You are on page 1of 12

Colloids and Surfaces A: Physicochem. Eng.

Aspects 269 (2005) 47–58

Understanding the relationship between geopolymer composition,


microstructure and mechanical properties
Peter Duxson a , John L. Provis a , Grant C. Lukey a , Seth W. Mallicoat b ,
Waltraud M. Kriven b , Jannie S.J. van Deventer a,∗
aDepartment of Chemical and Biomolecular Engineering, The University of Melbourne,
Vic. 3010, Australia
b Department of Material Science and Engineering, The University of Illinois at Urbana-Champaign,
Urbana 61801, IL, USA

Received 24 March 2005; received in revised form 15 June 2005; accepted 28 June 2005
Available online 18 August 2005

Abstract

A mechanistic model accounting for reduced structural reorganization and densification in the microstructure of geopolymer gels with high
concentrations of soluble silicon in the activating solution has been proposed. The mechanical strength and Young’s modulus of geopolymers
synthesized by the alkali activation of metakaolin with Si/Al ratio between 1.15 and 2.15 are correlated with their respective microstructures
through SEM analysis. The microstructure of specimens is observed to be highly porous for Si/Al ratios ≤1.40 but largely homogeneous for
Si/Al ≥1.65, and mechanistic arguments explaining the change in microstructure based on speciation of the alkali silicate activating solutions
are presented. All specimens with a homogeneous microstructure exhibit an almost identical Young’s modulus, suggesting that the Young’s
modulus of geopolymers is determined largely by the microstructure rather than simply through compositional effects as has been previously
assumed. The strength of geopolymers is maximized at Si/Al = 1.90. Specimens with higher Si/Al ratio exhibit reduced strength, contrary to
predictions based on compositional arguments alone. The decrease in strength with higher silica content has been linked to the amount of
unreacted material in the specimens, which act as defect sites. This work demonstrates that the microstructures of geopolymers can be tailored
for specific applications.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Geopolymer; Young’s modulus; Microstructure

1. Introduction has been published using different terminology including


‘low-temperature aluminosilicate glass’ [3], ‘alkali-activated
The term geopolymer was first applied by Davidovits [1] cement’ [4] and ‘hydroceramic’ [5], the term ‘geopolymer’
to alkali aluminosilicate binders formed by the alkali sili- is the generally accepted name for this technology. The back-
cate activation of aluminosilicate materials. Geopolymers are bone matrix of geopolymers is an X-ray amorphous analogue
often confused with alkali-activated cements, which were of the tetrahedral alkali aluminosilicate framework of zeo-
originally developed by Glukhovsky in the Ukraine dur- lites. Due to their inorganic framework, geopolymers are
ing the 1950s [2]. Glukhovsky worked predominantly with intrinsically fire resistant and have been shown to have excel-
alkali-activated slags containing large amounts of calcium, lent thermal stability far in excess of traditional cements
whereas Davidovits pioneered the use of calcium-free sys- [6]. Geopolymers have also been shown to exhibit superior
tems based on calcined clays. Although research in this field mechanical properties to those of Ordinary Portland Cement
(OPC) [7–9]. However, compared to other technologies there
∗ Corresponding author. Tel.: +61 3 8344 6619; fax: +61 3 8344 7707. is not yet a substantial body of research focused on under-
E-mail address: jannie@unimelb.edu.au (J.S.J. van Deventer). standing the relationships between composition, processing,

0927-7757/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2005.06.060
48 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

microstructure and the properties (e.g. mechanical strength) to affect significantly the rate and extent of dissolution dur-
of geopolymers. ing geopolymerization [19].
The majority of published studies on geopolymer systems The link between composition and strength has been inves-
have focused on composite fly ash/blast furnace slag systems. tigated previously for sodium silicate/metakaolin geopoly-
In most cases the analysis has been limited to observation of mers, and while it was hinted that there was a link between
X-ray diffractograms and the ultimate compressive strength, mechanical strength, composition and microstructure, none
which are standard techniques in cement science. Mean- was elucidated [20]. A geopolymer composition with opti-
while, microstructural detail has been less intensely inves- mized mechanical strength was identified to occur at an
tigated, due largely to the complexities involved in analysis intermediate Si/Al ratio. However it would be expected that
of materials formed from such highly inhomogeneous alumi- the strength of fully condensed tetrahedral aluminosilicate
nosilicate sources. The use of metakaolin (calcined kaolinite network structures should increase monotonically with sil-
clay) as an aluminosilicate source eliminates many of these ica content, due to the increased strength of Si O Si bonds
issues by providing a purer, more readily characterized start- in comparison to Si O Al and Al O Al bonds [21]. There-
ing material, thereby greatly enhancing the microstructural fore, the relationship between Si/Al ratio and the mechanical,
understanding that may be obtained by analysis of the final physical and microstructural properties of geopolymers needs
reaction products. Metakaolin-based geopolymers are a con- to be determined, with reference to a new mechanistic under-
venient ‘model system’ upon which analysis can be carried standing of geopolymerization.
out, without the unnecessary complexities introduced by the The most critical element of geopolymerization that has
use of fly ash or slag as raw materials. been explored only briefly is the transformation from liquid
The effect of different calcium containing raw-materials precursor to “solid” gel and the mechanisms of densification
[10,11], other ionic additives [12], curing conditions [13] [15]. This provides the key to controlling the nanostruc-
and post-curing chemical treatments [14] on compressive ture, porosity and properties of geopolymers so they may
and/or flexural strength have been investigated in some depth. be tailored for specific applications. Gelation results from
However, few other relevant mechanical properties, in par- hydrolysis–polycondensation of aluminum and silicon con-
ticular density and Young’s modulus, have been measured taining species, resulting in a complex network swollen by
in these studies. These properties are highly significant in water trapped in the pores. Aluminosilicate gels formed by
architectural and structural applications, as well as being a the sol–gel process are made of primary globular polymeric
valuable tool by which the relationship between structure entities 0.8–2.0 nm in diameter, which are densely packed
and properties may be understood. The general aim of initial according to the hydrolysis–polycondensation rate and the
investigations was to demonstrate the utility of geopoly- water content [22]. Structural reorganization of the network
mers in a broader context, but with limited analysis of the occurs by continued reaction and expulsion of the water into
underlying mechanisms. As such these investigations have larger pores. The effect of the main compositional parame-
proven valuable, but lack a systematic approach to deter- ter of geopolymers, the Si/Al ratio, on the gel transformation
mining the effects of basic compositional variables and pro- densification process and how this affects the physical prop-
cessing conditions on intrinsic geopolymer properties and erties of geopolymers has not been explored.
microstructure. The compositions of geopolymers in the current work
Initial studies of geopolymer microstructure focused on have been formulated to ensure that the Al/Na ratio is con-
identification of unreacted particles and determining the stant at unity, providing sufficient alkali to enable complete
chemical composition of the binder in systems synthesized charge balancing of the negatively charged tetrahedral alu-
from multi-component materials, such as blast furnace slag minium centres, while maintaining a constant H2 O/Na2 O
and fly-ash [11,15,16]. Geopolymers have been shown to ratio of 11. The composition of the geopolymers studied is
have a microporous framework, with the characteristic pore therefore controlled by varying the composition of the activat-
size being determined by the nature of the alkali cation or mix- ing solutions by addition of soluble silicate. The differences
ture of cations used in activation [17]. Studies of fly ash-based in microstructure between geopolymers of different compo-
geopolymeric systems identified quartz and mullite particles sition are able to be characterized by SEM and therefore
that act as micro-aggregates in the final matrix, with evidence correlated with basic macro-scale physical properties: ulti-
of unreacted glassy aluminosilicates. It is therefore thought mate compressive strength, Young’s modulus and superficial
that the glassy material acts as the source of aluminum and density. The relationship between composition and properties
silicon for the gel in these systems. Fracture surface analy- is to be explored by firstly confirming the trends in mechani-
sis of clay-based systems shows sheets of unreacted particles cal strength observed by Rowles and O’Connor [20] and then
lodged in the gel [16]. The presence of potentially reactive linking these results to the microstructure of the specimens.
aluminosilicate particles in hardened geopolymer indicates Furthermore, through investigation of the activating solution
that hardening is completed prior to complete dissolution by 29 Si NMR and interpretation of the resulting microstruc-
of raw materials [16,18]. As would be expected from sim- tures in terms of gel transformation, a greater understanding
ple mass transport considerations, the initial particle size of the mechanistic processes occurring during the latter stages
and/or specific surface area of metakaolin has been shown of geopolymerization can be achieved.
P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58 49

2. Experimental methods different coating thicknesses, a different coating medium


(osmium) and left uncoated (analyzed in a FEG-ESEM with
2.1. Materials 2 Torr pressure) to ensure microstructural detail was not
altered by sample coating. The sample coating routine finally
Metakaolin was purchased under the brand name of selected was found to accurately display the microstructure of
Metastar 402 from Imerys Minerals, UK. The metakaolin the geopolymer without affecting any details or introducing
contains a small amount of a high temperature form of mus- artefacts in the coating process. The TEM specimen was pre-
covite (PDF 46,0741) as impurity. The chemical composi- pared by Focussed Ion Beam (FIB) milling of a thin section
tion of metakaolin determined by X-ray fluorescence (XRF) using a Focused Ion Beam ×P200 (FEI Company, Hillsboro,
was 2.3·SiO2 ·Al2 O3 . The Brunauer–Emmett–Teller (BET) OR, USA). The specimen was analyzed by bright field (BF)
surface area [23] of the metakaolin, as determined by nitro- imaging.
gen adsorption on a Micromeritics ASAP2000 instrument, is
12.7 m2 /g, and the mean particle size (d50) is 1.58 ␮m. An 2.4. Compressive strength and density
XRD diffractogram of this material is available elsewhere
[24]. Ultimate compressive strength and Young’s modulus were
Sodium silicate solutions with composition SiO2 /Na2 O = determined using an Instron Universal Testing Machine mov-
R (0.0, 0.5, 1.0, 1.5 and 2.0) and H2 O/Na2 O = 11 were pre- ing at a constant cross-head displacement of 0.60 mm/min.
pared by dissolving amorphous silica (Cabosil M5, 99.8% Specimens were cylindrical, 25 mm in diameter and 50 mm
SiO2 ) in sodium hydroxide solutions of the required con- high to maintain a 2:1 aspect ratio. Sample surfaces were
centration until clear. Solutions were stored for a minimum polished flat and parallel to avoid the requirement for cap-
of 24 h prior to use to allow equilibration. Sodium hydrox- ping. All values presented in the current work are an average
ide solutions were prepared by dissolution of NaOH pel- of six samples with error reported as average deviation from
lets (Merck, 99.5%) in Milli-Q water, with all containers mean. Nominal sample density was measured by averaging
kept sealed wherever possible to minimize contamination by calculated density given by the weight of each of the six sam-
atmospheric carbonation. ples divided by their volume prior to compressive strength
testing.
2.2. Geopolymer synthesis

Geopolymer samples were prepared by mechanically mix- 2.5. NMR spectroscopy


ing stoichiometric amounts of metakaolin and alkaline sili- 29 Si
cate solution to give Al2 O3 /Na2 O = 1, forming a homogenous NMR spectra were obtained at a Larmor frequency
slurry. After 15 min of mechanical mixing the slurry was of 119.147 MHz with a Varian (Palo Alto, CA) Inova 600
vibrated for a further 15 min to remove entrained air before NMR spectrometer (14.1 T). Spectra were collected using
being transferred to Teflon moulds and totally sealed from a 10 mm Doty (Columbia, SC) broadband probe. Between
the atmosphere. Samples were cured in a laboratory oven 128 and 256 transients were acquired using a single 70◦
at 40 ◦ C and ambient pressure for 20 h before being trans- pulse of about 8 ␮s and recycle delays of typically 20 s to
ferred from moulds into sealed storage vessels. The samples ensure full relaxation of all species. The pulse sequence
were then maintained at ambient temperature and pressure described results in NMR spectra that are quantitative
until used in mechanical strength experiments. Specimens with respect to the concentration of 29 Si in different envi-
were synthesized with different Si/Al ratios by use of the five ronments. Spectra were referenced to monomeric silicate,
different concentrations of alkali activator solutions, R = 0.0, Si(OH)4 .
0.5, 1.0, 1.5 and 2.0. This resulted in a total of five different
specimen compositions with nominal chemical composition 2.6. Nitrogen adsorption
M (SiO2 )z AlO2 ·5.5 H2 O, where z is 1.15, 1.40, 1.65, 1.90
and 2.15. N2 adsorption/desorption plots of powdered specimens
were carried out with a Micromeritics Tristar 3000 (Nor-
2.3. Electron microscopy cross, GA). The air (water) desorption was performed at
100 ◦ C for typically 24 h. Surface areas were calculated
Electron microscopy was performed using an FEI XL-30 with an accuracy of 10%, from the isotherm data using the
FEG-SEM and a Phillips CM200 (FEI Company, Hillsboro, Brunauer–Emmet–Teller method [23]. Mesopore diameter
OR, USA). Samples were polished using consecutively finer distributions and cumulative pore volumes were determined
media, prior to final preparation using 1 ␮m diamond paste with the Barret–Joyner–Halenda (BJH) method [25] using the
on cloth. As geopolymers are intrinsically non-conductive, desorption data. The total pore volume Vp was derived from
samples were coated using a gold/palladium sputter coater to the amount of vapor adsorbed at a relative pressure close to
ensure that there was no arching or image instability during unity, by assuming that pores filled subsequently with con-
micrograph collection. A control sample was prepared using densed adsorbate in the normal liquid state.
50 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

essentially constant to within experimental uncertainty in the


region Si/Al ≥ 1.65.
SEM micrographs of geopolymers over the composition
range of interest exhibit significant change in microstruc-
ture with variation in Si/Al ratio (Fig. 2). The change in
microstructure appears most dramatic between Si/Al ratios
of 1.40 and 1.65. Specimens with Si/Al ≤1.40 exhibit
a microstructure comprising large interconnected pores,
loosely structured precipitates and unreacted material, corre-
sponding to low mechanical strength and Young’s modulus.
Geopolymers with Si/Al ratio ≥1.65 are categorized by a
largely homogeneous binder containing unreacted particles
and some smaller isolated pores a few microns in size.
The microstructures of geopolymers with Si/Al ratio
Fig. 1. Young’s moduli () and ultimate compressive strengths () of ≥1.65 do not change significantly with increasing Si/Al ratio.
geopolymers. Error bars indicate the average deviation from the mean over
the six samples measured.
However, there is a slight decrease in the observed porosity in
the specimen with Si/Al ratio of 1.90, which correlates with
the observed maxima in compressive strength and Young’s
3. Results and discussion modulus in this specimen (Fig. 2). Therefore, improvement
in microstructural homogeneity provides a strong reasoning
The average compressive strengths and Young’s moduli for the increase in mechanical properties at lower Si/Al ratios,
of the five different compositions of geopolymer studied in but there is nothing directly observable in the SEM micro-
the current work are summarized in Fig. 1. The compressive graphs that can explain what is responsible for the decrease in
strengths determined in the current work confirm the trends strength above the maximum. Theoretically, Si O Si link-
observed in similar previous work [20]. The Young’s modu- ages are stronger than Si O Al and Al O Al bonds [21],
lus of each sample was calculated from the linear stress/strain meaning that the strength of geopolymers should increase
response prior to failure. The observed variation in Young’s with Si/Al ratio since the density of Si O Si bonds increases
modulus for each composition is comparatively smaller than with Si/Al ratio [24]. The decrease in mechanical strength
that observed for the ultimate compressive strength, partic- between specimens with Si/Al ratio of 1.90 and 2.15 suggests
ularly at higher Si/Al ratios where the variation in strength that other factors begin to affect the mechanical properties.
between samples increases (It should be noted that error in However, the similarity in appearance of the microstructures
compressive strength of geopolymers has previously been of geopolymers with Si/Al ≥1.65 correlates well with the
notionally estimated to be ±5%). This suggests that the almost constant Young’s moduli of these specimens. There-
Young’s moduli of geopolymers are a more characteristic fore, it is apparent that the Young’s modulus of geopolymers
measure of the mechanical properties of each composition, is closely linked with the microstructure, whereas one or more
whereas the greater deviation in the measured ultimate com- other parameters must play a role in determining the ultimate
pressive strength suggests that the failure mechanism con- mechanical strength.
tributes significantly to the measured strength. Observed Geopolymers are known to contain amounts of unre-
ultimate compressive strength data should therefore be con- acted solid aluminosilicate source, metakaolin in this case
sidered as a distribution rather than a discrete value. Inves- [9,16,26], which is confirmed by the plate-shaped voids
tigations focused specifically on describing the distribution observed in the SEM micrographs in Fig. 2. These voids
of the ultimate compressive strength of geopolymers are cur- are produced during the polishing process as the soft, plate-
rently being undertaken, using much larger sample popula- like metakaolin particles remaining unreacted are torn from
tions to ensure that the observed distributions are statistically the binder phase. However, there is no definitive and accu-
sound. rate method for quantitatively determining the amount of
The compressive strength of geopolymers is observed unreacted material in a particular specimen. From the micro-
to increase by approximately 400% from Si/Al = 1.15 to graphs in the current work it can be seen that the level of
Si/Al = 1.90 before decreasing again at the highest Si/Al unreacted material varies between specimens, and would
ratio of 2.15. The improvement in mechanical strength is therefore be expected to have correspondingly varying effects
essentially linear over the region 1.15 ≤ Si/Al ≤ 1.90. How- on their mechanical properties. Metakaolin is weak and will
ever, the same trend is not observed in the Young’s mod- be expected to act as a point defect in the structure, locally
uli, where the Si/Al = 1.90 specimen displays only a minor intensifying the stress in the binder and precipitating failure.
increase above Si/Al = 1.65. This suggests that the improve- Therefore, for any qualitative or semi-quantitative descrip-
ment in mechanical strength and Young’s modulus in the tion of the effect of unreacted material on the strength of
region 1.15 ≤ Si/Al ≤ 1.90 may be related, but not intrinsi- geopolymers, a measure of the amount of unreacted material
cally linked. Indeed, the Young’s modulus may be said to be is required.
P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58 51

Fig. 2. SEM micrographs of Na-geopolymers: Si/Al ratio of (a) 1.15, (b) 1.40, (c) 1.65, (d) 1.90 and (e) 2.15.

27 Al MAS-NMR has been used to correlate the amount amount of unreacted phase has been observed to be at a
of Al(VI) and the amount of unreacted phase in metakaolin- maximum. Therefore, the reduction in mechanical strength
based geopolymers of compositions studied in the current of geopolymer with high Si/Al ratios can be understood
work [26]. While this method does not provide an unequivo- by incorporating the concept of a defect density resulting
cal quantification of the unreacted content, it is able to detect from unreacted material. It also stands to reason that with an
a trend in the amount of Al(VI) in all specimens studied, increased defect density, the number of potential pathways to
matching theoretical expectations. The amount of unreacted failure similarly increases. This would lead to an increased
material has been observed to increase with Si/Al ratio. It is distribution in the measured compressive strengths of indi-
thought that greater amounts of unreacted material increase vidual specimens, as observed in Fig. 1.
the defect density in the specimens and have a deleterious Pore sizes in the order of <5 ␮m are observed in the
effect on the mechanical strength of geopolymers. This effect micrographs of geopolymers with Si/Al ≥1.65 (Fig. 2). The
is particularly pronounced at high Si/Al ratios, where the binder at the interface of some of these pores can be seen
52 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

to have a layered texture. This apparent layered texture is an hysteresis loop characteristics indicates a change in the distri-
artefact created by particle pullout of the plate-structure in bution of pores within the specimens. 2 H and 1 H MAS-NMR
metakaolin during polishing as opposed to pores filled with have shown that the pore size in geopolymers decreases with
solution. Previous SEM micrographs of fracture surfaces of increasing Si/Al ratio [26].
clay derived geopolymers do not show the same large pores, The change in pore volume distributions of sodium
confirming the effect of polishing on the porosity observed geopolymers is summarized in Fig. 4. The pore volume dis-
in polished cross-sections [16]. The cross-sectional area of tribution of geopolymers can be observed to shift into smaller
the pores caused by particle pullout indicates that the amount pores as the Si/Al ratio increases. However, the pore size dis-
of unreacted material in the samples once cured is signifi- tribution of the specimen with Si/Al ratio of 1.15 is observed
cant. Unreacted particles can be seen to be loosely wedged to be bimodal, which can be explained by the large volume
in the structure of geopolymers with Si/Al <1.65 and do not of interconnected pores in combination with some level of
appear to be tightly adhered to the binder. Due primarily to crystallinity in alkali-activated specimens [24]. The nitrogen
the dramatic changes in microstructure with Si/Al ratio, it adsorption/desorption characteristics of geopolymers (Fig. 3)
is impossible to confirm from SEM micrographs whether confirm the observations in the SEM micrographs (Fig. 2) that
the trend in the amount of unreacted particles in geopoly- the increase in nominal Si/Al ratio results in large changes in
mers supports the theoretical predictions and trends observed the microstructure and pore distribution of geopolymers.
previously [26]. Furthermore, not all of the pores in the speci- The nominal densities of geopolymers with varying Si/Al
mens with Si/Al ≥1.65 appear to result from particle pullout. ratios are also presented in Table 1. The density of geopoly-
Some pores appear to be a result of pooling from regions of mers is seen to increase from 1.683 to 1.798 g/cm3 in the
water that are generated in the polycondensation and transfor- range 1.15 ≤ Si/Al ≤ 2.15. The increase in nominal density
mation step of geopolymerization. The sizes of these pores of geopolymers observed with increasing Si/Al ratio results
range from microns to less than 10 nm in diameter (below from the higher proportion of solid components due to addi-
the resolution of SEM) [27], further complicating attempts to tion of silicon to the activating solution. This provides an
gauge the amount of unreacted phase in metakaolin geopoly- activating solution of higher density, and so mixing a given
mers and provide corroboratory evidence to support previous amount (calculated on a solute-free basis to maintain constant
findings [26]. overall H2 O/Na2 O) of this solution with a particular amount
Nitrogen adsorption/desorption isotherms of the speci- of metakaolin will give a product of higher nominal density.
mens in the current work are shown in Fig. 3. All specimens The large decrease in pore volume of geopolymers with
have a type IV isotherm with a hysteresis loop, though the increasing Si/Al ratio (Table 1) infers that accompanying the
characteristic shape of the isotherms and volume of nitrogen change in pore distribution from large to small pores, the
adsorbed per unit volume of specimen change remarkably increase in Si/Al ratio results in a net increase in the volume
with Si/Al ratio. At Si/Al ratio of 1.15, the volume of nitrogen of gel for only a slight increase in nominal density. Pore
adsorbed initially is large, indicating the high volume of large volume is related to the skeletal density, which represents the
interconnected pores in the specimen as observed in Fig. 2. density of the geopolymer gel, and the nominal density by
At higher Si/Al ratios, the initial volume of nitrogen adsorbed the following relation:
is lower, indicating a characteristic change in pore distribu-
tion and a more reduced volume of freely accessible pores. 1 1
Vp = − (1)
The volume of nitrogen adsorbed decreases as the Si/Al ratio ρgel ρskeleton
increases, which results in a decrease in the pore volume, Vp ,
presented in Table 1. The pore volume is observed to decrease where Vp is the specific pore volume, ρgel equals the bulk
from 0.206 to 0.082 cm3 /g as the Si/Al ratio of the specimens gel density and ρskeleton equals the density of the solid
increases. The hysteresis loop measured between the adsorp- phase which comprises the skeletal framework. This rela-
tion and desorption isotherms is observed to become larger tion assumes that pores that are inaccessible to N2 during
and occurs at lower relative pressures with increasing Si/Al the adsorption/desorption experiment are part of the skeletal
ratio, with the exception of the specimen with Si/Al ratio framework. The calculated skeletal densities of the geopoly-
of 2.15, which has the smallest pore volume. The change in mer gel are shown in Table 1 and presented in Fig. 5. The
skeletal density is observed to decrease with increasing Si/Al
Table 1 ratio, while nominal density increases. The increase in appar-
Cumulative pore volume (Vp ), nominal gel density (ρgel ) and calculated ent gel volume in the polished micrograph cross-sections in
skeletal density (ρskeleton ) of geopolymer specimens Fig. 2 must therefore result from the decreased skeletal gel
Specimen Si/Al Vp ρgel (g/cm3 ) ρskeleton (g/cm3 ) density in these specimens, rather than a greater nominal den-
1.15 0.206 1.683 2.57 sity. This confirms that the porosity within the geopolymer
1.40 0.205 1.695 2.60 monoliths becomes more highly distributed in small pores
1.65 0.187 1.718 2.53 inaccessible to N2 as the Si/Al ratio increases. A decrease in
1.90 0.143 1.777 2.38 skeletal density with increasing Si/Al ratio, while maintain-
2.15 0.082 1.798 2.11
ing a relatively constant nominal density results in a larger
P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58 53

Fig. 3. N2 Isotherms of sodium geopolymers with Si/Al ratios of (a) 1.15, (b) 1.40, (c) 1.65, (d) 1.90 and (e) 2.15.

volume of gel. The larger gel volume leads to a progressively polycondensation in highly siliceous specimens [28], which
more homogenous microstructure as observed in the micro- hinders reorganization and reduces the permeability through
graphs in Fig. 2. The larger gel volume allows stress during aggregation of water in certain regions of the gel. Further-
compression to be spread over a larger area, resulting in less more, the observed differences in microstructure can be seen
strain and higher Young’s modulus. to affect other physical properties of the gel such as adsorp-
The change in pore distribution and localized gel density tion and desorption (Fig. 3), and be likely to also affect ion
must result from differences in the mechanism of geopoly- exchange and chemical encapsulation characteristics. The
merization under conditions of higher concentrations of ability to control microstructural characteristics of geopoly-
soluble silicon in the activating solution. The change in mers will allow future geopolymer formulations to be tailored
mechanism hinders aggregation of pores (syneresis) during on a microstructural and chemical level for specific applica-
polycondensation and hardening, leaving more small pores tions.
distributed around the gel framework, rather than smaller The largest change in the microstructure of geopolymers
numbers of large pores. Hindered syneresis is likely to result in the current work appears to occur between the specimens
from factors such as reduced lability of gel precursors during with Si/Al ratios of 1.40 and 1.65. SEM micrographs of
54 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

Fig. 5. Comparison of () nominal and () skeletal densities of sodium


geopolymers.

in the region between 1.40 and 1.65. Therefore, the fac-


Fig. 4. Pore volume distribution of sodium geopolymers.
tors influenced by the soluble silicon concentration in the
activator that directly affect microstructural evolution during
geopolymers with Si/Al ratios between 1.45 and 1.60 are reaction must be in a critical transition in this concentration
presented in Fig. 6, allowing closer analysis of the change region.
in microstructure observed in Fig. 2. The microstructures Dissolution studies of aluminosilicate materials have
comprised both homogeneous and porous regions of gel. found that the initial rate of aluminum dissolution is higher
The homogeneous regions are seen to comprise a greater than that of silicon, due to the formation of an aluminum
proportion of the cross-section as the Si/Al ratio increases. deficient layer, followed by stoichiometric release of sili-
The transition from the porous microstructure observed in con and aluminum [29,30]. Therefore, it is expected that
geopolymers with Si/Al ≤1.40 to the largely homogeneous the metakaolin used in this experiment will initially release
microstructure where Si/Al ≥1.65 is essentially continuous monomeric silicon and aluminum in the ratio of Si/Al <1

Fig. 6. SEM micrographs of geopolymers with Si/Al = (a) 1.45, (b) 1.50, (c) 1.55 and (d) 1.60.
P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58 55

Fig. 8. Connectivity histogram obtained by integrating 29 Si NMR spectra


of sodium silicate solutions for SiO2 /Na2 O = 0.5, 1.0, 1.5 and 2.0. The error
associated with each bar is ±2%.

that have been identified are available elsewhere [33]. The


regions of the spectra relating to each of the different types
of Q-centers are indicated in Fig. 7. Subscript c indicates that
the sites are present in a three-membered ring, which can be
observed separately from chains or larger rings due to the
deshielding effects of the ring strain in these species. It can
be observed that as the concentration of silicon increases, the
number of larger oligomers increases.
For the purposes of this investigation it is important to
have a quantitative view of the speciation of the sodium sil-
Fig. 7. 29 Si NMR spectra of sodium silicate solutions used in the synthesis
of geopolymer specimens in the current work with SiO2 /Na2 O ratios of (a)
icate solutions at the time of mixing with the metakaolin.
0.5, (b) 1.0, (c) 1.5 and (d) 2.0. The connectivity of the silicate solutions is summarized in
Fig. 8. A large change in speciation can be observed between
followed by a period of approximately equal release of silicon the solutions with SiO2 /Na2 O ratios of 0.5 and 1.0, with
and aluminum. A recent study of the leaching characteris- the amount of monomer decreasing by approximately 50%.
tics of metakaolin under conditions of geopolymerization has These solutions are those used in the synthesis of the spec-
confirmed this expectation [31]. Therefore, the amount of imens with Si/Al ratios of 1.40 and 1.65, respectively. Fur-
silicon available in solution from the alkaline silicate activa- thermore, the majority of all silicon centers are incorporated
tor at the point of initial mixing would be expected to play in non-monomeric species in all solutions except that with
a defining role in determining the speciation of aluminum SiO2 /Na2 O = 0.5. During reaction and prior to gelation, small
throughout geopolymerization, which has been shown to aluminate and silicate species are released by dissolution of
affect the incorporation of aluminium into the gel [26]. the solid aluminosilicate source, in this case metakaolin. The
The 29 Si NMR spectra of the sodium silicate activating Si/Al ratio of the solution during reaction will, therefore,
solutions used in preparation of each of the specimens ana- depend greatly on two factors: (1) the amount of aluminum
lyzed in the current work are presented in Fig. 7. The solution released prior to equimolar dissolution of silicon and alu-
used in the synthesis of the specimen with Si/Al = 1.15 con- minum from metakaolin, and (2) the initial concentration
tains no soluble silica, and so is not shown. The connectivity of silicon present in the activator solution. For geopolymers
of each silicon center can be described using the nomencla- synthesized using activating solutions with SiO2 /M2 O ≥1,
ture of Engelhardt et al. [32] Each site is designated Q, since the Si/Al ratio in the solution will always be greater than
each atom is coordinated with four oxygen atoms, with the unity, since the concentrations of silicon initially in the solu-
number of linkages with other silicon atoms indicated with tion are large compared to the amount of aluminum initially
a subscript and the degree of deprotonation ignored. There- dissolved. Dissolution increases the concentration of sol-
fore, Q0 denotes the monomer Si(OH)(4−x) Ox x− , Q1 indicates uble silicon and initiates the formation of aluminosilicate
each of the silicon atoms in a dimer and also terminal silicon oligomers identified elsewhere [33]. Therefore, the speciation
atoms on larger oligomers and so on. Full descriptions of the within the solutions will tend to become more polymerized
designation of the more than 20 different silicate oligomers as dissolution proceeds, and specimens activated with more
56 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

concentrated solutions will always be more polymerized than


less concentrated solutions. Oligomers link together to form
clusters, which is called gelation. The clusters then continue
to reorganize and react as the geopolymer gel develops and
hardens. The rates of the exchange processes occurring in the
solution phase between the species identified in Fig. 7 and the
aluminosilicate species thus formed [34] will therefore play
a major role in determining the structure and conformation
of the gel.
Silicon is several orders of magnitude less labile than
aluminum in solution at room temperature due to the total pro-
tonation of aluminum at high pH, which catalyzes exchange
processes [34]. Furthermore, once aluminum is incorporated
in stable cyclic species, its lability is greatly reduced [34].
Therefore, it has been found that in aluminosilicate solutions
where the Si/Al ratio is greater than 5, all aluminum is incor-
porated in stable cyclic and larger aluminosilicate species
[35]. In solutions where the Si/Al ratio is smaller, the bulk of
all aluminum is present as monomeric Al(OH)4 − [35]. Fur-
thermore, the reduced silicon concentration in these solutions
Fig. 9. BF TEM image of geopolymer with Si/Al ratio of 2.15.
leads to a less polymerized distribution of silicon species as
observed in the sodium silicate solution with SiO2 /Na2 O of
0.5 in Fig. 7. Hence, the solution phase of geopolymers with sibility of chemically bound water in the form of silanol or
solutions having a low SiO2 /Na2 O ratio in the initial activat- aluminol groups. The transition from a solution with suffi-
ing solution is expected to comprise large amounts of small cient lability to reorganize and densify can be observed to
labile species such as silicate and aluminate monomer and occur in the region from 0.5 < SiO2 /Na2 O < 1.0, where the
aluminosilicate dimer. In specimens with higher SiO2 /Na2 O amount of small silicate species decreases rapidly in favour
concentrations in the initial activating solution, the major- of larger silicate oligomers (Fig. 8). Furthermore, the reduced
ity of the aluminum liberated from dissolution is expected lability of the gel with increasing Si/Al ratio will tend to
to be incorporated in stable aluminosilicate species with the reduce the rate of dissolution of metakaolin and promote
remaining silicon to be consumed by large stable silicate lower conversion rates as expected and previously observed
oligomers. Hence, the lability of geopolymeric gel synthe- [26].
sized with low SiO2 /Na2 O ratios in the initial activating A TEM micrograph of a section of geopolymer with
solution will be much greater than that of specimens with Si/Al = 2.15 is presented in Fig. 9. This microstructure has
higher SiO2 /Na2 O ratios. been reported before [9]. Small clusters of aluminosilicate gel
The lability of the solution phase is critical in deter- can be seen to be dispersed within a highly porous network,
mining the microstructure of geopolymers. After gelation, confirming the expected morphology of the gel structure.
transformation occurs due to continued reaction or structural The sizes of these clusters are on an average approximately
reorganization, which causes the expulsion of fluid from the 5–10 nm. Although geopolymers are often termed as ‘amor-
interstices of the structure into the bulk. This process, synere- phous’, the small size of the clusters in Fig. 9 would result
sis, can result in the break-up of the gel into discrete regions in severe line broadening of peaks in X-ray diffraction even
of less porous gel [36], such as that observed in Fig. 2. Lower if crystalline phases were present within them. The concep-
SiO2 /Na2 O ratios have been shown to promote syneresis in tual framework of nanocrystal formation in geopolymers is
aluminosilicate grouts [36]. Therefore, the smaller and more dealt with elsewhere in detail [37], but the structure of the
labile species present in the solution phase and gel structure geopolymeric gel observed in Fig. 9 supports the contention.
of geopolymer with lower SiO2 /Na2 O ratios in the activating The structural ordering of these gel clusters, their intercon-
solution allow a greater degree of structural reorganization nectivity, their physical, thermal and chemical properties, and
and densification of the gel prior to hardening. In specimens their morphological changes over time will play a crucial
with higher soluble silicate concentrations, the reorganization role in understanding geopolymer science and its application-
of the gel structure is hindered by the slow rate of exchange specific formulation.
between the cyclic or cage-like oligomeric species present.
Hence, hardening will occur when the gel has only formed
small and perhaps not fully condensed and cross-linked clus- 4. Conclusions
ters. This means that the porosity appears uniformly dis-
tributed throughout the microstructure on a length scale that A new mechanistic model for the gel transforma-
is below observation using SEM, and also suggests the pos- tion process occurring during geopolymerization has been
P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58 57

proposed, accounting for changes observed in the microstruc- Acknowledgments


ture and mechanical properties of geopolymer specimens
formed by sodium silicate activation of metakaolin. This Financial support is gratefully acknowledged from the
demonstrates that the characteristics of geopolymers can Australian Research Council (ARC), the Particulate Fluids
be tailored for applications with requirements for spe- Processing Centre (PFPC), a Special Research Centre of the
cific microstructural, chemical, mechanical and thermal Australian Research Council, and the United States Air Force
properties. Office of Scientific Research (AFOSR), under STTR Grant
Specimens with Si/Al ratio ≤1.40 exhibit a microstruc- number F49620-02 C-010 in association with The Univer-
ture comprising clustered dense particulates with large sity of Illinois at Urbana-Champaign and Siloxo Pty Ltd.,
interconnected pores observed by SEM. Specimens with Melbourne, Australia.
Si/Al ≥1.65 appear homogeneous with porosity distributed
in small pores largely below observation in SEM micro-
graphs. Closer inspection of the microstructure of geopoly- References
mers with 1.40 ≤ Si/Al ≤ 1.65 revealed that the evolution of
the microstructure with increasing silicon content is rapid [1] J. Davidovits, J. Therm. Anal. 37 (8) (1991) 1633.
yet continuous within the small compositional region. The [2] V.D. Glukhovsky, Ancient, modern and future concretes, in: P.V.
change in microstructure has been shown by nitrogen adsorp- Krivenko (Ed.), Proceedings of the First International Conference
tion to be a result of an increased volume of gel in these on Alkaline Cements and Concretes, VIPOL Stock Company, Kiev,
1994, pp. 1–9.
specimens, as the skeletal density of the gel decreases. Inspec- [3] H. Rahier, B. Van Mele, M. Biesemans, J. Wastiels, X. Wu, J. Mater.
tion of the higher Si/Al ratio specimens using TEM revealed Sci. 31 (1) (1996) 71.
that the microstructure of the gel comprised clusters of gel in [4] A. Palomo, J.I.L. de la Fuente, Cem. Conc. Res. 33 (2) (2003) 281.
the order of 5–10 nm, interspersed by a highly distributed pore [5] D.D. Siemer, Mater. Res. Innov. 6 (2002) 96.
structure. The change in microstructure is a result of varia- [6] V.F.F. Barbosa, K.J.D. MacKenzie, Mater. Res. Bull. 38 (2) (2003)
319.
tion in the lability of silicate species within the sodium silicate [7] A. Palomo, F.P. Glasser, Br. Ceram. Trans. J. 91 (4) (1992) 107.
activating solutions that control the rate of structural reorga- [8] H. Xu, J.S.J. van Deventer, Int. J. Miner. Proc. 59 (3) (2000) 247.
nization and densification during geopolymerization. Greater [9] W.M. Kriven, J.L. Bell, M. Gordon, Ceram. Trans. 153 (2003)
lability allows extensive gel reorganization and densification, 227.
and facilitates pores to aggregate resulting in a microstruc- [10] H. Xu, J.S.J. van Deventer, Ind. Eng. Chem. Res. 42 (8) (2003)
1698.
ture comprising dense gel particles and large interconnecting [11] C.K. Yip, J.S.J. van Deventer, J. Mater. Sci. 38 (18) (2003) 3851.
pores, whereas reduced lability promotes a decreased local- [12] W.K.W. Lee, J.S.J. van Deventer, Ind. Eng. Chem. Res. 41 (18)
ized gel density and distributed porosity. Lability of the gel (2002) 4550.
during geopolymerization has been linked to the concen- [13] J.G.S. van Jaarsveld, J.S.J. van Deventer, G.C. Lukey, Chem. Eng.
tration of soluble silicon in the sodium silicate activating J. 89 (1–3) (2002) 63.
[14] A. Palomo, M.T. Blanco-Varela, M.L. Granizo, F. Puertas, T.
solution, with higher lability promoted by low silicon con- Vazquez, M.W. Grutzeck, Cem. Conc. Res. 29 (7) (1999) 997.
centration. [15] W.K.W. Lee, J.S.J. van Deventer, Colloid Surf. A 211 (1) (2002) 49.
The increase in gel volume allows for a greater cross- [16] H. Xu, J.S.J. van Deventer, Cem. Conc. Res. 32 (11) (2002) 1705.
section of gel to support compression loads, explaining [17] W.M. Kriven, J.L. Bell, Ceram. Eng. Sci. Proc. 25 (3–4) (2004) 99.
the increase of Young’s modulus until the microstructures [18] C.K. Yip, G.C. Lukey, J.S.J. van Deventer, Proceedings of the 28th
Conference on Our World in Concrete and Structures, Singapore,
become largely homogenous at Si/Al ratio of 1.65. There- 2003.
fore, an increase in Young’s modulus is thought to be mainly [19] H. Rahier, J.F. Denayer, B. van Mele, J. Mater. Sci. 38 (14) (2003)
a product of increased homogeneity of the microstructure 3131.
and not simply improvement in the strength of the actual [20] M. Rowles, B. O’Connor, J. Mater. Chem. 13 (5) (2003) 1161.
binder. There is a rapid increase in the compressive strength [21] B.H.W.S. de Jong, G.E. Brown, Geochim. Cosmochim. Acta 44 (3)
(1980) 491.
of geopolymers with increasing Si/Al ratio. However, spec- [22] P. Colomban, J. Mater. Res. 13 (4) (1998) 803.
imens with Si/Al = 2.15, the highest ratio achievable with [23] S. Brunauer, P.H. Emmett, E. Teller, J. Am. Chem. Soc. 60 (1938)
the synthesis technique used in this investigation, exhibit a 309.
reduced strength compared to those with Si/Al = 1.90. The [24] P. Duxson, S. Mallicoat, G.C. Lukey, W.M. Kriven, J.S.J. van Deven-
reduction in ultimate compressive strength of the highest ter, Ceram. Trans. 165 (2005) 71.
[25] E.P. Barrett, L.G. Joyner, P.P. Halenda, J. Am. Chem. Soc. 73 (1)
Si/Al ratio geopolymer is believed to be a result of the effects (1951) 373.
of unreacted material, which is very soft and acts as a defect [26] P. Duxson, G.C. Lukey, F. Separovic, J.S.J. van Deventer, Ind. Eng.
in the binder phase. Similar effects are not observed in the Chem. Res. 44 (4) (2005) 832.
Young’s moduli of the specimens due to the different struc- [27] J.L. Bell, W.M. Kriven, 62nd Annual Meeting of the Microscopy
tural parameters controlling each of these properties. There- Society of America, Microscopy Society of America, Savannah, vol.
10, 2004, p. 590.
fore, to achieve a geopolymer with high strength and high [28] G.W. Scherer, Cem. Conc. Res. 29 (8) (1999) 1149.
Young’s modulus, an intrinsically porous gel microstructure [29] E.H. Oelkers, S.R. Gislason, Geochim. Cosmochim. Acta 65 (21)
is required. (2001) 3671.
58 P. Duxson et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 269 (2005) 47–58

[30] E.H. Oelkers, J. Schott, J.L. Devidal, Geochim. Cosmochim. Acta [34] M.R. North, T.W. Swaddle, Inorg. Chem. 39 (12) (2000)
58 (9) (1994) 2011. 2661.
[31] D. Feng, H. Tan, J.S.J. van Deventer, J. Mater. Sci. 39 (2) (2004) [35] N. Azizi, R.K. Harris, A. Samadi-Maybodi, Magn. Reson. Chem. 40
571. (10) (2002) 635.
[32] G. Engelhardt, D. Zeigan, H. Jancke, D. Hoebbel, W. Wieker, Z. [36] S.A. Jefferis, A. Sheikh Bahai, Geotechnique 45 (1) (1995)
Anorg. Allg. Chem. 418 (1975) 17. 131.
[33] T.W. Swaddle, J. Salerno, P.A. Tregloan, Chem. Soc. Rev. 23 (5) [37] J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Chem. Mater. 17 (12)
(1994) 319. (2005) 3075.

You might also like