You are on page 1of 10

Construction and Building Materials 102 (2016) 260–269

Contents lists available at ScienceDirect

Construction and Building Materials


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

Effect of fly ash characteristics on delayed high-strength development of


geopolymers
J.G. Jang, H.K. Lee ⇑
Department of Civil and Environmental Engineering, Korea Advanced Institute of Science and Technology, 291 Daehak-ro, Yuseong-gu, Daejeon 305-701, South Korea

h i g h l i g h t s

 Effect of fly ash on the strength development of geopolymer was investigated.


 Densification of pore structures had little correlation with strength development.
 Delayed strength development mainly occurred due to the gel transformation.
 It was more evident as vitreous phase content and SiO2/Al2O3 ratio of fly ash increased.

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

Article history: This paper investigates the effect of fly ash characteristics on the strength development of fly ash-based
Received 4 May 2015 geopolymer, and reports an observation of delayed high-strength development of geopolymer. Physical
Received in revised form 7 October 2015 and chemical characteristics of fly ashes were determined by XRF, XRD and particle size analyzer.
Accepted 28 October 2015
Multi-technical characterizations using MIP, FT-IR and SEM & EDS were applied to give in-depth under-
Available online 4 November 2015
standing of the relationship between reaction products, microstructure and strength development
according to the maturity of geopolymer. Test results provided an insight into the delayed strength devel-
Keywords:
opment of geopolymer which mainly occurs due to the transformation of aluminosilicate gel from Al-rich
Geopolymer
Fly ash
gel to Si-rich gel, rather than densification of pore structures. As particle size of fly ash increased, the
Compressive strength strength development of geopolymer was delayed. The delayed high-strength development was more
Aluminosilicate gel evident as the vitreous phase content and SiO2/Al2O3 ratio of fly ash increased. Furthermore, the effect
Microstructure of crystal growth on the strength development was discussed.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction origin of the fly ash [1,2]. While Portland cement possesses
hydraulic properties due to the large amounts of reactive CaO
Fly ash, which is an industrial by-product generated from coal and SiO2, most fly ash itself is not hydraulic at room temperature.
combustion power plants, is a spherical powder with a specific However, alkali-materials can activate the reactivity of fly ash [3].
gravity range of 2.1–3.0 [1]. Its use in the cement and concrete Davidovits introduced a reaction between metakaolin and alkali-
fields has been thoroughly studied. When it is used as an admix- activator at high temperature to form aluminosilicate gel, which
ture for Portland cement concrete it can be expected to increase is a compound called geopolymer in recent years and is extensively
workability, reduce hydration heat, enhance long-term strength being studied due to its wide range of potential applications [4–8].
and durability (resistance against chloride penetration, alkali- Although the chemical composition of geopolymer is known to be
silica reactions and sulfate attack), and contribute a light-weight similar to zeolite, its main component is amorphous aluminosili-
effect [2]. Extensive studies focused on the reduction of CO2 emis- cate. The general formula of typical geopolymer composition is
sions and the practical use of wastes has evaluated fly ash as a con- expressed as M n ½ðSi  O2 Þz  Al  On wH2 O where M is an alkali
struction material. metal, z is 1, 2, or 3, and n is the degree of polymerization [9]. Many
The chemical composition of fly ash mainly consists of SiO2, studies have reported that geopolymer is highly resistive against
Al2O3 and Fe2O3, but it may contain some CaO depending on the acid, high temperature and alkali-silica reactions, and can exhibit
high strength at an early age when it is cured at high temperature
[7,10]. Moreover, utilizing fly ash as the source material for
⇑ Corresponding author. geopolymer is an attractive strategy for making environmentally
E-mail address: haengki@kaist.ac.kr (H.K. Lee).

http://dx.doi.org/10.1016/j.conbuildmat.2015.10.172
0950-0618/Ó 2015 Elsevier Ltd. All rights reserved.
J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269 261

friendly concrete, since the amount of CO2 released during the pro- obtained from thermal power plants in South Korea and used as
duction of geopolymer is less than that of Portland cement paste source materials to synthesize geopolymers. This allowed a
[3–5]. For those reasons, many studies focusing on reaction kinet- comparison of the effect of the different fly ash properties on the
ics, mechanical strength and durability have been conducted to compressive strength, phase and microstructure developments.
develop new alternatives to Portland cements. An alkali-activator was made by mixing 9 M NaOH solution and
To synthesize geopolymer, fly ash (alternatively, metakaolin) sodium silicate (Korean Industrial Standards KS Grade-3;
and an alkali activator such as sodium silicate, sodium hydroxide, SiO2 = 29 wt%, Na2O = 10 wt%, H2O = 61 wt%, specific grav-
potassium hydroxide, or calcium hydroxide, etc. are generally used ity = 1.38) with mass ratio of 1:1. The alkali-activator was cooled
[7]. It is a known fact that the physical and chemical natures of the at room temperature a day before casting. Geopolymer samples
fly ash and the alkali-activator have a significant impact on reac- were fabricated by mixing fly ash and alkali-activator at a constant
tion process. However, it should be emphasized that the chemical fly ash/activator mass ratio of 2. The fly ash and alkali-activator
composition of the fly ash changes dramatically depending on the were mixed for five minutes at room temperature, and then the
origin of the coal fuel and type of combustion system, and has a mixtures were cast into a cubic mold with a size of 5  5  5 cm.
greater effect overall than the quality of the alkali-activator [1]. The samples were sealed to prevent water evaporation, and cured
Previous studies dealing with the strength development of at 60 °C for 24 h in an oven. After initial curing, all samples were
geopolymer, have mainly focused on the effect of the type, compo- kept in a chamber with a relative humidity of 65% at 20 °C until
sition and concentration of the alkali-activator, and ionic additives, the planned test date. The labels ‘GP1’, ‘GP2’ and ‘GP3’ in this paper
curing conditions and calcium-rich admixture [6,11–16], while represent the geopolymer samples made from three types of fly
only a few studies have addressed the effect of the characteristics ashes, ‘FA1’, ‘FA2’ and ‘FA3’, respectively.
of fly ash on the strength development of geopolymer [17,18].
The curing temperature of geopolymer usually takes place at
2.2. Testing and characterization
40–95 °C and is slightly higher compared with that of Portland
cement which generally requires room temperature for curing
For chemical characterization of the raw fly ashes, chemical
[6,13,19,20]. High temperature curing of geopolymer promotes
composition was determined by X-ray fluorescence (XRF) spec-
the activation of fly ash, production of aluminosilicate gel and early
trometer. The Minipal 2 from PANanalytical was used for the XRF
gain in high strength [21]. This latter phenomenon was confirmed
analysis. The mineral composition of the raw fly ashes was
by the observation in previous studies that the ultimate strength of
analyzed by high resolution X-ray diffractometer (HR-XRD) with
geopolymer was achieved at a significantly early age. A previous
Rietveld refinement method. The Empyrean from PANanalytical
study regarding the role of Al2O3 and SiO2 in geopolymerization
with CuKa radiation was used for the XRD analysis. The XRD was
conducted by Silva et al. revealed that the strength of a geopolymer
operated at 40 kV and 30 mA with a scan range of 5–90° and a step
cured at a temperature of 60 °C showed no significant change
size of 0.02°. The physical characterization of the raw fly ashes
between 1 and 24 h of curing time [19]. It was also reported by
included a determination of the particle size distribution by parti-
Khale and Cuaudhary that 70% of the ultimate compressive
cle size analyzer. The Helos from Sympatech was used for this test.
strength was gained within the initial 3–4 h [6]. It is especially
The compressive strength of hardened geopolymer samples was
worth noting that geopolymer undergoes a contrasting strength
tested at 1, 3, 4, 28, 56, 91 days using a 3000 kN universal testing
gaining process, as Portland cement paste gains strength and a
machine in accordance with ASTM C109 [25]. The loading speed
hydration reaction continues proportional to the age of the cement
was set to 0.01 mm/s. The strengths were the average of three sep-
paste [6]. Similarly, the high compressive strength gain within a
arate tests. The microstructural characteristics of the hardened
few hours or days has been observed in a number of previous stud-
geopolymer samples were evaluated by means of mercury intru-
ies [3,13,17,22,23]. Consequently, the prediction of geopolymer
sion porosimetry (MIP) using an Autopore VI machine by
strength is often based on strength measurement at an early age,
Micromeritics Corp. The pressure range from 30 to 60,000 psia
while the characteristics of geopolymer strength as they develop
was selected to detect a pore diameter in the range of 3 nm to
over time, in a similar manner to Portland cement, are hardly
6 lm. Despite the fact that the use of MIP for determining the pore
focused on. Predicting and understanding the strength develop-
structure of a cementitious material has limitations [26], this
ment of geopolymer at later ages is important from both economic
method is still useful to estimate pore size distribution of cemen-
and technical points of view. Moreover, geopolymer has a wide
titious materials, if used with a proper care during sample prepara-
range of applications, as a construction material itself, and as a bin-
tion. The pore size distribution curve and pore structure obtained
der for concrete and ground reinforcement, and so the effect of
by MIP can provide a better understanding of the effect of fly ash
characteristics of fly ash on the behavior of strength development
characteristics on the connectivity of pore structure and porosity
should be understood.
of geopolymer [27,28].
In the present study, the effect of fly ash characteristics on the
The Fourier transform infrared (FT-IR) spectra of the raw fly
strength development of fly ash-based geopolymer was investi-
ashes and powdered geopolymer samples were obtained using
gated. A representative group of Korean class F fly ashes was used
FT-IR spectroscopy (Model FT-IR 4100, Jasco, Japan). Spectra were
for geopolymer synthesis with 9 M NaOH solution and sodium sil-
collected from 400 to 4000 cm1 at a resolution of 0.9642 cm1.
icate. Physical and chemical characteristics of the fly ashes were
The microstructure and phase of geopolymers were analyzed by
determined, and the relationships between the reaction product,
field emission scanning electron microscopy (SEM) and electron
microstructure, and strength development of geopolymer with dif-
dispersive X-ray spectroscopy (EDS). The Magellan400 from FEI
ferent maturities were discussed on the basis of multi-technical
company was used for the test.
analysis.

2. Experimental procedure 3. Test results

2.1. Materials and sample preparation 3.1. Characterization of raw fly ash

Three types of Class F fly ash as defined in ASTM C618 [24], The chemical composition of the fly ashes as determined by XRF
which had different chemical and mineral compositions, were is given in Table 1. Pozzolanic compounds of fly ashes in Table 1
262 J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269

Table 1 Q
Chemical composition of the fly ashes as determined by XRF.

Composition (wt%) FA1 FA2 FA3


Al2O3 27.6 22.2 22.0
SiO2 50.4 50.1 51.5
Q
CaO 2.9 5.2 5.3 M
M HH
Fe2O3 8.0 10.1 10.8 FA3 MMM M M M
K2O 1.6 1.3 1.2 Ma QQ Q M
M Q M Q Ma Q M Q
Na2O 0.6 0.5 0.6 FA2
MgO 2.0 2.0 2.0
SO3 0.9 1.0 0.7
Cl 0.1 0.1 0.3
FA1
TiO2 2.1 1.7 1.6
MnO 0.1 0.1 0.1
P2O5 0.3 0.4 0.2
5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90
Pozzolanic compounds (SiO2 + Al2O3 + Fe2O3) 86.0 82.4 84.3
SiO2/Al2O3 1.83 2.26 2.34 2-Theta (º)
Loss on ignition (up to 1000 °C) 3.57 2.40 2.58
Fig. 1. XRD patterns of the fly ashes: M, Q, H and Ma indicate mullite, quartz,
hematite and magnetite, respectively and a halo pattern recorded for 2 theta of 20–
35° indicates the presence of vitreous phase.

are calculated in accordance with ASTM C618 [24]. The most abun-
dant oxides in the fly ashes are SiO2 and Al2O3 followed by Fe2O3, Table 2
CaO, MgO, TiO2 and K2O. The pozzolanic compound content was Mineral composition of fly ashes as determined by XRD Rietveld refinement method.
the highest in FA1 with 86.0 wt%, followed by FA3 (84.3 wt%) and
Mullite Quartz Hematite Magnetite Vitreous phase
FA2 (82.4 wt%). The CaO content in FA2 and FA3 was 5.2 and
FA1 13.2 7.9 0.4 0.3 78.2
5.3 wt% respectively, indicating no significant difference, whereas
FA2 9.3 7.7 – 0.5 82.4
it was 2.9 wt% in FA1. SiO2/Al2O3 which is a dominant factor on FA3 7.8 7.6 – 0.7 84.0
the fresh and hardened properties of geopolymer [17,19,29,30]
was the highest in FA3 (2.34), followed by FA2 (2.26) and FA1
(1.83). The unburned content of material, which can affect the 100
reactivity of fly ash, was evaluated by measuring the amount of
Cummulative distribution (%)

90 (a) FA1
loss on ignition. The measured values of loss on ignition were all 80
less than 5%, implying that the fly ashes were suitable to be a 70
FA2
source material for geopolymer synthesis [17]. 60 FA3
XRD patterns of the fly ashes are shown in Fig. 1. The XRD anal- 50
ysis aims to quantify unreactive crystals such as mullite, quartz, 40
hematite and magnetite, and reactive vitreous phase for geopoly- 30
merization. Duxson et al. revealed that quartz and mullite particles 20
incorporated in fly ash behave as a micro-aggregate during the 10
geopolymerization process, and glassy materials (i.e., vitreous 0
phases) become a source material for the formation of aluminosil- 0.1 1 10 100 1000
icate gel [29]. Thus, the Rietveld refinement method was adopted Particle size ( m)
to quantitatively analyze the mineral composition of each material.
0.8
The vitreous phase was derived from a halo pattern observed at 2
0.7 (b)
Density distribution (%)

theta of 20–35° as shown in Fig. 1 [30,31]. The mineral composition


of the fly ashes as determined by the XRD Rietveld refinement 0.6
FA1
method is given in Table 2. FA1 contains 13.2% mullite, 7.9% quartz, 0.5
0.4% hematite, 0.3% magnetite and 78.2% vitreous phase; FA2 con- FA2
0.4
tains 9.3% mullite, 7.7% quartz, 0.5% magnetite and 82.4% vitreous FA3
0.3
phase; FA3 contains 7.8% mullite, 7.6% quartz, 0.7% magnetite and
84.0% vitreous phase, showing the highest vitreous phase. 0.2
Particle size distributions of the fly ashes are shown in Fig. 2. 0.1
Particle size distribution is one of the factors which most strongly 0.0
affects the reactivity of fly ash [17]. The measured mean particle 0 1 10 100 1000
diameters of FA1, FA2 and FA3 were 10.03, 13.13 and 16.49 lm, Particle size ( m)
respectively. It is worth mentioning that this value is relatively Fig. 2. Particle size distribution of the fly ashes: (a) cumulative distribution and (b)
small, as the mean particle size of raw fly ash used for geopolymer density distribution.
synthesis in many studies ranges from 10.9 to 44.4 lm (e.g., 41 lm
in [20], 10.9–19.1 lm in [32], 19 lm in [33], 14.4 lm in [34], and
44.4 lm in [35]). The BET surface areas of FA1, FA2 and FA3 were strengths at 91 days were 42.7, 48.7 and 53.0 MPa for GP1, GP2
estimated by this method to be 9853.8, 8467.8 and 7440.7 cm2/g, and GP3, respectively, showing that the highest compressive
respectively. strength was achieved by GP3. Although a logarithmic scale
increase in the compressive strength was experienced by all
3.2. Compressive strength development of geopolymer geopolymer samples, strength development was significantly dif-
ferent in each sample. The strength development of GP1 was rather
The compressive strength of geopolymer samples measured at fast compared to GP2 and GP3, and the incremental change in
1, 3, 7, 28, 56 and 91 days is shown in Fig. 3. The compressive strength was rather small in GP1 after passing 28 days. Although
J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269 263

60 Table 3
GP3 Additional details of geopolymers from previous studies referenced in this study.

50
GP2 Refs. Selected Specimen Materials used Curing condition
Compressive strength (MPa)

no. of code in (temp., time)


samples the Ref.
40
GP1 [3] 3 – Fly ash, NaOH, sand 45–85 °C, 24 h
[17] 5 C, L, PN, R, Fly ash, NaOH, sand 85 °C, 20 h
30
T
[19] 6 Si 30, Si Metakaolin, sodium 40 °C, 72 h
34, Si38, silicate, NaOH
20 Al06,
Al08, Al10
[20] 4 – Fly ash, NaOH 25–28 °C, until test
10 date
[22] 3 CFA, MFA, Fly ash, NaOH, 75 °C, 48 h
FFA sodium silicate, sand
0 [23] 4 – Fly ash, NaOH Pre-cured at room
13 7 28 56 91
temperature, 2 or
Time (d) 24 h; then 75 or
95 °C, until test date
Fig. 3. Compressive strength of geopolymer samples measured at 1, 3, 7, 28, 56 and
[30] 4 N, W15, Fly ash, NaOH; Fly 85 °C, until test date
91 days.
W50, ash, NaOH, sodium
W84 silicate
[36] 10 A1, A2, Fly ash, NaOH, 85 °C, 48 h
GP2 had similar strength development to GP1 at early age, the A3, A4, sodium silicate, sand
strength constantly increased even after 28 days. In contrast, the A5, S1, S2,
S3, S4, S5
strength development of GP3 was clearly delayed. Although the [37] 2 Sample A, Fly ash, geothermal 40 °C, until test date
GP3 strength at 28 days was lower than the other two, its strength Sample B silica, sodium
development gradually increased until 91 days, exhibiting delayed aluminate (or
high-strength development. amorphous alumina),
NaOH
The relative development of compressive strength of the
geopolymers tested in this study and previous studies
[3,17,19,20,22,23,30,36,37] is shown in Fig. 4. Additional details It was also notable that the UCS of GP3 was 1.6 times greater than
of geopolymers from previous studies referenced in this study its compressive strength at 28 days.
are provided in Table 3. It should be noted that the relative com-
pressive strength in Fig. 4 is the value relative to the ultimate com- 3.3. Porosity of geopolymer
pressive strength (UCS) given the age of the geopolymer sample. At
7 days, the relative values of compressive strength of GP1, GP2 and Pore size distributions of geopolymer samples measured at 3
GP3 were 61.0%, 49.5% and 31.2% of the UCS, respectively, indicat- and 91 days are shown in Fig. 5. In addition, the values of
ing that the strength development was relatively slower compared cumulative pore area, median pore diameter and porosity of the
to previous studies [3,17,19,23,36] where rapid strength develop- geopolymer samples are listed in Table 4. The MIP test result
ing was observed at very early age. Particularly, the strength devel- shows that the cumulative pore areas of GP1, GP2 and GP3 at
opment of GP3 was undoubtedly slower, and the relative value of 91 days were 22.1, 28.2 and 34.9 m2/g, respectively, and were
compressive strength of GP3 after 3 days showed a linear increase. reduced in comparison with that at 3 days. This result indicates

110
[3]
100 [17]
[19]
90 [20]
Relative compressive strength (%)

[22]
80
[23]
[30]
70
[36]
[37]
60
GP1
GP2
50
GP3
40

30

20

10

0
0
0.01 0.1 1 3 7 10 14 28 56 100 180 1000
Time (d)

Fig. 4. The relative development of compressive strength of the geopolymers tested in this study and previous studies [3,17,19,20,22,23,30,36,37].
264 J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269

0.25 0.5 Table 4


Cumulative pore area, median pore diameter and porosity of the geopolymer samples.
Cumulative intrusion (mL/g)

91 d (a) GP1
0.2 0.4 Sample code Age (d) Cumulative pore Median pore Porosity (%)

ΔV/ΔlogD (mL/g)
91 d
area (m2/g) diameter (nm)
0.15 3d 0.3 GP1 3 25.9 15.5 30.5
91 22.1 28.6 29.4
3d
0.1 0.2 GP2 3 32.9 12.2 29.8
91 28.2 20.0 33.3
0.05 0.1 GP3 3 41.3 12.4 29.1
91 34.9 19.2 31.4
0 0
1 10 100 1000
Pore diameter, D (nm)
shown in Fig. 5, the threshold pore diameter decreased signifi-
0.25 0.5 cantly in GP1 with age, while it was not significantly changed in
Cumulative intrusion (mL/g)

(b) GP2 GP2 and GP3. From the result of the threshold pore diameter, it
0.2 91 d 0.4 can be said that the microstructural change of geopolymer over

ΔV/ΔlogD (mL/g)
3d time is not significant, since the porosity obtained in this study
91 d
0.15 0.3 may contain errors originated from the MIP method.
3d
0.1 0.2
3.4. Phase development of geopolymer
0.05 0.1
The FT-IR spectra of raw fly ashes and geopolymer samples at 3
and 91 days are presented in Fig. 6. In addition, the major IR vibra-
0 0
1 10 100 1000
Pore diameter, D (nm)
70
0.25 0.5
60
(a) GP1
Cumulative inrtusion (mL/g)

(c) GP3 Raw fly ash


Absorbance (%)

0.2 0.4 50
91 d 91 d Geopolymer-3 d
ΔV/ΔlogD (mL/g)

3d 40
Geopolymer-91 d
0.15 0.3
30
3d
0.1 0.2 20

0.05 0.1 10

0
0 0 4000 3600 3200 2800 2400 2000 1600 1200 800 400
1 10 100 1000 Wavenumber (cm-1)
Pore diameter, D (nm)
70
Fig. 5. Pore size distribution of geopolymer samples measured at 3 days and (b) GP2
91 days. 60
Raw fly ash
Absorbance (%)

50
that geopolymerization continued and reaction products were pro- Geopolymer-3 d
duced even after 3 days. In contrast, the median pore diameter of 40
Geopolymer-91 d
all geopolymer samples increased with age. Furthermore, it was 30
shown that the porosity of GP2 and GP3 was significantly increased
with age, while that of GP1 was slightly decreased with age. It is 20

worth mentioning that the observed result was the opposite of 10


Portland cement paste, whose pore diameter and porosity gener-
0
ally reduce as hydration proceeds [38,39]. Furthermore, a close 4000 3600 3200 2800 2400 2000 1600 1200 800 400
correlation between the compressive strength and porosity of Wavenumber (cm-1)
cementitious materials has been noticed in a number of studies
[40,41]. However, there was no clear evidence in the relationship 70

between porosity and compressive strength of the geopolymer 60


(c) GP3
tested in this study. Ma et al. also reported similar observations
Absorbance (%)

Raw fly ash


50
on the pore characteristic of geopolymer [28]. This may be attrib- Geopolymer-3 d
uted to the different pore shapes and ‘‘ink-bottle” effects, meaning 40
Geopolymer-91 d
that the pore size revealed by MIP tests may not be the exact
30
porosity and pore size of the material [26]. Since geopolymer has
more complex pore structure, it is more likely that the ink-bottle 20
effect had larger influence on the porosity value measured by
10
MIP [42]. Nonetheless, there is a general agreement in the mea-
surement of ‘‘threshold pore diameter” when interpreting MIP test 0
4000 3600 3200 2800 2400 2000 1600 1200 800 400
results [26]. The threshold pore diameter is the first connected
Wavenumber (cm-1)
pore pathway in materials, and is defined as the inflection point
on the cumulative intrusion versus pore diameter curve [26]. As Fig. 6. FT-IR spectra of raw fly ashes and geopolymer samples at 3 and 91 days.
J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269 265

Table 5 position of the asymmetric Si–O–T stretching vibration at the ini-


Summary of the major IR vibrational bands observed in this study. tial stage shifting to lower frequency indicates the formation of
Wavenumber (cm1) Type of vibration Refs. Al-rich geopolymer gel, while the position of the asymmetric Si–
3585–3448 Stretching (–OH) [23,44,50] O–T stretching vibration shifting to higher frequency at a later
1654–1651 Bending (H–O–H) [23,44,50] stage indicates the gel transformation into Si-rich geopolymer gel
1455–1440 Stretching (O–C–O) [49,50] [37,52,54].
1085–1009 Asymmetric stretching (Si–O–T: [43,48]
T = tetrahedral Si or Al)
797–776 Symmetric stretching (Si–O–Si) [43,47]
562–550 Symmetric stretching (Al–O–Si) [43,46] 3.5. Morphology change in geopolymer
469–457 Bending (Si–O–Si and O–Si–O) [43,45–48]
The morphology of geopolymer samples observed by SEM at 3
and 91 days is shown in Fig. 8. At 3 days, the geopolymer consisted
of a weakly bonded matrix in which colloidal aluminosilicate was
loosely distributed, and a crystalline phase was not visible in GP1
1100
GP1 GP2 GP3
and GP2. However, the presence of crystalline phases and
amorphous phases surrounding a fly ash particle was observed
1080
from the GP3 sample at 3 days (see Fig. 9). The geopolymer matrix
Wavenumber (cm-1)

at 91 days had more densely distributed amorphous phases com-


1060 pared with that at 3 days and crystalline phases were visible in
all samples. In particular, it could be observed from the GP2 sample
1040 at 91 days that amorphous phases were present within or outside a
fly ash particle. In the GP3 sample at 91 days, very interesting mor-
1020 phology can be observed: a large amount of needle-shaped crys-
talline phases have grown within a fly ash particle and
1000
penetrated through the glass layer at the surface of the fly ash. In
Raw fly ash GP-3d GP-91d addition, polyhedral crystalline phases were observed from the
GP3 sample at 91 days. The EDS analysis results for the needle-
Fig. 7. Position shifts over time of the main Si–O–T asymmetric stretching band for
shaped crystal and polyhedral crystal are provided in Table 6.
GP1, GP2 and GP3.
The needle-shaped crystal is rich in Al and Si and has a low Na con-
tent. The formation of a new geopolymerization product, a crystal
tional bands observed in this study were summarized in Table 5 by similar to mullite (Al6Si2O13), is evidenced by its atomic ratio and
referring to previous studies [23,43–50]. The spectra of raw fly crystal shape. On the other hand, the polyhedral crystal contains
ashes display a main broad band attributed to asymmetric stretch- a high content of Fe, Mg, Al, and a low content of Si and Na. The
ing of Si–O–T (T = tetrahedral Si or Al) around 1085 cm1 and atomic ratio and crystal shape signify the occurrence of crystalliza-
another bands, around 790, 560 and 470 cm1, can be attributed tion similar to osumilite ((K,Na)(Fe,Mg)2(Al,Fe)3(Si,Al)12O30H2O).
to symmetric stretching of Si–O–Si, symmetric stretching of Al–
O–Si and the bending vibration of Si–O, respectively. These bands
can be considered evidence of mullite and quartz contained in 4. Discussion
the raw fly ash [43,51,52]. It should be noted that the intensity
of the band is not related to the crystallization of the raw fly ashes 4.1. Delayed strength development of geopolymer
[52]. On the other hand, a distinct difference was observed in the
FT-IR spectra of the geopolymer and raw fly ash. In all geopolymer It can be reasonably assumed that the graph plotted below GP1
samples, new bands appeared in the regions of 3585–3448 and in Fig. 4 represents relatively delayed strength development. The
1654–1651 cm1 which were attributed to the stretching vibration delayed strength development of geopolymers reported in previ-
of –OH and the bending vibration of H–O–H, respectively ous studies have been attributed to the curing temperature
[23,44,50]. These new bands indicate the presence of OH com- [3,20], the particle size of fly ash [20], the concentration of alkali-
bined due to geopolymerization and water adsorbed in the rings activator [23,30] and the composition of fly ash [37] (this may
of geopolymeric products [23,44,50,53]. Moreover, another new include SiO2/Al2O3 ratio, vitreous phase content, alumina release
band at 1455–1440 cm1, which was attributed to stretching rate, etc.). In the present study, the curing temperature, concentra-
vibration of O–C–O, indicates the presence of sodium bicarbonate tion of alkali-activator and fly ash-to-activator ratio were kept
[49,50]. identical for all samples, while the three types of fly ash had
The shift in the position of the main band around 1085 cm1, physicochemical characteristics different from each other. The
which was attributed to the asymmetric stretching of Si–O–T is results presented in this study show that the characteristics of fly
often referenced as an explanation of a geopolymer reaction pro- ash can significantly affect the strength development of the
cess [9,23,37,43,50,52,54]. The position shifts over time of the main geopolymer.
Si–O–T asymmetric stretching band for GP1, GP2 and GP3 are In particular, the results provide an insight into the delayed
shown in Fig. 7. The main band of the three types of raw fly ash strength development of geopolymer which mainly occurs due to
identically coincide at the frequency of 1086 cm1, while position the transformation of aluminosilicate gel from Al-rich gel to Si-
shifts of the frequency as the age of the geopolymer proceeded rich gel, rather than densification of pore structures. The MIP
were different for GP1, GP2 and GP3. The main band of GP1 shifted results shown in Fig. 5 reveal that the increase in age (i.e., contin-
to lower frequency as the time proceeded, while that of GP2 and ued geopolymerization) had no correlation with the decrease in
GP3 shifted to the lowest frequency at 3 days and then shifted to porosity. However, an increase was noticed in the median pore
slightly higher frequency at 91 days. In addition, the presence of diameter and porosity of GP2 and GP3, which had relatively
two slopes in the main band observed in GP3 at 91 days. The delayed strength development. This may have been caused
observed band shift of GP2 and GP3 can be explained as the change because a portion of fly ash which was not involved in the reaction
in Al content of aluminosilicate gel over the reaction time. The during the formation of the geopolymer matrix, became involved
266 J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269

GP1 – 3 d GP1 – 91 d

GP2 – 3 d GP2 – 91 d

GP3 – 3 d GP3 – 91 d

Fig. 8. Morphology of geopolymer samples observed by SEM at 3 and 91 days.

Table 6
GP3 – 3 d Elemental compositions of the needle-shaped crystal and the polyhedral crystal
observed in GP3 at 91 days.

at% Needle-shaped crystal Polyhedral crystal


O 56.66 62.13
Fe – 10.25
Na 3.07 2.37
Mg 0.78 12.58
Al 23.56 9.64
Si 18.92 3.03

in the reaction at a later age, breaking the glass layer surrounding


the fly ash particles and resulting in an increase in the porosity.
GP3 – 91 d On the other hand, the FT-IR spectra in Figs. 6 and 7 clearly rep-
resent the behavior of the aluminosilicate gel as the time pro-
gressed. Previous studies on the FT-IR spectra of geopolymers
found that the shift behavior exhibited by the main band of GP1
provided an evidence for the formation of an Al-rich aluminosili-
cate gel (e.g., –Si–O–Al–O–), and the presence of this type of gel
even at later age [37,50,52,54]. In contrast to GP1, the shift behav-
ior exhibited by the main band of GP3 shows that while Al-rich
aluminosilicate gel formed at early age, it underwent gel transfor-
mation into the Si-rich aluminosilicate gel (e.g., –Si–O–Al–O–Si–O–
Si–O–) [37,50,52,54]. Fernández-Jiménez et al. conducted a study
on the transformation of aluminosilicate gel and reported that
Fig. 9. Crystalline phases observed in GP3 at 3 and 91 days. although the formation of Al-rich aluminosilicate gel (i.e., low
J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269 267

70 55 55

Relative compressive strength at 7d (%)


(a) (b) (c)

Ultimate compressive strength (MPa)


Ultimate compressive strength (MPa)
GP1 GP3 GP3
60
50 50
50 GP2 GP2
GP2
40 45 45

GP3 GP1
30 GP1
40 40

20

35 35
10

0 30 30
5 10 15 20 1 2 3 70 75 80 85 90
Mean particle size of fly ash ( m) SiO2/Al2O3 ratio of fly ash Vitreous phase of fly ash (%)

Fig. 10. The effect of physicochemical characteristics of fly ash on strength development of geopolymer.

[Q4(3Al) + Q4(2Al)]/Q4(4Al) ratios) contributed toward the early shown in an SEM micrograph of GP3 at 3 days in Fig. 9. This could
age strength development, the transformation into Si-rich alumi- degrade the dissolution of fly ash into liquid and furthermore,
nosilicate gel (i.e., high [Q4(3Al) + Q4(2Al)]/Q4(4Al) ratios) had a could interfere with the strength development at early stages
more critical influence on the further increase in the compressive [23]. Meanwhile, an SEM micrography of GP3 at 91 days shows
strength at later age [55]. Generally, the bond strength of Si–O–Si the growth of a new needle-shaped crystal with high Al and Si con-
bonds is higher than that of Al–O–Al and Si–O–Al bonds [29]. Sim- tent further progressed. This crystal growth can be thought to have
ilarly, the strength of the tetrahedral network in an aluminosilicate a positive effect in the strength development at later age, similar to
gel depends on and increases with the Si content [29]. Thus, the the case of ettringite formation in Portland cement based materi-
delayed high-strength development exhibited by GP3 at later age als. In the case of Portland cement, the delayed ettringite formation
was driven by the gel transformation into Si-rich aluminosilicate may induce cracks, leading to the degradation of the cementitious
gel, rather than a change in porosity due to the additional produc- material (known as sulfate attack), but it is occasionally accepted
tion of reactants. that the ettringite formation could contribute to enhancement in
the strength [58]. Calcium sulfoaluminate cement is an example
4.2. Effect of fly ash characteristics on strength development where the ettringite formation was utilized to develop a high
strength cementitious material.
It is obvious that the fly ashes used in this study had different For similar reason, the needle-shaped crystal growth at later
physicochemical characteristics and thus resulted in different reac- age as observed in this study might have affected the increase in
tion mechanisms at early and later ages. The effect of physico- compressive strength of the geopolymer to some extent. However,
chemical characteristics of fly ash on strength development of the effect of the crystallinity of the geopolymerization reactants on
geopolymer is shown in Fig. 10. It can be concluded from the the strength development has not been quantitatively revealed.
experimental results that the strength development of geopolymer Previous studies noted that a thorough consideration is required
was delayed as the particle size of the fly ash increased, i.e., the to identify the effect of nanometer-scale crystallinity on the
surface area decreased. In particular, the delayed high-strength strength of the geopolymer, since its size is too small to act as an
development was more evident as the vitreous phase content aggregate within the matrix [59,60]. Thus, the findings of this
and SiO2/Al2O3 ratio of the fly ash increased. FA1 had the finest par- study in this respect remain as a probable hypothesis which needs
ticle, the lowest SiO2/Al2O3 ratio and the lowest vitreous phase further investigation.
content, whereas FA3 had the coarsest particle, the highest SiO2/
Al2O3 ratio and the highest vitreous phase content. Since particle 5. Concluding remarks
size is one of the factors which contribute to the granular reaction
rate [32], it may have been responsible for the high reactivity This study investigated the effect of fly ash characteristics on
exhibited by FA1 and the rather low reactivity exhibited by FA3 the strength development of fly ash-based geopolymer. Three
at early age. It is known that a compound between silicates in an types of class F fly ash activated with 9 M NaOH solution and
alkaline aluminosilicate solution takes longer to form than one sodium silicate were used for experiments, and multi-technical
between aluminate and silicate does [56,57], thus a compound characterizations using MIP, FT-IR spectroscopy and SEM & EDS
might have been formed between a silicate and aluminate at high were applied to obtain in-depth understanding of the relationship
reaction degree at early age in GP1. Although GP3, which was made between the reaction products, microstructure, and strength
with coarse FA3, showed rather slower production of Al-rich alu- development with the maturity of the geopolymer. The main con-
minosilicate gel compared to GP1 and GP2, its high SiO2/Al2O3 ratio clusions extracted from this work are as follows.
and vitreous phase content may have been responsible for the
transformation into Si-rich aluminosilicate gel at later age. Further (1) The characteristics of the fly ash can significantly affect the
information about the release rate and reaction kinetics of alumi- properties of geopolymer.
nate and silicate can be found in [19,29,30,37]. (2) The delayed strength development of geopolymer occurred
mainly due to the transformation of aluminosilicate gel into
4.3. Effect of crystal growth on strength development the Si-rich aluminosilicate gel, rather than the densification
of pore structures.
The effect of crystal growth on the strength development can be (3) The particle size of fly ash had a crucial influence on the
discussed from the results of SEM (see Figs. 8 and 9). New crys- strength development. As the particle size of fly ash
talline products were observed surrounding fly ash particles, as increased, i.e., as the surface area decreased, the strength
268 J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269

development of geopolymer was delayed. In particular, the [24] American Society for Testing and Materials, ASTM C618: Standard
Specification for Coal Fly Ash and Raw or Calcined Natural Pozzolan for Use
delay in high-strength development was more evident as
in Concrete, ASTM International, 2012.
the vitreous phase content and the SiO2/Al2O3 ratio of fly [25] American Society for Testing and Materials, ASTM C109: Standard Test
ash increased. Method for Compressive Strength of Hydraulic Cement Mortars, ASTM
(4) The effect of a new needle-shaped crystal growth on the International, 2013.
[26] S. Diamond, Mercury porosimetry: an inappropriate method for the
strength development at later age, similar to the ettringite measurement of pore size distributions in cement-based materials, Cem.
formation of Portland cement based materials, was noticed. Concr. Res. 30 (10) (2000) 1517–1525.
[27] E. Kamseu, B. Nait-Ali, M.C. Bignozzi, C. Leonelli, S. Rossignol, D.S. Smith, Bulk
composition and microstructure dependence of effective thermal conductivity
of porous inorganic polymer cements, J. Eur. Ceram. Soc. 32 (8) (2012) 1593–
Acknowledgments 1603.
[28] Y. Ma, J. Hu, G. Ye, The pore structure and permeability of alkali activated fly
ash, Fuel 104 (2013) 771–780.
This research was supported by a Grant from the Energy Tech- [29] P. Duxson, J.L. Provis, G.C. Lukey, S.W. Mallicoat, W.M. Kriven, J.S.J. van
nology Development Program (Grant No. 2013T100100021) Deventer, Understanding the relationship between geopolymer composition,
funded by the Ministry of Trade Industrial and Energy of the Kor- microstructure and mechanical properties, Colloid Surf. A 269 (1–3) (2005)
47–58.
ean government. The authors are grateful to Dr. Seung-Muk Bae [30] M. Criado, A. Fernández-Jiménez, A.G. de la Torre, M.A.G. Aranda, A. Palomo, An
from Center for research facilities in Kunsan National University, XRD study of the effect of the SiO2/Na2O ratio on the alkali activation of fly ash,
Mr. Sol-Moi Park and Ms. Hee-Jeong Kim from Structural Analysis Cem. Concr. Res. 37 (5) (2007) 671–679.
[31] E. Sakai, S. Miyahara, S. Ohsawa, S.H. Lee, M. Daimon, Hydration of fly ash
& Materials Research Laboratory in KAIST for the cooperation in the cement, Cem. Concr. Res. 35 (6) (2005) 1135–1140.
test involved in this study. [32] J.G.S. van Jaarsveld, J.S.J. van Deventer, G.C. Lukey, The characterisation of
source materials in fly ash-based geopolymers, Mater. Lett. 57 (7) (2003)
1272–1280.
References [33] U. Rattanasak, P. Chindaprasirt, Influence of NaOH solution on the synthesis of
fly ash geopolymer, Miner. Eng. 22 (12) (2009) 1073–1078.
[1] M. Ahmaruzzaman, A review on the utilization of fly ash, Prog. Energy [34] J. Temuujin, R.P. Williams, A. van Riessen, Effect of mechanical activation of fly
Combust. 36 (3) (2010) 327–363. ash on the properties of geopolymer cured at ambient temperature, J. Mater.
[2] R. Siddique, M.I. Khan, Supplementary Cementing Materials, Springer, 2011. Process. Technol. 209 (12–13) (2009) 5276–5280.
[3] A. Palomo, S. Alonso, A. Fernández-Jiménez, Alkaline activation of fly ashes: [35] S. Andini, R. Cioffi, F. Colangelo, T. Grieco, F. Montagnaro, L. Santoro, Coal fly
NMR study of the reaction products, J. Am. Ceram. Soc. 87 (6) (2004) 1141– ash as raw material for the manufacture of geopolymer-based products, Waste
1145. Manage. 28 (2) (2008) 416–423.
[4] J. Davidovits, Geopolymers, J. Therm. Anal. Calorim. 37 (8) (1991) 1633–1656. [36] R.N. Thakur, S. Ghosh, Effect of mix composition on compressive strength and
[5] P. Duxson, A. Fernández-Jiménez, J.L. Provis, G.C. Lukey, A. Palomo, J.S.J. van microstructure of fly ash based geopolymer composites, ARPN J. Eng. Appl. Sci.
Deventer, Geopolymer technology: the current state of the art, J. Mater. Sci. 42 4 (4) (2009) 68–74.
(9) (2007) 2917–2933. [37] A. Hajimohammadi, J.L. Provis, J.S.J. van Deventer, Effect of alumina release
[6] D. Khale, R. Chaudhary, Mechanism of geopolymerization and factors rate on the mechanism of geopolymer gel formation, Chem. Mater. 22 (2010)
influencing its development: a review, J. Mater. Sci. 42 (2007) 729–746. 5199–5205.
[7] J.L. Provis, Geopolymers and other alkali activated materials: why, how, and [38] R.A. Cook, K.C. Hover, Mercury porosimetry of hardened cement pastes, Cem.
what?, Mater Struct. 47 (2014) 11–25. Concr. Res. 29 (6) (1999) 933–943.
[8] J.G. Jang, Y.B. Ahn, H. Souri, H.K. Lee, A novel eco-friendly porous concrete [39] D. Winslow, D. Liu, The pore structure of paste in concrete, Cem. Concr. Res. 20
fabricated with coal ash and geopolymeric binder: heavy metal leaching (2) (1990) 227–235.
characteristics and compressive strength, Constr. Build. Mater. 79 (2015) 173– [40] E.P. Kearsley, P.J. Wainwright, The effect of porosity on the strength of foamed
181. concrete, Cem. Concr. Res. 32 (2) (2002) 233–239.
[9] A. Palomo, M.W. Grutzeck, M.T. Blanco, Alkali-activated fly ashes: a cement for [41] R. Kumar, B. Bhattacharjee, Porosity, pore size distribution and in situ strength
the future, Cem. Concr. Res. 29 (8) (1999) 1323–1329. of concrete, Cem. Concr. Res. 33 (1) (2003) 155–164.
[10] V.F.F. Barbosa, K.J.D. MacKenzie, Thermal behaviour of inorganic geopolymers [42] J.S.J. Sindhunata van Deventer, GC.. Lukey, H. Xu, Effect of curing temperature
and composites derived from sodium polysialate, Mater. Res. Bull. 38 (2) and silicate concentration on fly-ash-based geopolymerization, Ind. Eng.
(2003) 319–331. Chem. Res. 45 (2006) 3559–3568.
[11] N.K. Lee, J.G. Jang, H.K. Lee, Shrinkage characteristics of alkali-activated fly ash/ [43] W.K.W. Lee, J.S.J. van Deventer, Use of infrared spectroscopy to study
slag paste and mortar at early ages, Cem. Concr. Compos. 53 (2014) 239–248. geopolymerization of heterogeneous amorphous aluminosilicates, Langmuir
[12] W.K.W. Lee, J.S.J. van Deventer, Effects of anions on the formation of 19 (2003) 8726–8734.
aluminosilicate gel in geopolymers, Ind. Eng. Chem. Res. 41 (18) (2002) [44] X. Guo, H. Shi, W.A. Dick, Compressive strength and microstructural
4550–4558. characteristics of class C fly ash geopolymer, Cem. Concr. Compos. 32 (2)
[13] J.G.S. van Jaarsveld, J.S.J. van Deventer, G.C. Lukey, The effect of composition (2010) 142–147.
and temperature on the properties of fly ash- and kaolinite-based [45] M. Handke, W. Mozgawa, M. Nocuń, Specific features of the IR spectra of
geopolymers, Chem. Eng. J. 89 (1–3) (2002) 63–73. silicate glasses, J. Mol. Struct. 325 (1994) 129–136.
[14] J.G. Jang, N.K. Lee, H.K. Lee, Fresh and hardened properties of alkali-activated [46] B.T. Poe, P.F. McMillan, C.A. Angell, R.K. Sato, Al and Si coordination in SiO2–
fly ash/slag pastes with superplasticizers, Constr. Build. Mater. 50 (2014) 169– Al2O3 glasses and liquids: a study by NMR and IR spectroscopy and MD
176. simulations, Chem. Geol. 96 (3–4) (1992) 333–349.
[15] N.K. Lee, H.K. Lee, Setting and mechanical properties of alkali-activated fly ash/ [47] M.Y.A. Mollah, T.R. Hess, D.L. Cocke, Surface and bulk studies of leached and
slag concrete manufactured at room temperature, Constr. Build. Mater. 47 unleached fly ash using XPS, SEM, EDS and FTIR techniques, Cem. Concr. Res.
(2013) 1201–1209. 24 (1) (1994) 109–118.
[16] M. Steveson, K. Sagoe-Crentsil, Relationships between composition, structure [48] R.K. Vempati, A. Rao, T.R. Hess, D.L. Cocke, H.V. Lauer Jr, Fractionation and
and strength of inorganic polymers, J. Mater. Sci. 40 (2005) 4247–4259. characterization of Texas lignite class ‘F’ fly ash by XRD, TGA, FTIR, and SFM,
[17] A. Fernández-Jiménez, A. Palomo, Characterisation of fly ashes. Potential Cem. Concr. Res. 24 (6) (1994) 1153–1164.
reactivity as alkaline cements, Fuel 82 (18) (2003) 2259–2265. [49] J.C. Swanepoel, C.A. Strydom, Utilisation of fly ash in a geopolymeric material,
[18] A. Fernández-Jiménez, A.G. de la Torre, A. Palomo, G. López-Olmo, M.M. Appl. Geochem. 17 (8) (2002) 1143–1148.
Alonso, M.A.G. Aranda, Quantitative determination of phases in the alkali [50] W.K.W. Lee, J.S.J. van Deventer, The effects of inorganic salt contamination on
activation of fly ash. Part I. Potential ash reactivity, Fuel 85 (5–6) (2006) 625– the strength and durability of geopolymers, Colloid Surf. A 211 (2–3) (2002)
634. 115–126.
[19] P.D. Silva, K. Sagoe-Crenstil, V. Sirivivatnanon, Kinetics of geopolymerization: [51] A. Fernández-Jiménez, M. Monzó, M. Vicent, A. Barba, A. Palomo, Alkaline
role of Al2O3 and SiO2, Cem. Concr. Res. 37 (4) (2007) 512–518. activation of metakaolin–fly ash mixtures: obtain of zeoceramics and
[20] K. Somna, C. Jaturapitakkul, P. Kajitvichyanukul, P. Chindaprasirt, NaOH- zeocements, Microporous Mesoporous Mater. 108 (1–3) (2008) 41–49.
activated ground fly ash geopolymer cured at ambient temperature, Fuel 90 [52] A. Fernández-Jiménez, A. Palomo, Mid-infrared spectroscopic studies of alkali-
(6) (2011) 2118–2124. activated fly ash structure, Microporous Mesoporous Mater. 86 (1–3) (2005)
[21] G. Kovalchuk, A. Fernández-Jiménez, A. Palomo, Alkali-activated fly ash: effect 207–214.
of thermal curing conditions on mechanical and microstructural development. [53] A. Palomo, M.T. Blanco-Varela, M.L. Granizo, F. Puertas, T. Vazquez, M.W.
Part II, Fuel 86 (3) (2007) 315–322. Grutzeck, Chemical stability of cementitious materials based on metakaolin,
[22] P. Chindaprasirt, T. Chareerat, S. Hatanaka, T. Cao, High-strength geopolymer Cem. Concr. Res. 29 (7) (1999) 997–1004.
using fine high-calcium fly ash, J. Mater. Civil. Eng. 23 (3) (2011) 264–270. [54] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Attenuated total reflectance
[23] T. Bakharev, Geopolymeric materials prepared using Class F fly ash and fourier transform infrared analysis of fly ash geopolymer gel aging, Langmuir
elevated temperature curing, Cem. Concr. Res. 35 (6) (2005) 1224–1232. 23 (2007) 8170–8179.
J.G. Jang, H.K. Lee / Construction and Building Materials 102 (2016) 260–269 269

[55] A. Fernández-Jiménez, A.G. de la Torre, A. Palomo, G. López-Olmo, M.M. [58] P.K. Mehta, Mechanism of sulfate attack on Portland cement concrete –
Alonso, M.A.G. Aranda, Quantitative determination of phases in the alkaline another look, Cem. Concr. Res. 13 (3) (1983) 401–406.
activation of fly ash. Part II. Degree of reaction, Fuel 85 (14–15) (2006) 1960– [59] J.W. Phair, J.S.J. van Deventer, J.D. Smith, Interaction of sodium silicate with
1969. zirconia and its consequences for polysialylation, Colloid Surf. A 182 (1–3)
[56] M.R. Anseau, J.P. Leung, N. Sahai, T.W. Swaddle, Interactions of silicate ions (2001) 143–159.
with zinc (II) and aluminium (III) in alkali aqueous solution, Inorg. Chem. 44 [60] J.L. Provis, G.C. Lukey, J.S.J. van Deventer, Do geopolymers actually contain
(22) (2005) 8023–8032. nanocrystalline zeolites? A reexamination of existing results, Chem. Mater. 17
[57] M.R. North, T.W. Swaddle, Kinetics of silicate exchange in alkaline (2005) 3075–3085.
aluminosilicate solutions, Inorg. Chem. 39 (12) (2000) 2661–2665.

You might also like