Deventer. Springer/RILEM, Dordrecht, The Nether- “X-ray Microtomography Shows Pore Structure and Tortuosity in Alkali-Acti-
lands. 2014. vated Binders,” Cem. Concr. Res., 42 [6] 855–64 (2012).
8 35
R. J. Myers, S. A. Bernal, R. San Nicolas, and J. L. Provis, “Generalized Struc- A. Favier, G. Habert, J. B. d’Espinose de Lacaillerie, and N. Roussel,
tural Description of Calcium-Sodium Aluminosilicate Hydrate Gels: The Cross- “Mechanical Properties and Compositional Heterogeneities of Fresh Geopoly-
linked Substituted Tobermorite Model,” Langmuir, 29 [17] 5294–306 (2013). mer Pastes,” Cem. Concr. Res., 48, 9–16 (2013).
9 36
I. G. Richardson, A. R. Brough, G. W. Groves, and C. M. Dobson, “The S. Diamond, “Particle Morphologies in Fly Ash,” Cem. Concr. Res., 16 [4]
Characterization of Hardened Alkali-Activated Blast-Furnace Slag Pastes and 569–79 (1986).
April 2014 Durability of Alkali‐Activated Materials 1007
37 64
J. C. Qian, E. E. Lachowski, and F. P. Glasser, “Microstructure and M. Palacios, P. F. G. Banfill, and F. Puertas, “Rheology and Setting of
Chemical Variation in Class F Fly Ash Glass;” pp. 45–54 in Fly Ash and Coal Alkali-Activated Slag Pastes and Mortars: Effect of Organic Admixture,” ACI
Conversion By-Products: Characterization, Utilization, and Disposal IV, Vol. Mater. J., 105 [2] 140–8 (2008).
65
113, Edited by G. J. McCarthy, F. P. Glasser and D. M. Roy. Materials A. Kashani, J. L. Provis, J. Xu, A. Kilcullen, P. Duxson, G. G. Qiao, and
Research Society Symposium Proceedings, Warrendale, PA, 1988. J. S. J. van Deventer, “Effects of Different Polycarboxylate Ether Structures
38
H. S. Pietersen, S. P. Vriend, R. E. P. Poorter, and J. M. Bijen, “Micro- on the Rheology of Alkali Activated Slag Binders,” pp. 279–89 in Tenth Inter-
probe Study of Chemical Inter-Grain Variation in Class F Fly Ash: Applica- national Conference on Superplasticizers and Other Chemical Admixtures in
tion of Fuzzy C-Means Cluster Analysis and Non-Linear Mapping;” pp. 115– Concrete (American Concrete Institute Special Publication SP-288), Prague,
26 in Fly Ash and Coal Conversion By-Products: Characterization, Utilization, Czech Republic. 2012.
66
and Disposal VI, Vol. 178, Edited by R. L. Day and F. P. Glasser. Materials D. Marchon, U. Sulser, A. Eberhardt, and R. J. Flatt, “Molecular Design
Research Society Symposium Proceedings, Pittsburgh, PA, 1990. of Comb-Shaped Polycarboxylate Dispersants for Environmentally Friendly
39
B. Golosio, A. Simionovici, A. Somogyi, C. Camerani, and B. M. Stee- Concrete,” Soft Matter, 9, 10719–28 (2013).
67
nari, “X-ray Fluorescence Tomography of Individual Fly Ash Particles,” J. M. Palacios, P. Bowen, M. Kappl, H. J. Butt, M. Stuer, C. Pecharroman,
Phys. IV Fr., 104, 647–50 (2003). U. Aschauer, and F. Puertas, “Fuerzas de Repulsi on de Aditivos Superplastifi-
40
J. Duchesne, L. Duong, T. Bostrom, and R. Frost, “Microstructure Study cantes en Sistemas de Escoria Granulada de Horno Alto en Medios Alcalinos,
of Early in Situ Reaction of Fly Ash Geopolymer Observed by Environmental Desde Medidas de AFM a Propiedades Reol ogicas (Repulsion Forces of
Scanning Electron Microscopy (ESEM),” Waste Biomass Valoriz., 1 [3] 367–77 Superplasticizers on Ground Granulated Blast Furnace Slag in Alkaline
(2010). Media, from AFM Measurements to Rheological Properties),” Mater. Constr.,
41
R. R. Lloyd, J. L. Provis, and J. S. J. van Deventer, “Microscopy and 62 [308] 489–513 (2012).
68
Microanalysis of Inorganic Polymer Cements. 2: The Gel Binder,” J. Mater. A. Kashani, J. L. Provis, J. Xu, A. R. Kilcullen, G. G. Qiao, and J. S. J.
Sci., 44 [2] 620–31 (2009). van Deventer, “Effect of Molecular Architecture of Polycarboxylate Ethers on
42
J. L. Provis, V. Rose, R. P. Winarski, and J. S. J. van Deventer, “Hard X- Plasticizing Performance in Alkali Activated Slag Paste,” J. Mater. Sci. (2014).
ray Nanotomography of Amorphous Aluminosilicate Cements,” Scripta (in press), doi: 10.1007/s10853-013-7979-0.
69
Mater., 65 [4] 316–9 (2011). B. Huet, V. L’Hostis, F. Miserque, and H. Idrissi, “Electrochemical
43
R. E. Davis, R. W. Carlson, J. W. Kelly, and H. E. Davis, “Properties of Behavior of Mild Steel in Concrete: Influence of pH and Carbonate Content
Cements and Concretes Containing Fly Ash,” J. Am. Concr. Inst., 33, 577–612 of Concrete Pore Solution,” Electrochim. Acta, 51 [1] 172–80 (2005).
70
(1937). R. R. Lloyd, J. L. Provis, and J. S. J. van Deventer, “Pore Solution Com-
44
A. Bilodeau and V. M. Malhotra, “High-Volume Fly Ash System: Concrete position and Alkali Diffusion in Inorganic Polymer Cement,” Cem. Concr.
Solution for Sustainable Development,” ACI Mater. J., 97 [1] 41–8 (2000). Res., 40 [9] 1386–92 (2010).
45 71
S. Donatello, A. Fernandez-Jimenez, and A. Palomo, “Very High Volume R. R. Lloyd, “Accelerated Ageing of Geopolymers;” pp. 139–66 in Geo-
Fly Ash Cements. Early Age Hydration Study Using Na2SO4 as an Activa- polymers: Structure, Processing, Properties and Industrial Applications, Edited
tor,” J. Am. Ceram. Soc., 96 [3] 900–6 (2013). by J. L. Provis and J. S. J. van Deventer. Woodhead, Cambridge, UK, 2009.
46 72
A. Fern andez-Jimenez, A. Palomo, and M. Criado, “Microstructure A. R. Sakulich and D. P. Bentz, “Mitigation of Autogenous Shrinkage in
Development of Alkali-Activated Fly Ash Cement: A Descriptive Model,” Alkali Activated Slag Mortars by Internal Curing,” Mater. Struct., 46 [8]
Cem. Concr. Res., 35 [6] 1204–9 (2005). 1355–67 (2013).
47 73
C. A. Rees, J. L. Provis, G. C. Lukey, and J. S. J. van Deventer, “Attenu- ACI Committee 201, “Guide to Durable Concrete,” American Concrete
ated Total Reflectance Fourier Transform Infrared Analysis of Fly Ash Geo- Institute, Farmington Hills, MI, 1982.
74
polymer Gel Aging,” Langmuir, 23 [15] 8170–9 (2007). M. Fernandez-Bertos, S. J. R. Simons, C. D. Hills, and P. J. Carey, “A
48
C. A. Rees, J. L. Provis, G. C. Lukey, and J. S. J. van Deventer, “In Situ Review of Accelerated Carbonation Technology in the Treatment of Cement-
ATR-FTIR Study of the Early Stages of Fly Ash Geopolymer Gel Forma- Based Materials and Sequestration of CO2,” J. Hazard. Mater., 112, 193–205
tion,” Langmuir, 23 [17] 9076–82 (2007). (2004).
49 75
C. A. Rees, J. L. Provis, G. C. Lukey, and J. S. J. van Deventer, “The D. W. Hobbs, “Concrete Deterioration: Causes, Diagnosis, and Minimis-
Mechanism of Geopolymer Gel Formation Investigated through Seeded ing Risk,” Int. Mater. Rev., 46 [3] 117–44 (2001).
76
Nucleation,” Colloids Surf. A, 318 [1–3] 97–105 (2008). S. A. Bernal, R. San Nicolas, J. L. Provis, R. Mejıa de Gutierrez, and J.
50
A. Hajimohammadi, J. L. Provis, and J. S. J. van Deventer, “Time- S. J. van Deventer, “Natural Carbonation of Aged Alkali-Activated Slag Con-
Resolved and Spatially-Resolved Infrared Spectroscopic Observation of Seeded cretes,” Mater. Struct. (2014). (in press). doi: 10.1617/s11527-013-0089-2.
77
Nucleation Controlling Geopolymer Gel Formation,” J. Colloid Interface Sci., M. Palacios and F. Puertas, “Effect of Carbonation on Alkali-Activated
357 [2] 384–92 (2011). Slag Paste,” J. Am. Ceram. Soc., 89 [10] 3211–21 (2006).
51 78
S. A. Bernal, J. L. Provis, R. Mejıa de Gutierrez, and V. Rose, “Evolution S. A. Bernal, R. San Nicolas, R. J. Myers, R. Mejıa de Gutierrez, F. Puer-
of Binder Structure in Sodium Silicate-Activated Slag-Metakaolin Blends,” tas, J. S. J. van Deventer, and J. L. Provis, “MgO Content of Slag Controls
Cem. Concr. Compos., 33 [1] 46–54 (2011). Phase Evolution and Structural Changes Induced by Accelerated Carbonation
52
S. D. Wang and K. L. Scrivener, “Hydration Products of Alkali-Activated in Alkali-Activated Binders,” Cem. Concr. Res., 57, 33–43(2014).
79
Slag Cement,” Cem. Concr. Res., 25 [3] 561–71 (1995). S. A. Bernal, R. Mejıa de Gutierrez, A. L. Pedraza, J. L. Provis, E. D.
53
S. A. Bernal, J. L. Provis, V. Rose, and R. Mejıa de Gutierrez, “High-Res- Rodrıguez, and S. Delvasto, “Effect of Binder Content on the Performance of
olution X-ray Diffraction and Fluorescence Microscopy Characterization of Alkali-Activated Slag Concretes,” Cem. Concr. Res., 41 [1] 1–8 (2011).
80
Alkali-Activated Slag-Metakaolin Binders,” J. Am. Ceram. Soc., 96 [6] 1951–7 F. G. da Silva, P. Helene, P. Castro-Borges, and J. B. L. Liborio, “Sources
(2013). of Variations When Comparing Concrete Carbonation Results,” J. Mater.
54
I. G. Richardson, “Tobermorite/Jennite- and Tobermorite/Calcium Civ. Eng., 21 [7] 333–42 (2009).
81
Hydroxide-Based Models for the Structure of C-S-H: Applicability to Hard- S. A. Bernal, R. Mejıa de Gutierrez, V. Rose, and J. L. Provis, “Effect of
ened Pastes of Tricalcium Silicate, Β-Dicalcium Silicate, Portland Cement, and Silicate Modulus and Metakaolin Incorporation on the Carbonation of Alkali
Blends of Portland Cement with Blast-Furnace Slag, Metakaolin, or Silica Silicate-Activated Slags,” Cem. Concr. Res., 40 [6] 898–907 (2010).
82
Fume,” Cem. Concr. Res., 34 [9] 1733–77 (2004). F. Puertas, M. Palacios, and T. Vazquez, “Carbonation Process of Alkali-
55
A. Gruskovnjak, B. Lothenbach, L. Holzer, R. Figi, and F. Winnefeld, Activated Slag Mortars,” J. Mater. Sci., 41, 3071–82 (2006).
83
“Hydration of Alkali-Activated Slag: Comparison with Ordinary Portland H. Xu, J. L. Provis, J. S. J. van Deventer, and P. V. Krivenko, “Charac-
Cement,” Adv. Cem. Res, 18 [3] 119–28 (2006). terization of Aged Slag Concretes,” ACI Mater. J., 105 [2] 131–9 (2008).
56 84
A. R. Brough and A. Atkinson, “Sodium Silicate-Based, Alkali-Activated S. A. Bernal, J. L. Provis, A. Fernandez-Jimenez, P. V. Krivenko, E.
Slag Mortars: Part I. Strength, Hydration and Microstructure,” Cem. Concr. Kavalerova, M. Palacios, and C. Shi, “Binder Chemistry – High-Calcium
Res., 32 [6] 865–79 (2002). Alkali-Activated Materials,” pp. 59–91 in Alkali-Activated Materials: State-of-
57
A. A. Melo Neto, M. A. Cincotto, and W. Repette, “Drying and Autoge- the-Art Report, Rilem TC 224-AAM, Edited by J. L. Provis and J. S. J. van
nous Shrinkage of Pastes and Mortars with Activated Slag Cement,” Cem. Deventer. Springer/RILEM, Dordrecht, 2014.
85
Concr. Res., 38, 565–74 (2008). D. W. S. Ho and R. K. Lewis, “Carbonation of Concrete and Its Predic-
58
R. R. Lloyd, J. L. Provis, K. J. Smeaton, and J. S. J. van Deventer, “Spatial tion,” Cem. Concr. Res., 17 [3] 489–504 (1987).
86
Distribution of Pores in Fly Ash-Based Inorganic Polymer Gels Visualised by S. Song and H. M. Jennings, “Pore Solution Chemistry of Alkali-Acti-
Wood’s Metal Intrusion,” Microporous Mesoporous Mater., 126 [1–2] 32–9 (2009). vated Ground Granulated Blast-Furnace Slag,” Cem. Concr. Res., 29, 159–70
59
P. Duxson, J. L. Provis, G. C. Lukey, S. W. Mallicoat, W. M. Kriven, (1999).
87
and J. S. J. van Deventer, “Understanding the Relationship between Geopoly- F. Puertas, A. Fernandez-Jimenez, and M. T. Blanco-Varela, “Pore Solu-
mer Composition, Microstructure and Mechanical Properties,” Colloids Surf. tion in Alkali-Activated Slag Cement Pastes. Relation to the Composition and
A, 269 [1–3] 47–58 (2005). Structure of Calcium Silicate Hydrate,” Cem. Concr. Res., 34, 139–48 (2004).
60 88
J. L. Provis, R. M. Harrex, S. A. Bernal, P. Duxson, and J. S. J. van H. P. Eugster, “Sodium Carbonate-Bicarbonate Minerals as Indicators of
Deventer, “Dilatometry of Geopolymers as a Means of Selecting Desirable Fly PCO2,” J. Geophys. Res., 71 [14] 3369–77 (1966).
89
Ash Sources,” J. Non-Cryst. Solids, 358 [16] 1930–7 (2012). S. A. Bernal, J. L. Provis, D. G. Brice, A. Kilcullen, P. Duxson, and J. S.
61
C. Shi, P. V. Krivenko, and D. M. Roy, Alkali-Activated Cements and J. van Deventer, “Accelerated Carbonation Testing of Alkali-Activated Bind-
Concretes. Taylor & Francis, Abingdon, UK, 2006. ers Significantly Underestimates Service Life: The Role of Pore Solution
62
D. Krizan and B. Zivanovic, “Effects of Dosage and Modulus of Water Chemistry,” Cem. Concr. Res., 42 [10] 1317–26 (2012).
90
Glass on Early Hydration of Alkali–Slag Cements,” Cem. Concr. Res., 32 [8] M. Criado, A. Fernandez-Jimenez, and A. Palomo, “Alkali Activation of
1181–8 (2002). Fly Ash. Part III: Effect of Curing Conditions on Reaction and its Graphical
63
M. Palacios and F. Puertas, “Effect of Superplasticizer and Shrinkage- Description,” Fuel, 89 [11] 3185–92 (2010).
91
Reducing Admixtures on Alkali-Activated Slag Pastes and Mortars,” Cem. T. Bakharev, J. G. Sanjayan, and Y. B. Cheng, “Sulfate Attack on Alkali-
Concr. Res., 35 [7] 1358–67 (2005). Activated Slag Concrete,” Cem. Concr. Res., 32 [2] 211–6 (2002).
1008 Journal of the American Ceramic Society—Bernal and Provis Vol. 97, No. 4
92
I. Ismail, S. A. Bernal, J. L. Provis, S. Hamdan, and J. S. J. van Deventer, Microstructural Development of Alkali Activated Blast-Furnace Slags,” Cem.
“Microstructural Changes in Alkali Activated Fly Ash/Slag Geopolymers with Concr. Res., 41 [3] 301–10 (2011).
108
Sulfate Exposure,” Mater. Struct., 46 [3] 361–73 (2013). S. Miyata, “Anion-Exchange Properties of Hydrotalcite-Like Com-
93
S. Thokchom, P. Ghosh, and S. Ghosh, “Performance of Fly Ash Based Geo- pounds,” Clays Clay Miner., 31 [4] 305–11 (1983).
109
polymer Mortars in Sulphate Solution,” J. Eng. Sci. Technol., 3 [1] 36–40 (2010). A. Fernandez-Jimenez and A. Palomo, “Factors Affecting Early Com-
94
S. Thokchom, P. Ghosh, and S. Ghosh, “Effect of Na2O Content on pressive Strength of Alkali Activated Fly Ash (OPC-Free) Concrete,” Mater.
Durability of Geopolymer Pastes in Magnesium Sulfate Solution,” Can. J. Constr., 57 [287] 7–22 (2007).
110
Civ. Eng., 39 [1] 34–43 (2011). T. H€akkinen, “The Influence of Slag Content on the Microstructure, Per-
95
T. Bakharev, “Durability of Geopolymer Materials in Sodium and Magne- meability and Mechanical Properties of Concrete: Part 2. Technical Properties
sium Sulfate Solutions,” Cem. Concr. Res., 35 [6] 1233–46 (2005). and Theoretical Examinations,” Cem. Concr. Res., 23 [3] 518–30 (1993).
96 111
P. Chindaprasirt, U. Rattanasak, and S. Taebuanhuad, “Resistance to E. I. Diaz-Loya, E. N. Allouche, and S. Vaidya, “Mechanical Properties
Acid and Sulfate Solutions of Microwave-Assisted High Calcium Fly Ash of Fly-Ash-Based Geopolymer Concrete,” ACI Mater. J., 108 [3] 300–6 (2011).
112
Geopolymer,” Mater. Struct., 46 [3] 375–81 (2013). J. Wongpa, K. Kiattikomol, C. Jaturapitakkul, and P. Chindaprasirt,
97
D. M. Roy, W. Jiang, and M. R. Silsbee, “Chloride Diffusion in Ordinary, “Compressive Strength, Modulus of Elasticity, and Water Permeability of
Blended, and Alkali-Activated Cement Pastes and Its Relation to Other Prop- Inorganic Polymer Concrete,” Mater. Des., 31 [10] 4748–54 (2010).
113
erties,” Cem. Concr. Res., 30, 1879–84 (2000). D. Bondar, C. J. Lynsdale, N. B. Milestone, N. Hassani, and A. A. Ram-
98
V. Saraswathy, S. Muralidharan, K. Thangavel, and S. Srinivasan, “Influ- ezanianpour, “Engineering Properties of Alkali-Activated Natural Pozzolan
ence of Activated Fly Ash on Corrosion-Resistance and Strength of Con- Concrete,” ACI Mater. J., 108 [1] 64–72 (2011).
114
crete,” Cem. Concr. Compos., 25 [7] 673–80 (2003). E. Douglas, A. Bilodeau, and V. M. Malhotra, “Properties and Durabil-
99
J. M. Miranda, A. Fern andez-Jimenez, J. A. Gonzalez, and A. Palomo, ity of Alkali-Activated Slag Concrete,” ACI Mater. J., 89 [5] 509–16 (1992).
115
“Corrosion Resistance in Activated Fly Ash Mortars,” Cem. Concr. Res., 35 R. Anderson and H.-E. Gram, “Properties of Alkali-Activated Slag Con-
[6] 1210–7 (2005). crete,” Nord. Concr. Res., 6, 7–18 (1987).
100 116
D. Law, A. Adam, T. Molyneaux, and I. Patnaikuni, “Durability Assess- S. D. Wang, “Review of Recent Research on Alkali-Activated Concrete
ment of Alkali Activated Slag (AAS) Concrete,” Mater. Struct., 45 [9] 1425–37 in China,” Mag. Concr. Res., 43 [154] 29–35 (1991).
117
(2012). T. Bakharev, J. G. Sanjayan, and Y. B. Cheng, “Effect of Elevated Tem-
101
D. Ravikumar and N. Neithalath, “Electrically Induced Chloride Ion perature Curing on Properties of Alkali-Activated Slag Concrete,” Cem. Con-
Transport in Alkali Activated Slag Concretes and the Influence of Microstruc- cr. Res., 29 [10] 1619–25 (1999).
118
ture,” Cem. Concr. Res., 47, 31–42 (2013). S. Bernal, R. Mejıa de Gutierrez, E. Rodriguez, S. Delvasto, and F. Puer-
102
D. Ravikumar and N. Neithalath, “An Electrical Impedance Investigation tas, “Mechanical Behaviour of Steel Fibre-Reinforced Alkali Activated Slag
into the Chloride Ion Transport Resistance of Alkali Silicate Powder Activated Concrete,” Mater. Constr., 59 [293] 53–62 (2009).
119
Slag Concretes,” Cem. Concr. Compos., 44, 58–68 (2013). D. Penteado Dias and C. Thaumaturgo, “Fracture Toughness of Geo-
103
K. Kupwade-Patil and E. N. Allouche, “Examination of Chloride polymeric Concretes Reinforced with Basalt Fibers,” Cem. Concr. Compos., 27
Induced Corrosion in Reinforced Geopolymer Concretes,” J. Mater. Civ. [1] 49–54 (2005).
120
Eng., 25 [10] 1465–76 (2013). F. Collins and J. G. Sanjayan, “Microcracking and Strength Develop-
104
C. Shi, Another Look at the Rapid Chloride Permeability Test (ASTM ment of Alkali Activated Slag Concrete,” Cem. Concr. Compos., 23 [4–5] 345–
C1202 or ASSHTO T277). FHWA Resource Center, Baltimore, MD, 2003. 52 (2001).
105 121
Q. Yuan, C. Shi, G. De Schutter, K. Audenaert, and D. Deng, “Chloride M. Palacios and F. Puertas, “Effect of Shrinkage-Reducing Admixtures
Binding of Cement-Based Materials Subjected to External Chloride Environ- on the Properties of Alkali-Activated Slag Mortars and Pastes,” Cem. Concr.
ment – A Review,” Constr. Build. Mater., 23 [1] 1–13 (2009). Res., 37 [5] 691–702 (2007).
106 122
O. Kayali, M. S. H. Khan, and M. Sharfuddin Ahmed, “The Role of F. Collins and J. G. Sanjayan, “Cracking Tendency of Alkali-Activated
Hydrotalcite in Chloride Binding and Corrosion Protection in Concretes with Slag Concrete Subjected to Restrained Shrinkage,” Cem. Concr. Res., 30 [5]
Ground Granulated Blast Furnace Slag,” Cem. Concr. Compos., 34 [8] 936–45 791–8 (2000).
123
(2012). J. R. Yost, A. Radli nska, S. Ernst, M. Salera, and N. J. Martignetti,
107
M. Ben Haha, G. Le Saout, F. Winnefeld, and B. Lothenbach, “Influence “Structural Behavior of Alkali Activated Fly Ash Concrete. Part 2: Structural
of Activator Type on Hydration Kinetics, Hydrate Assemblage and Testing and Experimental Findings,” Mater. Struct., 46 [3] 449–62 (2013). h

Dr. Susan A. Bernal is a Research Prof. John L. Provis holds BE


Fellow in the Department of Materi- (Hons.) (2002) and PhD (2006)
als Science and Engineering, Univer- degrees in Chemical Engineering
sity of Sheffield, UK. She holds from the University of Melbourne,
BEng (2004) and DEng (2009) Australia, and a BSc (2002) in
degrees in Materials Engineering Applied Mathematics from the same
from Universidad del Valle, Colom- institution. He was appointed to a
bia. From 2009 to 2010 she was a Chair in the Department of Materi-
postdoctoral researcher in the als Science and Engineering at the
Instrument Center for Solid-State University of Sheffield in 2012, hav-
NMR Spectroscopy at Aarhus Uni- ing previously led the Geopolymer
versity, Denmark, and then worked and Minerals Processing Group in
as a Research Fellow at the Univer- the Department of Chemical & Bio-
sity of Melbourne, Australia from molecular Engineering, University of
2010-2012, before relocating to Sheffield. Her research inter- Melbourne. He is Associate Editor of Cement and Concrete
ests involve the development and assessment of low-CO2 Research, and has edited a technical book and a RILEM
cementitious materials, especially alkali-activated binders, state of the art report on geopolymers and other alkali-acti-
particularly focusing on valorization of wastes derived from vated binder systems. He is the 2013 holder of the RILEM
industrial and mining processes, characterization of cements Robert L\x92Hermite Medal, awarded annually to the lead-
on the nano- and macro-scale, and durability. ing researcher under the age of 40 active in the areas of con-
struction materials, systems or structures, and chairs a
RILEM Technical Committee (TC 247-DTA) focused on
durability analysis of non-Portland cement concretes. His
research is focused on the characterization of novel low-CO2
cements and concretes, and the development of conceptual
and computational models for the fluid-solid interaction pro-
cesses which control the formation and durability of cementi-
tious materials in infrastructure and nuclear waste
immobilization applications.

Vous aimerez peut-être aussi