Correlating Mechanical and Thermal Properties of Sodium Silicate-Fly

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Colloids and Surfaces A: Physicochem. Eng.

Aspects 336 (2009) 57–63

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Correlating mechanical and thermal properties of sodium silicate-fly


ash geopolymers
John L. Provis ∗ , Chu Zheng Yong, Peter Duxson, Jannie S.J. van Deventer
Department of Chemical and Biomolecular Engineering, The University of Melbourne, Victoria 3010, Australia

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

Article history: The correlation between mechanical and dilatometric properties of aluminosilicate geopolymer binders
Received 22 September 2008 is highlighted by analysis of a set of samples synthesised from a single ash source using different acti-
Received in revised form 6 November 2008 vating solution compositions and liquid/solid ratios. The geopolymers which display the best strength
Accepted 9 November 2008
performance also show a small expansion in the temperature range 700–800 ◦ C, which is identified as cor-
Available online 21 November 2008
responding to the swelling of a high-silica phase present as pockets within the geopolymeric gel structure.
Systems in which this phase is absent (made using hydroxide activating solutions or very low liquid/solid
Keywords:
ratios) generally show a low extent of binder formation and do not achieve high strength, while systems
Geopolymer
Aluminosilicate
in which the expansive phase dominates have a sub-optimally structured geopolymer phase and also
Fly ash correspondingly reduced strength.
Dilatometry © 2008 Elsevier B.V. All rights reserved.
Strength

1. Introduction aluminosilicate gel chemistry obtained by the study of metakaolin


geopolymers will be of value in this investigation.
Geopolymeric cements are becoming the focus of increasing Geopolymer synthesis involves a sequence of interdependent
research efforts as the need to reduce global CO2 emissions inten- dissolution/reprecipitation reactions, as well as speciation equilib-
sifies. Displaying excellent mechanical strength and resistance ria within the solution phase [11]. These reaction steps are almost
to attack by aggressive environments, these materials represent impossible to isolate for individual study, as the process of isolation
an opportunity to simultaneously improve both environmental of a single step will almost invariably change the reaction conditions
and engineering performance compared to traditional technolo- so much that the step of interest either does not occur or plays a
gies [1,2]. Geopolymers are synthesised by alkaline and/or silicate very different role. It is therefore necessary to take a somewhat less
activation of a solid aluminosilicate source, resulting in a highly direct approach to the study of specific aspects of geopolymerisa-
cross-linked, largely X-ray amorphous gel binder phase bearing tion, and this is where the use of model systems becomes important.
significant similarities to zeolite precursor gels [3]. Geopolymer Recent work has provided critical information with regard to the
synthesis from metakaolin was the focus of much early study [4], role played by dissolved silicate speciation in determining geopoly-
however the possibilities of using coal fly ash as an aluminosilicate mer gel microstructure and kinetics [12–14], and to the respective
source began to be explored in detail over a decade ago [5,6]. Fly ash roles of different alkali cations in geopolymer gel formation [15–17].
geopolymerisation has since been observed to provide probably the Detailed kinetic studies of both fly ash and metakaolin geopoly-
greatest opportunities for commercial utilisation of this technology merisation have shown similar multistep reaction processes, and in
due to the plentiful worldwide raw material supply derived from particular the sequential transformation of an initial gel phase into
coal-fired electricity generation. However, the chemistry of fly ash is the final, highly cross-linked gel [18–20]. One of the aims of this
complex [7,8], and so metakaolin geopolymerisation is widely stud- paper is therefore to transfer as much as possible of the knowledge
ied as a ‘model system,’ whereby the relatively simpler chemistry obtained by systematic studies of metakaolin geopolymerisation to
of the purer system may be used to understand certain aspects of the more industrially relevant fly ash-based systems. Many workers
fly ash geopolymerisation [9]. While there are some significant dif- have previously published sets of mechanical strength data for fly
ferences in the products obtained from these two different classes ash geopolymers along with some microstructural analysis [21–26].
of precursor [10], the ability to use some of the understanding of However, the coupling of mechanical and dilatometric performance
to provide an understanding of binder structure has not previously
been presented.
∗ Corresponding author. Tel.: +61 3 8344 8755; fax: +61 3 8344 4153. Dilatometric analysis of metakaolin-based geopolymers has
E-mail address: jprovis@unimelb.edu.au (J.L. Provis). been undertaken previously [27–32]. In general, a well-reacted

0927-7757/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2008.11.019
58 J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63

metakaolin geopolymer shows some initial shrinkage with loss of Table 2


Activator/ash mass ratios (kg activator per kg fly ash) of all samples.
pore water below 300 ◦ C, a period of very gradual shrinkage up to
700 ◦ C associated with dehydroxylation, and then rapid densifica- r m (moles Na2 O/100 g fly ash in r = 0 sample)
tion above this. However, a detailed dilatometric characterisation 0.075 0.100 0.125 0.150 0.175 0.200 0.225
of fly ash geopolymers has not previously been published. Bakharev
0.0 0.195 0.260 0.325 0.390 0.455 0.520 0.626
[33] presented some data showing shrinkage and residual strength 0.5 0.240 0.320 0.400 0.480 0.560 0.641 0.761
of a selection of fly ash geopolymers after heating to 800, 1000 or 1.0 0.285 0.380 0.475 0.571 0.666 0.761 0.896
1200 ◦ C, and Kong et al. [34] discussed the changes in pore struc- 1.5 0.330 0.440 0.551 0.661 0.771 0.881 1.032
ture and residual strength of fly ash geopolymers after heating to 2.0 0.375 0.501 0.626 0.751 0.876 1.001 1.167
800 ◦ C. Kong et al. [35] have also presented a single, low resolu-
tion dilatometric trace for a fly ash geopolymer up to 800 ◦ C, which
shows a sharp shrinkage at ∼200 ◦ C and further shrinkage beyond to the r = 0 sample (determined empirically). This necessitates the
600 ◦ C. Their discussion was focussed on the mismatch between use of slightly increased solution volumes to account for the higher
binder and aggregate behaviour in the geopolymer concretes stud- viscosity of the higher-silica solutions, so that the r = 0.5 sample in
ied, resulting in the increased strength of heated mortar specimens each set had 10% extra solution added, r = 1.0 had 18%, r = 1.5 had
not being reproduced in concretes as the aggregates expanded 24.5% and r = 2.0 had 30%. In the discussions throughout this paper,
while the binder shrank [35]. each set of samples will be referred to by the activator/ash ratio m
The primary aim of this paper is therefore to highlight for the corresponding to the r = 0 sample, but it must be understood that
first time the valuable information that can be obtained by corre- this ratio is only exact for that specific sample within each set.
lating mechanical and dilatometric investigations of geopolymers. The collection of samples tested here spans the entire range in
In terms of determining the strength development of a geopolymer, which a workable, solid geopolymer can be produced from Glad-
there is clearly no substitute for actual strength testing. However, stone fly ash. Ratios below m = 0.075 provide insufficient liquid to
such tests are not always possible where only a small amount of wet the ash, whereas ratios above m = 0.225 do not set. This method
material is available – for example in screening of a large number of of specifying activator/ash ratios provides a more consistent chem-
ash sources, or where a synthetic aluminosilicate precursor is only ical basis for the understanding of fly ash geopolymerisation than
available in small quantities – and so the availability of an alter- would a more traditional mass proportion, because simple mass
native method of analysis may be desirable to provide indicative proportions do not take into account the different solution compo-
measures of strength in such situations. sitions and the need for Na+ to act as a charge-balancing cation in
the aluminosilicate gel structure. For the sake of comparison, the
activator/ash mass ratios of all samples are given in Table 2.
2. Materials and methods
Compressive strength testing was undertaken using mortar
samples, 50 mm cubes, with an ash/sand mass ratio of 0.47. Samples
The work described in this paper utilises fly ash derived from
were cured in moulds sealed into plastic bags with a small amount
a single source, Gladstone Power Station in Queensland, Aus-
of water to maintain a humid atmosphere at 40 ◦ C for 3 days prior
tralia. This is a Class F ash (ASTM C618) with oxide composition,
to strength testing. The results shown are the average of three sam-
determined by XRF, as given in Table 1. Detailed characterisa-
ples, and a mean deviation of <10% was observed where error bars
tion of this fly ash, including identification of the crystalline and
are not shown explicitly. Dilatometry (PerkinElmer Diamond Ther-
non-crystalline phases present, has been presented by Keyte [8].
momechanical Analyser (TMA)) was conducted on hardened paste
The ash contains small quantities of crystalline quartz, mullite,
samples approximately 5 mm in diameter and 10 mm in length,
and iron oxides (hematite/maghemite/magnetite), and is predom-
cured in sealed moulds at 40 ◦ C for 3 days and then stored at
inantly (∼75%) X-ray amorphous. The d50 of Gladstone fly ash is
room temperature for a further 4–7 days. The TMA instrument was
around 10 ␮m, and 5 vol.% of the ash particles are retained on a
operated using a nitrogen purge (i.e. an inert environment in the
45 ␮m sieve.
temperature range of interest) at a flowrate of 200 mL/min and a
Geopolymer samples were prepared by mixing the fly
heating rate of 10 ◦ C/min. X-ray diffractometry (XRD; Philips PW-
ash with sodium silicate activating solutions of composition
1800; Cu K␣ radiation) utilised crushed paste samples.
Na2 O·rSiO2 ·11H2 O, with r = 0, 0.5, 1.0, 1.5 and 2.0. Activator/ash
ratios for each ‘set’ of samples are specified in terms of m, the num-
ber of moles of Na2 O added per 100 g of ash in the sample with r = 0, 3. Results and discussion
and fall in the range 0.075 ≤ m ≤ 0.225. Subsequent samples in each
set, with r = 0.5. . .2.0, are synthesised with comparable workability 3.1. Mechanical strength

Fig. 1 shows a contour plot of compressive strength data for all


Table 1 specimens studied. This plot shows that there is a clear maximum in
Oxide composition of Gladstone fly ash.
mechanical performance in the region 1 ≤ m ≤1.5, 0.125 ≤ r ≤ 0.175.
Oxide Mass % The data for these activator/ash ratios are therefore replotted as
SiO2 46.4
Fig. 2 to enable more detailed analysis.
Al2 O3 28.3 It is of interest to note that the highest strength observed at
Fe2 O3 11.7 each liquid/solid ratio is generated by the use of an activating solu-
TiO2 1.4 tion with r ∼ 1.0–1.5. This corresponds closely to the composition
MnO 0.2
which gives the highest strength in sodium silicate-metakaolin
CaO 5.1
MgO 1.4 geopolymers [12], which was attributed to microstructural details
K2 O 0.6 of the geopolymer gel. In particular, this solution composition gave a
Na2 O 0.3 microstructure that appeared ‘smooth’ and uniform under scanning
P2 O5 1.0
electron microscopy, whereas higher SiO2 /Na2 O ratios gave prod-
SO3 0.3
LOI 3.3
ucts with more obvious unreacted particles, and lower SiO2 /Na2 O
ratios gave a gel microstructure that was clearly particulate and
Total 100.1
porous on a length scale of microns [12]. While there is clearly a
J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63 59

most likely by reducing the reprecipitation of aluminosilicate gel


particles onto the fly ash surfaces [10,37].
Conversely, in the presence of a high-silica activating solution,
there are few silicate monomers present and so the dissolution of
fly ash is relatively slow [10,13]. Coupled with the high degree of
supersaturation with respect to most aluminosilicate phases that
is generated almost as soon as Al starts to dissolve from the ash
into the solution, at least some fraction of the (alumino)silicate
gel phases that form in the higher-Si systems shown in Fig. 2 will
be expected to show a relatively low degree of connectivity when
compared to the generally fully connected structure of the bulk
geopolymer gel. Such less-connected gels have previously been
observed to lead to a rapid expansion upon heating of very high-
silica metakaolin-derived geopolymers [38]. Similar gels, often with
the addition of some alkali metal carbonate (as will be observed in
geopolymers exposed to atmospheric carbonation), are also used
to provide fire resistance to glass laminates due to their significant
Fig. 1. Contour plot of compressive strength data for all samples. swelling upon heating to fire conditions [39]. Expansive behaviour
was therefore expected to be observable by dilatometric analysis
of the fly ash geopolymers investigated here, in the cases where
significant difference between the geopolymerisation behaviour of
part of the geopolymer structure consists of a poorly reacted alkali
metakaolin and fly ash, due primarily to the different nanostruc-
silicate gel. Non-uniformity of the gel binder has also been pro-
tures and particle morphologies of these two materials, the fact
posed as an explanation for some of the unusual behaviour of
that the same activating solution gives the strongest product in
low-silica geopolymer systems seeded with small quantities of
both instances may prove to be of importance. It will be neces-
alumina nanoparticles [40]. This raises a number of interesting
sary to conduct further analysis on a wider range of fly ash samples
fundamental questions regarding the definitions of ‘single-phase’
(rather than the single ash tested here) before such correlations
or ‘multi-phase’ in the context of a gel, in the absence of clear
can be attributed to anything more than coincidence, and such
boundaries as are observed in phase-separated glasses, but such
investigations are currently ongoing.
discussion is largely beyond the scope of this work.
Fig. 2 shows that, as the activator/ash ratio increases, the opti-
mum silicate content of the activating solution decreases. There
is very little difference between the strengths attained at these 3.2. Dilatometric behaviour as a function of solid/liquid ratio
three activator/ash ratios using high-silica activating solutions
(m > 1). However, low-silica (and particularly zero-silica, i.e. NaOH) Fig. 3 shows the change in dilatometric performance for geopoly-
activators gave significantly higher binder strengths at higher acti- mers with r = 1.5, as m is increased from 0.075 to 0.175 (i.e. from
vator/ash ratios. This is most likely a pH effect; the low-silica the lowest ratio that was able to be mixed, to the highest ratio
activating solutions have a much higher pH than those with more that hardened during curing at 40 ◦ C). The r = 1.5 samples in general
SiO2 added. This will mean that the extent of ash dissolution give the highest strengths for each mix ratio m. All samples show
increases with increasing liquid addition in the lower-silica sys- little shrinkage below 100 ◦ C as free water is released (identified
tems, as the rate of gel formation in these systems will be strongly as “Region I” by Duxson et al. [31]), shrinkage from 100 to 250 ◦ C
controlled by the kinetics of reactions involving silicate monomers, as pore water is released from the gel (Region II), very gradual (if
and the low-silicate activating solutions contain a very high con- any) shrinkage in the range 250–600 ◦ C and attributed to the loss
centration of these species [36]. It has also been observed that of bound hydroxyls (Region III), and then, in most cases, densifica-
the presence of silicate monomers accelerates fly ash dissolution, tion attributed to a process resembling viscous sintering (Region IV)
commencing at around 600 ◦ C. Thermogravimetry (data not shown)
supports the identification of these regions, with the vast majority
of the mass loss taking place below 250 ◦ C in all cases, and attributed
to water loss.

Fig. 2. Compressive strength data for sample sets 0.125 ≤ m ≤ 0.175, where m is the
activator/ash mass ratio of the r = 0 sample in each set. Fig. 3. Dilatometric data for samples with r = 1.5 and varying values of m as marked.
60 J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63

Fig. 4. Expansion peak onset and maximum temperatures, and size of expansion peaks, for samples with r = 1.5, as a function of activator/ash ratio.

Fig. 3 also shows that the use of a higher liquid/solid ratio gives in itself dramatically detrimental to mechanical performance. In
geopolymer products which show a sharp and rapid expansion at fact, it appears that the best mechanical performance is observed
700–800 ◦ C, with the onset temperature decreasing and magnitude in the samples which show a slight expansion at above 800 ◦ C—but
of the expansion increasing with liquid/solid ratio. The higher liq- whether there is any causation in this relationship (i.e. whether
uid/solid samples then melt at around 1000 ◦ C. There may also be the expansive gel contributes markedly to mechanical strength)
effects due to glass devitrification in the presence of alkalis above remains an open question which requires further analysis. It
800 ◦ C in some samples, but this is difficult to distinguish clearly appears unlikely that this gel does contribute to strength, but rather
in the samples studied. However, the dimensional stability of the that a system which has a sufficiently high liquid/solid ratio to
lower liquid/solid samples is quite remarkable, with almost neg- enable satisfactory conversion of fly ash to geopolymer gel will
ligible change in length from 200 to 1000 ◦ C. The geopolymer gel result also in the formation of some small amount of this phase.
binder phase is known to be capable of showing very good nanos- In the case of metakaolin geopolymers, which have a much higher
tructural stability up to around 1000 ◦ C [41], and this appears to level of available aluminium from the solid precursor than do the
be reflected in the dilatometric data for the samples with lower fly ash samples tested here, some samples showed a similar expan-
liquid/solid ratio. sion peak at around 800 ◦ C superimposed on a very rapid shrinkage
The important parameters of the 700–800 ◦ C expansion peaks attributed to viscous sintering processes [30,31]. However, this peak
for all samples shown in Fig. 3 are summarised in Fig. 4. The peak did not appear in a consistent or predictable manner in the data for
onset is defined as the point at which expansion commences (i.e. metakaolin geopolymers, in contrast to the clear trends observed
the temperature at which the maximum shrinkage is observed, here.
immediately prior to the expansion peak), and the height of the Whether the relationship between slight expansion at elevated
expansion peak is the difference between the length at the onset and high strength is causative or simply a correlation, the fact that
temperature and the greatest length measured. Data are not plot- it exists suggests that dilatometry may be a useful tool in screening
ted for m = 0.075, as no distinct peak is observed in this sample. The geopolymer mixes and/or ash sources. While it may seem strange to
onset and peak maximum temperatures decrease monotonically, suggest the use of an instrumental analysis technique as an alter-
and the peak height similarly increases, with the addition of more native to simple strength testing for screening purposes, the fact
activating solution to the ash. However, while there is very little dif- that dilatometry can be conducted on much smaller samples (<1 g
ference between the onset or maximum temperatures for m = 0.15 of ash was used per sample tested here, compared to ∼300 g for a
or 0.175, the peak height changes significantly. This indicates that set of three mortar samples for compressive strength testing) may
the structure of the phase which causes this peak is very similar in be an advantage in some instances. In particular, where a sample of
these two samples, but it is present at a higher concentration in the a specific potential aluminosilicate source is to be sent to an exter-
m = 0.175 sample. This further assists in the association of this peak nal laboratory for testing, the ability to obtain results from an easily
with poorly connected silicate gels, possibly derived from pock- transported quantity of a few grams is significantly advantageous.
ets of activating solution which are far enough from any reactive The use of dilatometry as a method for screening of different fly
fly ash particle surface that they do not show a significant extent ash sources will be explored in detail in a forthcoming publication
of Al incorporation. This is not likely to be related to mixing defi- [42], making use of the methodology presented and validated in
ciencies; the samples were homogenised using a high-shear mixer, this paper.
and mixing problems would be expected to be more notable in the
low-liquid samples which have less favourable rheology. Rather, the 3.3. Dilatometric behaviour as a function of activator composition
higher liquid/solid ratio, in conjunction with the high viscosity of
the r = 1.5 silicate solution, will mean that there are regions of acti- Figs. 5–7 present the dilatometric data obtained for the sample
vating solution to which dissolved aluminate is unable to diffuse sets with m = 0.125, 0.15 and 0.175 respectively, as a function of sil-
during the reaction process. ica content r in each case. Further information regarding the likely
It is also notable here that the samples which show the high- causes of the expansion peak and other features in the dilatometric
est strength (m = 0.0125 or 0.015) do show a marked expansion; traces can be obtained by examination and comparison of these
it is clear that the presence of the expansive high-silica gel is not plots and Fig. 4. For instance, while the data sets in Fig. 4 appear
J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63 61

Fig. 5. Dilatometric data for the sample set with activator/ash ratio m = 0.125 and Fig. 7. Dilatometric data for the sample set with activator/ash ratio m = 0.175 and
varying SiO2 /Na2 O ratios r as marked. varying SiO2 /Na2 O ratios r as marked. Note the change in vertical scale compared to
Figs. 4 and 5.

to show a contraction at a temperature roughly corresponding to


the ␣–␤ quartz transition at 573 ◦ C, the r = 0.0 data sets in Figs. 5–7
(which contain the highest mass percentage fly ash, and therefore
the highest remnant quartz content, of all the samples studied)
do not show a corresponding contraction. Cracking of the sample
accompanying this transition is therefore unlikely to contribute to
the observed shrinkage in the samples studied here.
It is also interesting to note that the difference in the magnitude
of the expansion peaks in the high-silica (r ≥ 1.5) samples in the
m = 0.125 and 0.15 sample sets is significantly less than the differ-
ence between the m = 0.15 and 0.175 sets. Adding extra activating
solution to move from m = 0.125 to 0.15 resulted in the formation of
additional geopolymer binder in these samples, meaning that the
added silica did not result in the formation of correspondingly more
Fig. 6. Dilatometric data for the sample set with activator/ash ratio m = 0.15 and of the expansive phase. However, adding more activating solution
varying SiO2 /Na2 O ratios r as marked. to the higher-silica samples led to a decrease in strength (Fig. 2),

Fig. 8. X-ray diffractograms of the sample set with m = 0.175 and r ranging from 0.0 to 2.0. Stick patterns for phases identified are shown below the diffractograms; hematite
peaks overlap with those due to other iron oxides, and the peak resolution was not sufficient to definitively distinguish hydrosodalite from hydroxysodalite, so these phases
are not shown individually.
62 J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63

meaning that the extra solution did not lead to the formation of use of dilatometry provides a method of screening potential pre-
viable geopolymer binder, and instead gave the expansive silicate cursors for geopolymerisation, which may prove more convenient
gel. The same is not true for the lower-silica (r ≤ 1.0) activating than the synthesis of larger samples for mechanical strength testing
solutions; moving from m = 0.15 to 0.175 in these systems gave an in cases where the transport of large quantities of materials proves
increase in strength, and the expansion peak correspondingly did problematic.
not grow markedly.
The X-ray diffractograms presented in Fig. 8 may also pro- Acknowledgments
vide some insight into the factors which control the behaviour
of samples synthesised at higher silica content. The identifica- This work was funded in part by the Australian Research Council
tion of zeolitic phases in the low-silica samples (chabazite-Na and (ARC), including funding through the Discovery Grants program and
hydrosodalite and/or hydroxysodalite in the r = 0.0 sample, and fau- also via the Particulate Fluids Processing Centre, a Special Research
jasite in the r = 0.5 sample) is consistent across the sample sets with Centre of the ARC.
m = 0.125, 0.15 and 0.175; only the m = 0.175 data are shown here.
It should also be noted that there are multiple Powder Diffraction
File (PDF) entries available for each of these three zeolite structures, References
and it was not possible to definitively select a single ‘best’ PDF card
[1] P. Duxson, A. Fernández-Jiménez, J.L. Provis, G.C. Lukey, A. Palomo, J.S.J. van
for each phase due to the broad and low-intensity peaks present in Deventer, Geopolymer technology: the current state of the art, J. Mater. Sci. 42
each case. The phases are therefore identified here by framework (9) (2007) 2917–2933.
type rather than PDF card number. [2] P. Duxson, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, The role of inorganic poly-
mer technology in the development of ‘Green concrete’, Cement Concr. Res. 37
However, the identification of specific zeolite phases in some (12) (2007) 1590–1597.
samples is of interest, as it provides a potential explanation for [3] J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Do geopolymers actually contain
the differences in behaviour observed at ∼600 ◦ C. The chabazite- nanocrystalline zeolites?—a reexamination of existing results, Chem. Mater.
17 (12) (2005) 3075–3085.
Na phase identified in the r = 0.0 sample is known to convert to [4] J. Davidovits, Geopolymers—inorganic polymeric new materials, J. Therm. Anal.
a sodalite framework structure above ∼400 ◦ C [43], with a corre- 37 (8) (1991) 1633–1656.
sponding decrease in open porosity and thus total volume. The [5] J.G.S. van Jaarsveld, J.S.J. van Deventer, L. Lorenzen, Factors affecting the immo-
bilization of metals in geopolymerized flyash, Metall. Mater. Trans. B 29 (1)
hydrosodalite/hydroxysodalite (i.e. “non-basic” or “basic” hydroso-
(1998) 283–291.
dalite) phases become dehydrated at 400–630 ◦ C, with either a [6] J. Wastiels, X. Wu, S. Faignet, G. Patfoort, Mineral polymer based on fly ash, in:
marked expansion or a small contraction depending on the intra- Proceedings of the 9th International Conference on Solid Waste Management,
Philadelphia, PA, 1993, 8 pp.
cage occupancies of the crystal structure [44]. The small expansion
[7] R. Mohapatra, J.R. Rao, Some aspects of characterisation, utilisation and envi-
observed in the r = 0.0 samples in this temperature region may ronmental effects of fly ash, J. Chem. Technol. Biotechnol. 76 (2001) 9–26.
therefore be tentatively assigned to the expansion of the non-basic [8] L.M. Keyte, Ph.D. Thesis, University of Melbourne, Australia, 2008.
hydrosodalite component of the geopolymer. The slight shrinkage [9] J.S.J. van Deventer, J.L. Provis, P. Duxson, G.C. Lukey, Reaction mechanisms in
the geopolymeric conversion of inorganic waste to useful products, J. Hazard.
above 600 ◦ C may then be attributed to restructuring of the gel Mater. A139 (3) (2007) 506–513.
and pore network as observed in metakaolin geopolymers previ- [10] R.R. Lloyd, Ph.D. Thesis, University of Melbourne, Australia, 2008.
ously [31], along with the dehydration of any basic hydrosodalite [11] J.L. Provis, J.S.J. van Deventer, Geopolymerisation kinetics. 2. Reaction kinetic
modelling, Chem. Eng. Sci. 62 (9) (2007) 2318–2329.
(hydroxysodalite) present, as this is known to be stable to higher [12] P. Duxson, J.L. Provis, G.C. Lukey, S.W. Mallicoat, W.M. Kriven, J.S.J. van Deventer,
temperatures than its non-basic counterpart. Understanding the relationship between geopolymer composition, microstruc-
The r = 0.5 samples show the formation of faujasite as the only ture and mechanical properties, Colloids Surf. A 269 (1–3) (2005) 47–58.
[13] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Attenuated total reflectance
significant zeolite phase. The structural collapse of faujasite has pre- Fourier transform infrared analysis of fly ash geopolymer gel ageing, Langmuir
viously been identified in differential thermal analysis (DTA) data 23 (15) (2007) 8170–8179.
for metakaolin-based geopolymers, but did not give a significant [14] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, In situ ATR-FTIR study of
the early stages of fly ash geopolymer gel formation, Langmuir 23 (17) (2007)
peak in dilatometry [31]. These, and the r = 1.0 and 1.5 phases which
9076–9082.
do not show notable zeolitic crystallinity, begin to shrink notably at [15] P. Duxson, G.C. Lukey, F. Separovic, J.S.J. van Deventer, The effect of alkali cations
∼600 ◦ C prior to the onset of the expansion peak at ∼700–800 ◦ C. on aluminum incorporation in geopolymeric gels, Ind. Eng. Chem. Res. 44 (4)
(2005) 832–839.
The r = 2.0 samples do not actually show a clear shrinkage prior
[16] P. Duxson, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, F. Separovic, Z.H. Gan, 39 K
to the onset of expansion, however this may be attributed to the NMR of free potassium in geopolymers, Ind. Eng. Chem. Res. 45 (26) (2006)
early and rapid onset of the expansion of the high-silica phase 9208–9210.
overwhelming the shrinkage of the remainder of the gel phase. [17] P. Duxson, S.W. Mallicoat, G.C. Lukey, W.M. Kriven, J.S.J. van Deventer, The
effect of alkali and Si/Al ratio on the development of mechanical properties
of metakaolin-based geopolymers, Colloids Surf. A 292 (1) (2007) 8–20.
4. Conclusions [18] C.A. Rees, Ph.D. Thesis, University of Melbourne, 2007.
[19] J.L. Provis, J.S.J. van Deventer, Geopolymerisation kinetics. 1. In situ energy dis-
persive X-ray diffractometry, Chem. Eng. Sci. 62 (9) (2007) 2309–2317.
The correlation between mechanical and dilatometric perfor- [20] A. Fernández-Jiménez, A. Palomo, I. Sobrados, J. Sanz, The role played by the
mance of sodium silicate-fly ash geopolymers has been identified, reactive alumina content in the alkaline activation of fly ashes, Micropor. Meso-
and rationalised in terms of the expansion of a high-silica phase por. Mater. 91 (1–3) (2006) 111–119.
[21] A. Palomo, M.W. Grutzeck, M.T. Blanco, Alkali-activated fly ashes—a cement for
which comprises part of the gel binder in samples with high levels of the future, Cement Concr. Res. 29 (8) (1999) 1323–1329.
activating solution addition, or very high-silica activating solutions. [22] J.G.S. van Jaarsveld, J.S.J. van Deventer, Effect of the alkali metal activator on the
Samples which show either a very large or a negligible expansion in properties of fly ash-based geopolymers, Ind. Eng. Chem. Res. 38 (10) (1999)
3932–3941.
the temperature range 700–800 ◦ C do not generally show the high-
[23] Z. Xie, Y. Xi, Hardening mechanisms of an alkaline-activated class F fly ash,
est strengths; optimal mechanical performance is provided by the Cement Concr. Res. 31 (9) (2001) 1245–1249.
mixes which show a small expansion in this temperature range. It [24] H. Rostami, W. Brendley, Alkali ash material: a novel fly ash-based cement,
Environ. Sci. Technol. 37 (15) (2003) 3454–3457.
is believed that this trend is due to a correlation of physical proper-
[25] F. Škvára, T. Jílek, L. Kopecký, Geopolymer materials based on fly ash, Ceram.-
ties rather than being a strictly causative effect—i.e., the expansive Silik. 49 (3) (2005) 195–204.
phase is not directly responsible for high strength development, but [26] M. Steveson, K. Sagoe-Crentsil, Relationships between composition, structure,
remains as pockets in the geopolymer binder structure as a result and strength of inorganic polymers. Part 2. Fly ash-derived inorganic polymers,
J. Mater. Sci. 40 (16) (2005) 4247–4259.
of the high activating solution viscosity and relatively high solution [27] H. Rahier, B. van Mele, J. Wastiels, Low-temperature synthesized aluminosili-
content of the geopolymers which show the highest strength. The cate glasses. 2. Rheological transformations during low-temperature cure and
J.L. Provis et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 336 (2009) 57–63 63

high-temperature properties of a model compound, J. Mater. Sci. 31 (1) (1996) Conference on Pozzolan, Concrete and Geopolymer, Khon Kaen University,
80–85. Khon Kaen, Thailand, 2006, pp. 105–119.
[28] H. Rahier, W. Simons, B. van Mele, M. Biesemans, Low-temperature synthesized [36] J.L. Provis, P. Duxson, G.C. Lukey, F. Separovic, W.M. Kriven, J.S.J. van Deventer,
aluminosilicate glasses. 3. Influence of the composition of the silicate solution Modeling speciation in highly concentrated alkaline silicate solutions, Ind. Eng.
on production, structure and properties, J. Mater. Sci. 32 (9) (1997) 2237–2247. Chem. Res. 44 (23) (2005) 8899–8908.
[29] V.F.F. Barbosa, K.J.D. MacKenzie, Thermal behaviour of inorganic geopolymers [37] W.K.W. Lee, J.S.J. van Deventer, Structural reorganisation of class F fly ash in
and composites derived from sodium polysialate, Mater. Res. Bull. 38 (2) (2003) alkaline silicate solutions, Colloids Surf. A 211 (1) (2002) 49–66.
319–331. [38] R.A. Fletcher, K.J.D. MacKenzie, C.L. Nicholson, S. Shimada, The composi-
[30] P. Duxson, G.C. Lukey, J.S.J. van Deventer, Thermal evolution of metakaolin tion range of aluminosilicate geopolymers, J. Eur. Ceram. Soc. 25 (9) (2005)
geopolymers. Part 1. Physical evolution, J. Non-Cryst. Solids 352 (52–54) (2006) 1471–1477.
5541–5555. [39] K.S. Varma, J.R. Holland, D.W. Holden, Production of fire resistant laminates,
[31] P. Duxson, G.C. Lukey, J.S.J. van Deventer, Characteristics of thermal shrinkage World Patent WO/2002/024445, Pilkington PLC, 2002.
and weight loss in Na-geopolymer derived from metakaolin, J. Mater. Sci. 42 (9) [40] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Geopolymer gel formation
(2007) 3044–3054. with seeded nucleation, Colloids Surf. A 318 (1–3) (2008) 97–105.
[32] A. Subaer, van Riessen, Thermo-mechanical and microstructural characterisa- [41] J.L. Bell, P. Sarin, J.L. Provis, R.P. Haggerty, P.E. Driemeyer, P.J. Chupas, J.S.J.
tion of sodium-poly(sialate-siloxo) (Na-PSS) geopolymers, J. Mater. Sci. 42 (9) van Deventer, W.M. Kriven, Atomic structure of a cesium aluminosilicate
(2007) 3117–3123. geopolymer: a pair distribution function study, Chem. Mater. 20 (14) (2008)
[33] T. Bakharev, Thermal behaviour of geopolymers prepared using class F fly ash 4768–4776.
and elevated temperature curing, Cement Concr. Res. 36 (2006) 1134–1147. [42] J.L. Provis, R.M. Harrex, P. Duxson, J.S.J. van Deventer, Dilatometry of geopoly-
[34] D.L.Y. Kong, J.G. Sanjayan, K. Sagoe-Crentsil, Comparative performance of mers as a means of selecting desirable fly ash sources, in preparation.
geopolymers made with metakaolin and fly ash after exposure to elevated [43] S. Cartlidge, W.M. Meier, Solid state transformations of synthetic CHA-and EAB-
temperatures, Cement Concr. Res. 37 (12) (2007) 1583–1589. type zeolites in the sodium form, Zeolites 4 (3) (1984) 218–225.
[35] D.L.Y. Kong, J.G. Sanjayan, K. Sagoe-Crentsil, The behaviour of geopolymer paste [44] J. Felsche, S. Luger, Phases and thermal decomposition characteristics of hydro-
and concrete at elevated temperatures, in: P. Chindaprasirt (Ed.), International sodalites Na6+x [AlSiO4 ]6 (OH)x ·nH2 O, Thermochim. Acta 118 (1987) 35–55.

You might also like