You are on page 1of 9

Journal of Asian Ceramic Societies 3 (2015) 130138

H OSTED BY

Contents lists available at ScienceDirect

Journal of Asian Ceramic Societies


journal homepage: www.elsevier.com/locate/jascer

Effect of the rate of calcination of kaolin on the properties of


metakaolin-based geopolymers
B.B. Kenne Diffo a,b , A. Elimbi a, , M. Cyr b , J. Dika Manga c , H. Tchakoute Kouamo a
a
b
c

Universit de Yaound 1, Laboratoire de Chimie Inorganique Applique, Facult des Sciences, B.P. 812, Yaound, Cameroon
Universit de Toulouse, Laboratoire Matriaux et Durabilit des Constructions, INSA/UPS Gnie Civil, 31077 Toulouse CEDEX 04, France
Universit de Douala, Laboratoire de Chimie Bio-Inorganique, Analytique et Structurale, Facult des Sciences, B.P. 24157, Douala, Cameroon

a r t i c l e

i n f o

Article history:
Received 7 September 2014
Received in revised form
18 November 2014
Accepted 2 December 2014
Available online 20 December 2014
Keywords:
Rate
Calcination
Residual kaolin
Metakaolin
Setting time
Compressive strength

a b s t r a c t
Kaolin samples of the same mass were treated at 700 C for the same duration of 30 min by varying the rate
of calcination (1, 2.5, 5, 10, 15 and 20 C/min) in order to obtain metakaolins which were used to produce
geopolymers. Depending on the nature of each type of material, kaolin, metakaolins and geopolymers
were characterized using thermal analysis, chemical analysis, XRD, FTIR, particle size distribution, specic
surface area, bulk density, setting time and compressive strength. FTIR and XRD analyses showed that
metakaolins except at 1 C/min contained residual kaolinite whose quantity increased with the rate of calcination of kaolin and which inuenced the characteristics of geopolymers. Thus as the rate of calcination
of kaolin increased, the setting time increased (226 min (rate of 1 C/min)773 min (rate of 20 C/min))
while the compressive strength reduced (49.4 MPa (rate of 1 C/min)20.8 MPa (rate of 20 C/min)). From
the obtained results the production of geopolymers having high compressive strength along with low
setting time requires that the calcination of kaolin be carried out at a low rate.
2014 The Ceramic Society of Japan and the Korean Ceramic Society. Production and hosting by
Elsevier B.V. All rights reserved.

1. Introduction
The term geopolymer refers to inorganic aluminosilicate
polymers. These materials nd applications in areas such as construction and civil engineering [13], encapsulation of certain toxic
waste [47], removal of certain heavy metal cations from aqueous solutions [68] or the development of biomaterials [9]. These
inorganic polymers are generally obtained by polymerization reaction between aluminosilicate and strongly basic alkaline solution
[10]. Aluminosilicate raw materials commonly used for the synthesis of geopolymers are y ash, blast furnace slag, volcanic ash
or metakaolin [1,4,11,12]. The variability of chemical composition of y ash, blast furnace slag or volcanic ash usually leads
to geopolymers whose physical and mechanical properties vary
from one aluminosilicate raw material to another [7,11]. Regarding
metakaolin, the characteristics of resulting inorganic aluminosilicate polymers depend on the conditions of calcination of kaolin

Corresponding author. Tel.: +237 677 612 623; fax: +237 222 234 496.
E-mail address: aelimbi2002@yahoo.fr (A. Elimbi).
Peer review under responsibility of The Ceramic Society of Japan and the Korean
Ceramic Society.

[11,13,14]. Elimbi et al. [11] studied the effect of temperature of


calcination of kaolin on the properties of inorganic aluminosilicate
polymers and found 700 C as the temperature at which geopolymers with optimal characteristics are obtained. Castelein et al. [15]
investigated the effect of heating rate of kaolin to reveal its inuence on the formation of mullite. Elsewhere, Castelein et al. [16]
studied both inuence of temperature and rate of ring of kaolin
on the distribution of iron neighboring kaolinite. Ptcek et al. [17]
studied the effect of the rate of calcination of kaolin to investigate kinetic decomposition of kaolinite under non-isothermal
conditions. The obtained results showed that dehydroxylation of
kaolinite is controlled by the rate of third-order reaction under
applied condition and the median of overall activation energy and
frequency factor correspond respectively to 227 1 kJ mol1 and
(9.4 0.3) 109 s1 .
In this study, samples of kaolin of the same mass were treated at
700 C for the same duration by varying the rate of calcination. The
metakaolins obtained were used for the synthesis of geopolymers
whose characteristics were determined with a view to evaluate the
effect of the rate of calcination of kaolin on their properties. To that
end, kaolin was treated at the following calcination speeds: 1, 2.5,
5, 10, 15 and 20 C/min. Depending on the materials, thermal and
chemical analyses, particle size distribution, specic surface area,

http://dx.doi.org/10.1016/j.jascer.2014.12.003
2187-0764 2014 The Ceramic Society of Japan and the Korean Ceramic Society. Production and hosting by Elsevier B.V. All rights reserved.

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

131

were respectively labeled as MK1 , MK2.5 , MK5 , MK10 , MK15 and


MK20 .
The alkaline solution was prepared by mixing sodium silicate
and sodium hydroxide solution (12 M) so as to obtain Na2 O/SiO2
molar ratio of 0.7. The sodium hydroxide solution was obtained by
dissolving sodium hydroxide pellets with a purity of 99% in distilled
water and sodium silicate was made up of (mass%) SiO2 (26.5), Na2 O
(8.0) and H2 O (65.5).
2.2. Paste and concrete samples preparation and analytical
techniques

Fig. 1. Particle size distribution of the clay fraction (K).

FTIR, XRD were performed on kaolin, metakaolins and geopolymers. The setting time and compressive strength were determined
as well on geopolymers.
2. Materials and methods
2.1. Materials processing
Metakaolins used in this study were obtained from kaolin provided by the NUBRU HOLDING Group which is involved in the
valorization of certain local raw materials in Cameroon. Before
being used, the kaolin was enriched with kaolinite by the sedimentometry process based on Stokess law [18]. The clay fraction
obtained labeled as K was rst maintained at ambient laboratory
temperature for a week, cured at 105 C until its mass became constant and then it was ground and sifted through a sieve of mesh
90 m. The resulting material was calcinated at 700 C, the choice
of this temperature being based on the work of Elimbi et al. [11].
Soaking of a mass of 1.5 kg of K at 700 C in an electric mufe furnace on ceramic plates (Nabertherm Mod LH 60/14) lasted 30 min
at the following rates of calcination: 1 C/min, 2.5 C/min, 5 C/min,
10 C/min, 15 C/min and 20 C/min. The obtained metakaolins

The preparation of geopolymer paste samples consisted of


mixing alkaline solution with metakaolin powder according to liquid/solid mass ratio of 1:1 in an automatic Hobart mixer (Controlab)
according to EN 196-1 standard [19]. Setting time of geopolymer
pastes was determined using Vicat apparatus according to EN 1961 standard [19]. For the determination of compressive strength,
cubic samples of (2 2 2 cm3 ) were made by mixing standardized sand [19] and metakaolin powder and alkaline solution in
mass ratio of 3:1:1. The molded mortar specimens were cured
at 20 C in a controlled room at 20 2 C and 98% RH condition.
Demolding was carried out 24 h later and the cubic samples were
covered with polyethylene lm and stored at ambient temperature (20 2 C) under 98% RH condition for 28 days. According to
the rate of calcination of kaolin, the cubic geopolymer specimens
were respectively labeled as GPC1 , GPC2.5 , GPC5 , GPC10 , GPC15 and
GPC20 . The cubic specimens aged of 28 days were subjected to
compressive strength using an automatic electro-hydraulic press
(IGM) operating at an average rate of 0.2 kN/s. Hardened geopolymer pastes aged of 28 days were crushed and sifted through a
sieve of mesh 80 m and powder was subjected to XRD (Siemens
D 5000 diffractometer using CoK radiation with a rear monochomator) and FTIR analysis using a Bruker Alpha-p IR spectrophotometer
operating in absorbance mode (interval of wave number ranging
was 4000400 cm1 ). The particle size distribution was determined using a granulometer (CILAS LIQUID 1090) whose interval
of investigation ranged between 0.04 and 500 m. The thermal
analyses (TG and DSC) of K were performed using a NETZSCH
STA-449F3 operating at the rate of 20 C/min in air. The chemical
analyses of K and metakaolins were performed according to the ICPOES process. Specic surface areas of metakaolins powders were

Fig. 2. Thermal analysis of the clay fraction (K).

132

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

Fig. 3. XRD pattern of the clay fraction (K).

Fig. 4. IR spectrum of the clay fraction (K).

Table 1
Bulk density and specic surface area of metakaolins.
Metakaolin

Bulk density (g/cm )


Specic surface area (cm2 /g)

MK1

MK2.5

MK5

MK10

MK15

MK20

2.54 0.02
128,710

2.54 0.02
128,200

2.54 0.02
129,480

2.56 0.02
125,490

2.54 0.02
128,720

2.5 6 0.02
127,790

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

133

determined according to Blaine method in a comparative goal and


their bulk densities were assessed by hydrostatic weighing.
3. Results and discussion
3.1. Characterization of raw materials

Fig. 5. Particle size distribution of metakaolins.

3.1.1. Characteristics of the clay fraction


The results of particle size distribution of K are shown in Fig. 1.
The curve represents the cumulative distribution of each class of
particles. The average particle size (d50 ) was 14 m and the chemical composition (mass%) of K was as follows: SiO2 (47.2), Al2 O3
(35.1), Na2 O (<0.1), K2 O (0.46), Fe2 O3 (0.46), TiO2 (0.49), CaO (<0.1),
SO3 (<0.01) and loss on ignition 14.94. Fig. 2 presented the results
of thermal analysis (TG and DSC) performed on K. The TG curve
showed three endothermic phenomena which correlated well with
those indicated by the DSC. The rst phenomenon went up to 100 C
and corresponded to elimination of water of hydration. The second one lay between 225 and 300 C with mass loss of 2.09% and
corresponds to the dehydroxylation of gibbsite [20]. The mass percentage of gibbsite was evaluated at 11.2%. The third phenomenon
between 450 and 550 C with mass loss of 9.49% represents the

Fig. 6. XRD patterns of metakaolins.

134

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

Table 2
Chemical compositions (mass%) of metakaolins.
Chemical composition

SiO2
Al2 O3
CaO
MgO
Fe2 O3
Na2 O
K2 O
TiO2
SO3
Loss on ignition
Total
SiO2 /Al2 O3 (molar ratio)

Metakaolin
MK1

MK2.5

MK5

MK10

MK15

MK20

54.60
40.60
0.12
0.22
0.53
0.12
0.53
0.57
0.01
2.26
99.57
2.28 0.01

53.60
39.86
0.11
0.22
0.52
0.11
0.52
0.56
0.01
5.11
100.64
2.28 0.01

52.50
39.04
0.11
0.21
0.51
0.11
0.51
0.55
0.01
6.92
100.49
2.28 0.01

52.10
38.74
0.11
0.21
0.51
0.11
0.51
0.54
0.01
6.93
99.78
2.28 0.01

52.00
38.67
0.11
0.21
0.51
0.11
0.51
0.54
0.01
6.95
99.62
2.28 0.01

51.40
38.22
0.11
0.21
0.50
0.11
0.50
0.53
0.01
7.73
99.34
2.28 0.01

dehydroxylation of kaolinite and gives rise to the mass percentage


of kaolinite to 81%. The crystalline phases revealed by the X-ray
diffractogram (Fig. 3) indicated mainly the presence of kaolinite,
quartz, gibbsite and muscovite. Fig. 4 showed the FTIR analysis of

K. The absorption bands at 36883619 cm1 express the stretching


vibrations of OH groups of kaolinite and gibbsite networks [21,22].
The bands located at 3450 and 1650 cm1 corresponded respectively to stretching vibrations of water molecules [23] while those

Fig. 7. IR spectra of metakaolins.

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

135

Fig. 8. XRD patterns of geopolymers.

at 1114 and 1025 cm1 and then at 998 and 789 cm1 expressed
respectively the vibrations of Si O Si and Si O Al groups of the
network. The bands at 908 and at 789 cm1 indicated the presence
of the stretching vibration of Al OH with Al in coordination VI. The
band at 522 cm1 indicated the vibrations of Si O Si and Si O Al
groups of the network.
3.1.2. Characteristics of metakaolins
Thermal treatment of kaolin generally causes transformations
which effects are among other things increasing of specic surface area and average particle grain size [20]. From Table 1 it
was clear that the variation of the rate of calcination of kaolin
had no signicant effect on the bulk density of metakaolins. As
for the specic surface area of metakaolins, an outline observation on the different values did not allow to draw meaningful
conclusion on the effect of the rate of calcination of kaolin. According to the curves of Figs. 1 and 5, metakaolins and powder of K
almost exhibited the same particle size distribution tendency. The
chemical composition (Table 2) showed that the mass percentages of SiO2 and Al2 O3 decreased contrary to the loss on ignition.
This meant that when the rate of calcination of kaolin increased,
the minerals in the kaolin (e.g. gibbsite and kaolinite) were less
decomposed and led to a lower production of active compounds

for the geopolymer synthesis. With regard to kaolinite particularly,


its transformation into metakaolin fell away with the increase of
the rate of calcination (Fig. 6: increase of intensity of the main
peak of kaolinite). Consequently loss on ignition became increasingly high because of the presence of increasing amount of water
released during the dehydroxylation of residual kaolinite. Decrease
of mass percentage of SiO2 and Al2 O3 with increase of the rate
of calcination correlated well with the increasing of the loss on
ignition since it was taken into account for the determination of
mass percentage of oxides present in metakaolins. However, this
decrease did not affect the molar ratio Al2 O3 /SiO2 (2.28 0.01)
which remained constant because of conservation of matter. The
X-ray diffractograms of metakaolins (Fig. 6) showed diffuse halo
peak with 2 between 18 and 38 which is characteristic of amorphous phase present in metakaolins [24,25]. All minerals initially
present in K remained in metakaolins except gibbsite (MK1 to
MK20 ) and kaolinite (MK1 ). Indeed except for MK1 where it was
of kaolinite was observed in all
absent, the main peak (7.15 A)
diffractograms with intensity which increased with the rate of
calcination. This behavior was in accordance with the variation
of the loss on ignition in metakaolins (Table 2). Thus, the lower
the rate of calcination, the more complete was the transformation of kaolin into metakaolin. According to Castelin et al. [15],

136

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

Fig. 9. IR spectra of geopolymers.

during the heating of kaolin, kaolinite dehydroxylation, metakaolinite structure change, exothermic structural organization along
with mullite formation are very sensitive to the heating temperature rate. In particular, the lower the rate of calcination of kaolin,
the lower is the temperature of commencement of dehydroxylation
of kaolinite [15]. Accordingly, when kaolin was calcined at 700 C
for the same soaking time, kaolinite was more dehydroxylated for
a low heating rate. FTIR spectra of metakaolins (Fig. 7) also conrmed that except for MK1 , metakaolins contained kaolinite. In fact
the absorption bands assigned to kaolinite (36883648, 3619, 908
and 788 cm1 ) were observed in the spectra of MK2.5 MK20 . Conversely, the FTIR spectrum of MK1 was mainly characterized by the
band at 1046 cm1 which corresponded to stretching vibration of
Si O Si and Si O Al groups of the metakaolin network [26].
3.2. Characterization of geopolymers
3.2.1. XRD
The X-ray diffractograms of geopolymer pastes (Fig. 8) showed
the presence of all the crystalline phases initially encountered

in metakaolins (Fig. 6). However, the comparison between


Figs. 6 and 8 revealed certain changes with regard to the halo peak
representing the amorphous phase. The halo peak with 2 between
18 and 38 for metakaolins (Fig. 6) was now between 20 and 45
for geopolymers (Fig. 8) which is ngerprint of geopolymerization
[27,28].
3.2.2. FTIR
In FTIR spectra of geopolymers (Fig. 9) the absorption bands at
34503480 and 1650 cm1 respectively corresponded to stretching and deformation vibrations of OH and H O H bonds of water
molecules [2931]. Also the absorption band at 13911410 cm1
referred to stretching vibrations of C O bond of carbonate [32,33]
in sodium carbonate which generally induces eforescence within
geopolymers. The absorption band at 1005 cm1 expressed the
asymmetric and symmetric vibrations of Si O Al and Si O Si
bonds of SiO4 and AlO4 tetrahedrons of the geopolymer network
[34,35]. However, contrary to GPC1 , the FTIR spectra of the geopolymers GPC2.5 GPC20 exhibited absorption bands at 36803640 cm1
which are the ngerprint of the network of kaolinite. Thus, as for

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

137

3.2.4. Compressive strength


There was a clear relation between the compressive strength
of geopolymer mortars and the loss on ignition of metakaolins
(Fig. 11). Unlike the loss on ignition, the compressive strength was
more and more high when the rate of calcination was low. According to the spectra of X-ray diffraction of metakaolins (Fig. 6), the
of kaolinite was stronger when
intensity of the main peak (7.15 A)
the rate of calcination was higher. In this context, residual kaolinite was mainly responsible for the increase of the loss on ignition
which was a good indication of the presence of reactive phase: the
higher the loss on ignition, the lesser metakaolin contained reactive
phase allowed for geopolymer synthesis. The presence of abundant reactive phase in aluminosilicate is known to lead to higher
value of compressive strength [11,39]. Thus, compressive strength
of geopolymers was higher for metakaolin which did not contain
residual kaolinite and this was achieved when kaolin is calcinated
at a low speed.
Fig. 10. Relation between setting time (geopolymers) and loss on ignition
(metakaolins).

XRD, FTIR analysis showed that the calcination of kaolin at a high


speed led to geopolymers which still contained kaolinite as a result
of incomplete dehydroxylation. In fact geopolymers can be synthesized between kaolin and alkali activator solution. However,
kaolinite always shows low reactivity [36,37].
3.2.3. Setting time
There was increase of setting time of geopolymer pastes with
increase of the rate of calcination of kaolin and it appeared as a
correlation between setting time and loss on ignition (Fig. 10). The
increase in loss on ignition with the rate of calcination (Table 2) was
inuenced among other things by the dehydroxylation of residual
kaolinite contained in metakaolins. It is well known that amorphous phase in metakaolin is the phase which is reactive in alkaline
medium to produce geopolymers [32,34,38]. Figs. 6 and 10 lead up
to consider that during the geopolymer synthesis, increase in setting time with respect to the rate of calcination is regulated not only
by an amorphous phase but also by the presence of quasi no reactive kaolinite. Thus, both amorphous phase and residual kaolinite
were limiting factors in geopolymerization. The lower was residual
kaolinite in metakaolin, the faster was the geopolymer synthesis
and this was better accomplished when kaolin was calcined at a
low speed.

4. Conclusion
Metakaolins obtained at 700 C for the same duration (30 min)
by varying the rate of calcination (120 C/min) of samples of equal
mass of kaolin were used to produce geopolymers. Chemical analysis of metakaolins revealed that loss on ignition increased with
the rate of calcination of kaolin. This resulted from the presence
of residual kaolinite in metakaolins and whose quantity increased
with the rate of calcination of kaolin. XRD and FTIR analyses of
geopolymers also revealed the presence of residual kaolinite. Thus
the synthesis of geopolymers was affected by both amorphous
phase and quasi no residual kaolinite of metakaolins. This simultaneous effect impacted on the setting time and the compressive
strength: the setting time increased while the compressive strength
fell away. To produce geopolymer having low setting time and
high mechanical strength, the calcination of kaolin needs to be
performed at a low speed.
Acknowledgments
One of us (B.B. Kenne Diffo) is delighted to acknowledge the
French Government for granting a scholarship for PhD studies
through the Service de la Coopration et dAction Culturelle
(SCAC) of the Embassy of France in Cameroon and the staff of the
Laboratoire Matriaux et Durabilit des Constructions (LMDC) for
the reception and infrastructural support he received during his
stay at INSAUPS Toulouse (France). Also, nancial supports from
the Bureau Afrique Centrale et des Grands Lacs de lAgence Universitaire de la Francophonie (BACGL-AUF) under the grant N S0020
*** 10406 and The World Academy of Sciences (TWAS) under the
grant N 13-018 RG/CHE/AF/AC G - UNESCO FR: 3240277723 are
acknowledged.
References

Fig. 11. Relation between compressive strength (geopolymers) and loss on ignition
(metakaolins).

[1] M.J. Sumajouw, D. Hardjito, S.E. Wallah and B.V. Rangan, J. Mater. Sci., 42,
31243130 (2007).
[2] N. Ganesan, P.V. Indira and A. Santhakumar, Constr. Build. Mater., 48, 9197
(2013).
[3] A. Kumar and S. Kumar, Constr. Build. Mater., 38, 865871 (2013).
[4] M. Izquierdo, X. Querol, J. Davidovits, D. Antenucci, H. Nugteren and C.
Fernndez-Pereira, J. Hazard. Mater., 166, 561566 (2009).
[5] T.W. Cheng, M.L. Lee, M.S. Ko, T.H. Ueng and S.F. Yang, Appl. Clay Sci., 56, 9096
(2012).
[6] A. Rooses, P. Steins, A. Dannoux-Papin, D. Lambertin, A. Poulesquen and F.
Frizon, Appl. Clay Sci., 73, 8692 (2012).
[7] M.B. Ogundiran, H.W. Nugteren and G.J. Witkamp, J. Hazard. Mater., 248249,
2936 (2013).
[8] K. Al-Zboon, M.S. Al-Harahsheh and F.B. Hani, J. Hazard. Mater., 188, 414421
(2011).

138

B.B. Kenne Diffo et al. / Journal of Asian Ceramic Societies 3 (2015) 130138

[9] H. Oudadesse, A.C. Derrien, M. Leoch and J. Davidovits, J. Mater. Sci., 42,
30923098 (2007).
[10] J. Davidovits, Geopolymer Chemistry and Applications, 2nd ed., Institut
Gopolymre, Saint-Quentin (2008).
[11] A. Elimbi, H.K. Tchakoute and D. Njopwouo, Constr. Build. Mater., 25,
28052812 (2011).
[12] K.H. Tchakoute, A. Elimbi, D.B.B. Kenne, J.A. Mbey and D. Njopwouo, Ceram. Int.,
39, 269276 (2013).
[13] C. Kuenzel, T.P. Neville, S. Donatello, L. Vandeperre, A.R. Boccaccini and C.R.
Cheeseman, Appl. Clay Sci., 8384, 308314 (2013).
[14] R.S. Nicolas, M. Cyr and G. Escadeillas, Appl. Clay Sci., 8384, 253262 (2013).
[15] O. Castelein, B. Soulestin, J.P. Bonnet and P. Blanchart, Ceram. Int., 27, 517522
(2001).
[16] O. Castelein, L. Aldon, J. Olivier-Fourcade, J.C. Jumas, J.P. Bonnet and P. Blanchart,
J. Eur. Ceram. Soc., 22, 17671773 (2002).

[17] P. Ptcek, D. Kubtov, J. Havlica, J. Brandstetr, F. Soukal


and T. Opravil, Powder
Technol., 204, 222227 (2010).
[18] R. Dupain, R. Lanchon and J.C. Saint Arroman, Granulats, sols, ciments et btons,
Casteilla, Paris (2000).
[19] European Standard Institution [NF EN196-1], Methods of Testing Cement:
Determination of Strength (2006).
[20] B. Fabbri, S. Gualtieri and C. Leonardi, Appl. Clay Sci., 73, 210 (2013).
[21] E. Gasparini, S.C. Tarantino, P. Ghigna, M.P. Riccardi, E.I. Cedillo-Gonzlez, C.
Siligardi and M. Zema, Appl. Clay Sci., 8081, 417425 (2013).
[22] E. Yusiharni and R. Gilkes, Appl. Clay Sci., 24, 6174 (2012).
[23] R.A. Antunes, B. Santa, A. Michael, H. Gracher and N. Cabral, J. Clean. Prod., 57,
302307 (2013).

[24] D. Panias, I.P. Giannopoulou and T. Perraki, Colloids Surf. A, 301, 246254
(2007).
[25] C. Bich, J. Ambroise and J. Pra, Appl. Clay Sci., 44, 194200 (2009).
[26] A. Sathonsaowaphak, P. Chindaprasirt and K. Pimraksa, J. Hazard. Mater., 168,
4450 (2009).
[27] A. Palomo and F.P. Glasser, Br. Ceram. Trans. J., 91, 107112 (1992).
[28] P. Rovnanik, Constr. Build. Mater., 24, 11761183 (2010).
[29] V.F.F. Barbosa, K.J.D. Mackenzie and C. Thaumaturgo, Int. J. Inorg. Mater., 2,
309317 (2000).
[30] P. Duxson, S.W. Mallicoat, G.C. Lukey, W.M. Kriven and J.S.J. van Deventer,
Colloids Surf. A: Physicochem. Eng. Aspects, 292, 1820 (2007).
[31] U. Rattanasak and P. Chindaprasirt, Miner. Eng., 22, 10731078 (2009).
[32] A.R. Sakulich, E. Anderson, C. Schauer and M.W. Barsoum, Constr. Build. Mater.,
23, 29512957 (2009).
[33] I. Lancellotti, M. Catano, C. Ponzoni, F. Bollino and C. Leonelli, J. Solid State
Chem., 200, 341348 (2013).
[34] D. Akolekar, A. Chaffee and R.F. Howe, Zeolites, 19, 359365 (1997).
[35] M. Gougazeh and J.-C. Buhl, Appl. Clay Sci., 15, 3542 (2014).
[36] C.Y. Heah, H. Kamarudin, A.M. Mustafa Al Bakri, M. Bnhussain, M. Luqman, I.
Khairul Nizar, C.M. Ruzaidi and Y.M. Liew, Constr. Build Mater., 35, 912922
(2012).
[37] A.D. Housni, G.L. Lecomte-Nana, G. Djtli and P. Blanchart, Constr. Build.
Mater., 42, 105113 (2013).
[38] M.L. Granizo, S. Alonso, M.T. Blanco-Varela and A. Palomo, J. Am. Ceram. Soc.,
85, 225231 (2012).
[39] P. Duxson, The Structure and Thermal Evolution of Metakaolin Geopolymers
(Ph.D. thesis), University of Melbourne, Melbourne (2006).

You might also like