You are on page 1of 10

Construction and Building Materials 81 (2015) 303–312

Contents lists available at ScienceDirect

Construction and Building Materials


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

Reactivity and reaction products of alkali-activated, fly ash/slag paste


N.K. Lee, H.K. Lee ⇑
Department of Civil and Environmental Engineering, Korea Advanced Institute of Science and Technology, Guseong-dong, Yuseong-gu, Daejeon 305-701, South Korea

h i g h l i g h t s

 The amount of added slag affected the formation of reaction product.


 As the amount of added slag increased, the amount of C-S-H gel increased.
 The aluminosilicate gel with the Q4 units was similar to a Ca-based geopolymer.

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

Article history: Few studies have described the reaction products of an alkali-activated, two-source binder and their
Received 23 August 2014 characteristics due to the complexity of the mechanism. In this study, the microstructure, reaction prod-
Received in revised form 14 January 2015 ucts, and reactivity of alkali-activated, fly ash/slag binders synthesized at various mixture ratios of two
Accepted 18 February 2015
raw materials were examined using various experimental techniques (NMR, ICP-OES, EDS, FT-IR and
Available online 4 March 2015
TGA) to systematically investigate the complex reaction mechanism of the binders. It was also intended
to help assess durability of the binders. It was found that the amount of added slag primarily affected the
Keywords:
amount of reaction product and its silicate structure, and as the amount of added slag increased, the
Alkali activated fly ash/slag
Reactivity
amount of C-S-H gel increased and the amount of aluminosilicate gel decreased. Considering chemical
Reaction product composition and silicate structure, the aluminosilicate gel with the Q4 (nAl) units was found to be similar
Ca-based geopolymer to a Ca-based geopolymer (N-C-A-S-H). The chemical composition ratios of the Ca-based geopolymer
N-C-A-S-H were nearly the same as those of C-S-H whereas their silicate structures were different.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction one-source mixture was effective for improving the mechanical


properties of the geopolymers made with metakaolin [43,44] and
Geopolymer materials are known as alternatives to ordinary fly ash [23]. The authors investigated mechanical properties such
Portland cement (OPC); the geopolymer binders can reduce CO2 as compressive strength, elastic modulus, splitting tensile strength,
emissions by 80% compared to the OPC [11]. The main sources of and setting properties of alkali-activated, fly ash/slag concrete at
the geopolymers are metakaolin, fly ash, and blast furnace slag. various slag contents, and concluded that the slag content of 20–
Reaction products formed through an alkali-activation process 30% relative to the total binder weight was optimal considering
are completely different depending on chemical composition of setting, flowability, and strength [27,28,20].
the raw sources [24]. The main reaction product of alkali-activated There are two different reaction products of the alkali-activated,
blast furnace slag, which is rich in Ca, is C-S-H gel with a low Ca/Si two-source binder; both products had higher reactivity than that
ratio (0.8–1.1) [24,16,17], whereas the main reaction product of of the one-source binder. Puertas et al. [37] showed that two differ-
alkali-activated metakaolin or fly ash, which is rich in Si and Al, ent reaction products existed in alkali-activated pastes. Corre-
is amorphous aluminosilicate gel with a three-dimensional frame- spondingly, Buchwald et al. [6] showed in an NMR study that
work of SiO4 and AlO4 tetrahedra linked through shared O atoms alkali-activated metakaolin/slag binders had higher reactivity than
[35,41,10]. the alkali-activated metakaolin binder, and established the types
Several studies have reported the alkali-activation of two- and quantities of the reaction products. The main reaction prod-
source mixtures that consist of fly ash/slag or metakaolin/slag ucts were calcium silicate hydrate rich in Al that contains Na in
[37]; [47,31,3,38,5,19]. The addition of blast furnace slag in the its structure and alkaline aluminosilicate hydrate with a three-
dimensional structure [37]. Yip and Van Deventer [45] and
Buchwald et al. [6] concluded, after SEM observations and NMR
⇑ Corresponding author. Tel.: +82 42 350 3623; fax: +82 42 350 3610.
spectroscopic studies, that the C-S-H and aluminosilicate gel
E-mail address: leeh@kaist.ac.kr (H.K. Lee).

http://dx.doi.org/10.1016/j.conbuildmat.2015.02.022
0950-0618/Ó 2015 Elsevier Ltd. All rights reserved.
304 N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312

coexist in the alkali-activated metakaolin/slag binders. In addition, Table 2


the chemical interaction may arise between the elements released Mix proportion of alkali-activated fly ash/slag paste.

by the dissolution of fly ash and slag particles [25], and the two Sample Water/ SiO2/Na2O ratio of Slag/(fly Type of
phases (C-S-H and aluminosilicate) may not only be combined bindera sodium silicate ash + slag)a sodium
but also may chemically interact [6,25]. On the contrary, the silicate

authors stated that the discrete formation of C-S-H gel and PS10 0.4 2.12 0.1 Powder
aluminosilicate gel in an SEM length scale was not observed from PS30 0.4 2.12 0.3 Powder
PS50 0.4 2.12 0.5 Powder
their experiment, and the separation between calcium-rich and LS10 0.4 0.94 0.1 Liquid
calcium-deficient regions was not identified [26]. Nevertheless, if LS30 0.4 0.94 0.3 Liquid
the separation is observed, it may be due to inadequate mixing LS50 0.4 0.94 0.5 Liquid
during the manufacturing process [26]. Recently, it was found that a
All values are given as mass ratios.
C-A-S-H gel and N-A-S-H gel are reaction products in a fly ash/slag
geopolymer [19,5,29], and the N-C-A-S-H gel may be a hybrid-type
phase of the N-A-S-H gel and the C-A-S-H gel [19]. use of liquid sodium silicate, fly ash, and slag were dry-mixed and then the liquid
The alkali-activation of a two-source mixture has a more com- sodium silicate was added to the mixture. The AFS paste samples prepared by the
process were immediately cast into 50 mm cubic molds. All of the samples were
plex mechanism than that of a one-source mixture. Few studies
cured at 20 °C and at a relative humidity of 50% in a room with constant tem-
have described the reaction products of an alkali-activated, two- perature and humidity levels. After one day, the samples were removed from their
source binder and its characteristics due to the complexity of the molds and were stored in a conditioning room until the day of testing.
mechanism [5,38,19]. Quantitatively evaluating the reaction prod-
ucts and reactivity of alkali-activated, fly ash/slag binder synthe- 2.3. Experimental details
sized at various mixture ratios of two raw materials is important
to figure out the complex mechanism. The AFS samples for the X-ray powder diffraction (XRD) test were prepared by
mechanical grinding. The XRD data were recorded on a Rigaku D/MAX-2500
In the present study, the microstructure, reaction products, and
machine using Cu Ka radiation at a scanning rate of 2°/min from 2° to 160° in 2h.
reactivity of alkali-activated, fly ash/slag binders synthesized at To analyze the polished surfaces of the samples on a NOVA 230 device (FEI
various mixture ratios of two raw materials were examined using Company), scanning electron microscopy (SEM) with back-scattered electron
various experimental techniques (NMR, ICP-OES, EDS, FT-IR and (BSE) images and energy-dispersive spectroscopy (EDS) analyses were performed.
TGA) in order to systematically investigate the complex reaction The samples were impregnated by using low-viscosity epoxy resin, polished with
SiC paper, and coated with gold.
mechanism of the binders. The test results will help assess the Powdered AFS samples were dried at 80 °C in an oven for 1 day before the ther-
durability of the alkali-activated, fly ash/slag binders with different mogravimetric analyses (TGA). The TGA measurements were applied to the pow-
amounts of the reaction products and different reactivity levels. dered samples under N2 gas at 10 K/min up to 900 °C.
Inductively coupled plasma optical emission spectroscopy (ICP-OES) was used
to determine the water-soluble silicon, aluminum, calcium, and sodium contents
2. Experimental program of the AFS samples in distilled water. The powdered AFS samples were sonicated
in distilled water for more than 2 h and the amounts of water-soluble elements
2.1. Materials were then measured by means of ICP-OES. The reactivity of the alkali-activator
(sodium silicate powder or liquid) was evaluated indirectly through this quantita-
Class F fly ash (containing more than 70% pozzolanic compounds (SiO2, Al2O3, tive analysis.
and Fe2O3) in accordance with ASTM C618) and blast furnace slag were used as bin- 29
Si nuclear magnetic resonance (NMR) spectroscopy (Bruker AVANCE 400WB)
der materials. The chemical compositions of these materials are listed in Table 1. was used to record the solid-state NMR spectra with the purpose of evaluating the
Two types of alkali-activators were prepared: sodium silicate powder with a reactivity and content of the reaction products (i.e., C-S-H gel and aluminosilicate
chemical composition of SiO2 (53.4%) and Al2O3 (25.2%), a bulk density of 0.62 g/ gel. The 29Si resonance frequency was 79.42 MHz and the spinning rate was
3
cm and a molar ratio of 2.18 (Ms = SiO2/Na2O); and sodium silicate liquid 5 kHz. All of the measurements were taken at room temperature with tetramethyl-
(SiO2/Na2O = 1.0) mixed with 4 M NaOH and water glass (Korean industrial silane (TMS) as an external standard. The spectra were acquired using a pulse
standards (KS) Grade-3; SiO2 (29%), Na2O (10%), H2O (61%), specific gravity length of 1.5 ls, and a short repetition time of 20 s was chosen. The powdered sam-
1.38 g/mL). Distilled water was used to dissolve the solid NaOH (98%); the latter ples were also analyzed by Fourier transform infrared (FT-IR) spectroscopy (Model
alkali activator was prepared by mixing the NaOH solution with water glass at FT-IR 4100, JASCO, Japan).
the SiO2/Na2O ratios developed by the authors [27].

2.2. Mixture proportions 3. Results

The mixture proportions are provided in Table 2 and are labeled with specific 3.1. X-ray powder diffraction (XRD)
codes. The labels ‘P’, ‘L’ and ‘S’ represent powder, liquid, and slag, respectively.
The numbers ‘10’, ‘30’, and ‘50’ refer to the percentage of slag to the total binder
(fly ash + slag) by mass.
Fig. 1 shows the XRD patterns of the AFS samples, raw fly ash
The method for producing the alkali-activated, fly ash/slag (AFS) paste is as fol- and raw slag. The raw fly ash shows peaks related to the presence
lows: In the case of use of powder sodium silicate, fly ash, slag, and powder sodium
silicate were dry-mixed for 2 min to ensure the homogeneity of the mixture. Once
water was added to the mixture, it was mixed for an additional 2 min. In the case of Q C-S-H, Calcite

M M Q
M
Table 1
Slag
Intensity

Chemical composition of the binder materials.

Oxide (wt.%) Fly ash (FA) Blast furnace slag (BFS) LS50

CaO 4.41 42.47 LS30


SiO2 67.26 35.17
Al2O3 14.76 13.93 LS10
Fe2O3 4.07 0.58 Fly ash
SO3 – 2.03 0 10 20 30 40 50 60 70
MgO 1.29 4.12
Two- Theta Angle (deg.)
Na2O 2.04 0.15
K2O 1.39 0.46
Fig. 1. XRD patterns of raw fly ash, raw slag and AFS samples activated by sodium
LOI 3.57 0.18
silicate liquid.
N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312 305

of quartz and mullite, and the raw slag clearly shows a diffuse band of the LS50 sample. In the LS samples, a peak at approximately
at approximately 30° in 2h due to its amorphous nature. 450–500 °C related to the presence of Ca(OH)2 was not identified,
The LS sample with the liquid activator shows a strong peak at which is consistent with the results of the XRD analysis, as shown
approximately 29–30° in 2h regardless of the amount of slag in the in Fig. 2(a).
mixture. It is similar to the peaks related to the presence of C-S-H Fig. 2(b) shows the results of the DTG analysis for the LS and PS
or calcite, as observed in the previous study [29] using the pow- samples. The maximum weight loss peaks for the two types of
dered alkali-activator. The peaks corresponding to mullite and samples are slightly different (100 °C for PS30 and 150 °C for
quartz were observed due to the presence of unreacted fly ash in LS30, respectively). Very weak peaks were observed between 500
the LS samples. As the amount of added slag increased, the width and 700 °C in the LS samples, while significant peaks were
of the peak at 29–30° in 2h became larger, meaning that the LS observed in the PS samples due to the formation of silica gel from
sample is amorphous. The XRD patterns were slightly different at the decomposition of carbonate (564–658 °C) [39,33] or the pres-
20–30° in 2h between the LS10 (i.e., sample with 10% of slag) ence of unreacted silica gel in the sodium silicate used as the pow-
and the LS30 or LS50 (sample with 30% or 50% of slag) samples. der activator [39,40,36]. In previous studies [2,4], the DTG results
The LS10 sample showed a larger diffuse peak at 20–25° in 2h as for metakaolin/slag blends showed that the intensity of the peaks
compared to the LS30 and LS50 samples due to the formation of at temperatures between 590 and 690 °C increased as the amount
aluminosilicate gel or silica as a reaction product [42,18]. The exis- of added slag increased due to the presence of calcium carbonate.
tence of the aluminosilicate gel will be explained later in this Thus, the weight loss peak between 500 and 700 °C may be due to
study. the formation of silica gel from the decomposition of carbonate
(564–658 °C). However, it is not clear whether the silica arises
from the carbonation or unreacted silica gel from the sodium sili-
3.2. Derivative thermal gravimetry (DTG) cate powder. The cause will be discussed in the discussion section
with the results of the FT-IR spectra.
The results of the derivative thermal gravimetric (DTG) analysis
of the LS samples are shown in Fig. 2(a). A reduction of weight of
the sample between 150 and 400 °C was observed due to the 3.3. Fourier transform-infrared (FT-IR) spectroscopy
evaporation of the physically absorbed water from the reaction
products (i.e., C-S-H gel and aluminosilicate gel) [22,4]. The weight Fig. 3(a) shows the results of the FT-IP spectra of the PS samples.
loss of the LS10 sample was lower than that of LS30 and LS50 sam- The bands at 3440 cm1 and 1650 cm1 are related to the O–H
ples over the entire temperature range; the weight loss of LS30 stretching and bending vibrations of molecular water, respectively
sample was similar to that of the LS50 sample between 150 and [24,12]. The band absorption for the PS10 sample was very weak at
300 °C and was slightly lower between 350 and 600 °C than that 1650 cm1, while that for the LS10 sample was stronger, as shown
in Fig. 3(a) and (b).
An absorption band at 790 cm1 appeared for all samples. In
0
particular, as the amount of added slag decreased, the absorption
-0.05 band at 790 cm1 became strong, as shown in Fig. 3(a) and (b). This
band is related to the symmetric stretching vibrations of the Si–O–
-0.1
Si (Al) bridges [24,12] due to the presence of the aluminosilicate
DTG (%/min)

LS10
-0.15 gel and the glassy phase of the sample [12].
LS30 The level of carbonation in the AFS samples could be deter-
-0.2 mined through the intensity of the absorption band at
LS50
-0.25 1465 cm1. The PS30 and PS50 samples contained carbonate spe-
cies characterized by a large absorption band near 1450 cm1
-0.3 [24] related to the anti-symmetric stretching vibration of the
CO2
3 ions [46,32]; the PS10 sample showed very weak band
-0.35
0 100 200 300 400 500 600 700 800 900 1000 absorption. In Fig. 2(a), as the amount of added slag increased in
Temperature ( C) the mixture, the absorption band of the PS samples at 1450 cm1
associated with atmospheric carbonation became strong, while
(a) Effect of slag content in AFS sample with liquid activator
that of the LS samples was nearly constant.
The absorption bands at approximately 1200, 1100, 950, 900,
0.05
and 850 cm1 are associated with the Si–O–Si stretching vibrations
0 of the SiQn units for n = 4, 3, 2, 1, and 0, respectively [7]. These val-
-0.05 ues move into lower wavenumbers as the amount of aluminum
substitution for silicon increases [24]. The wavenumbers corre-
DTG (%/min)

-0.1
LS30
sponding to the maximum peaks of the PS10, PS30 and PS50 sam-
-0.15 ples are 1079, 1018, and 991 cm1 in Fig. 3(c), respectively. The
PS30
-0.2
wavenumbers related to the maximum peaks of the LS10, LS30,
and LS50 samples are 1023, 995, and 975 cm1, respectively, as
-0.25 shown in Fig. 3(c). In both the LS and PS samples, as the amount
-0.3 of slag increased, the maximum peak shifted toward a lower
-0.35 wavenumber. These results reveal the distribution of the Q1 and
0 100 200 300 400 500 600 700 800 900 1000 1100 Q2 units for the PS30, LS30, PS50, and LS50 samples, while the shift
Temperature ( C) toward a higher wave number for the PS10 and LS10 samples indi-
(b) Comparison between PS50 sample with powder activator and LS50
cates the presence of a higher degree of polymerized units (Q3 and
Q4 units) [24]. Accordingly, the more the amount of slag was added
sample with liquid activator
to the samples, the less the polymerized structure of the AFS phase
Fig. 2. Differential thermogravimetric analysis (DTG) results. was formed.
306 N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312

0.9
795
1650 0.8 S10
S20
Transmission (%)

0.7
PS10
0.6 S30

Al/Si ratio
PS30 S40
0.5
S50
PS50 0.4
0.3
1465
0.2
465 960-1000 3450
0.1
0 1000 2000 3000 4000 5000 0
Wavenumbers (cm-1) 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
Ca/Si ratio
(a) AFS samples with powder activator
(a) Al/Si ratio versus Ca/Si ratio

795 0.35
1650
S10
Transmission (%)

LS10
0.3
S20
LS30 0.25 S30

Na/Si ratio
S40
LS50 0.2
S50
0.15

1465 0.1
465 960-1000 3450
0.05
0 1000 2000 3000 4000 5000
0
Wavenumbers (cm-1)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4 1.5
(b) AFS samples with liquid activator Ca/Si ratio
(b) Na/Si ratio versus Ca/Si ratio

Fig. 4. EDS analysis of reaction products in PS samples.


Transmission (%)

PS10
trend of the S20 sample was similar to that of the S10 sample.
AS10
PS30 Thus, the reaction products measured by EDS are likely to be the
AS30 N-A-S-H gel and the C-(N)-A-S-H gel in the S10 and S20 samples.
PS50 The reaction products identified in the S30, S40 and S50 samples
AS50 are mainly the C-N-A-S-H gel, the compositions of which are
included in the dashed rectangle on the right-hand side in
Fig. 4(b). Nevertheless, it was difficult to find the N-A-S-H gel in
S30, S40 and S50 samples through an EDS analysis.
The average of atomic ratio is given in Fig. 5. The atomic ratios
800 900 1000 1100 1200 1300
of the LS sample with the liquid activator are not significantly dif-
Wavenumbers (cm-1) ferent from that of the PS sample with the powder activator. The
Al/Si ratios were 0.27 and 0.27 in the LS30 and PS30 samples,
(c) 900~1200 cm-1 wavenumbers of AFS samples with powder activator
respectively; the Na/Si ratios were 0.09 and 0.17, and the Ca/Si
and liquid activator
ratios were 0.64 and 0.62. As the amount of added slag increased
Fig. 3. FT-IR spectra of AFS samples with different slag contents (10%, 30% and 50%). from 10% to 50%, the Ca/Si ratio increased from 0.21 to 0.81 in
the LS and PS samples. The Al/Si ratio decreased from 0.38 to
0.25 as the amount of slag increased in the PS sample, but was
3.4. Scanning electron microscopy (SEM) with energy-dispersive nearly constant at 0.25 in the LS sample. The Al/Si atomic ratio of
spectroscopy (EDS) raw fly ash was 0.12 and that of raw slag was 0.22, as listed in
Table 1. Although the Al/Si ratio of raw slag is higher than that of
Fig. 4 shows the atomic ratios of the reaction products in PS the raw fly ash, the Al/Si ratio of the PS sample was decreased by
samples with the powder activator. The atomic ratio of each sam- increasing the amount of slag. This has occurred since the degree
ple was measured for 20 times. Two apparent groups are shown in of Al substitution for Si was decreased as the amount of added slag
Fig. 4(a); N-A-S-H gel and C-(N)-A-S-H gel, as indicated by Ismail increased. The higher the degree of the Al substitution for the Si
et al. [19]. As the amount of slag increased from 10% to 50% in was, the higher the degree of polymerized units (Q3 and Q4 units)
the mixture, the main reaction product was N-C-A-S-H rather than in the AFS samples occurred. This is also supported by the fact that
N-A-S-H. In Fig. 4(b), there are three groups depending on the Ca/Si the AFS sample with the low amount of added slag contains the
ratio and the Na/Si ratio. From the S10 sample in Fig. 4(b), it has reaction products with a highly polymerized structure, as shown
been seen that as the Ca/Si ratio of the reaction products increased, in the NMR spectroscopy results listed in Table 4.
the Na/Si ratio decreased slightly. The average of Na/Si ratios in the In the PS samples, the amount of added slag did not remarkably
group (i.e., the dashed rectangle on the left-hand side of Fig. 4(b)) affect the Na/Si ratio (0.08–0.17), and the Na/Si ratio was the high-
was higher than that in the middle of the groups. Meanwhile the est when the amount of added slag was 30%. In the LS samples, the
N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312 307

0.9 with raw materials (i.e., fly ash and slag) to form aluminosilicate
0.8 Al/Si gel or C-S-H gel. The soluble Si or Na element content is defined
0.7 Na/Si as the alkali-activator content that did not react with raw materi-
Ca/Si als, but dissolved in distilled water.
Element ratios

0.6
In Table 3, the insoluble Si and Na contents did not significantly
0.5
change regardless of the amount of slag when the liquid alkali-
0.4 activator was used. However, the insoluble Si and Na contents
0.3 were slightly higher in the LS30 sample than in the LS10 and
0.2 LS50 samples. The insoluble Si content of the alkali-activator was
higher than the insoluble Na content of the alkali-activator in all
0.1
samples except for the PS 10 sample. The high insolubility of Si
0
0 10 20 30 40 50 60 (over 80%) can result in an increase in the Si/Al ratio of the reaction
Slag amount (%) products by using sodium silicate instead of using sodium hydrox-
ide as an alkali-activator. The high contents of insoluble Al and Ca
(a) PS sample with powder activator
(above 98%) indicate that the raw fly ash and slag particles were
not soluble in distilled water, in contrast with sodium silicate.
0.8
Al/Si
0.7 3.6. Nuclear magnetic resonance (NMR) spectroscopy
Na/Si
0.6 Ca/Si
Fig. 6(a) shows the 29Si NMR spectra of the raw materials (slag
Element ratios

0.5
and fly ash). The spectrum of the raw fly ash contains peaks at 94,
0.4 99, 104, and 109 ppm, which were associated with the initial
0.3 vitreous material [14,15], while most of the peaks appearing above
108 ppm were assigned to crystalline silica phases (Q4 (0Al)
0.2 units), such as quartz (108 ppm) [13]. The spectrum of the raw
0.1 slag contains a dominant peak at 77 ppm.
0 Fig. 6(b) shows the spectrum of the PS samples with the pow-
0 10 20 30 40 50 60 dered activator. As the amount of added slag increased, the inten-
Slag amount (%) sity of the broad spectrum between 80 and 110 ppm increased.
(b) LS sample with liquid activator The PS10 and PS30 samples showed slightly higher intensity
between 80 and 110 ppm compared to that of the raw fly ash,
Fig. 5. Average element ratios of reaction product phases in AFS samples depending whereas the PS50 sample showed a strong broad peak between
on the slag content.
the maximum peaks of the raw slag and the raw fly ash. Fig. 6(c)
presents the spectrum of the LS samples with the liquid activator.
The spectra of the LS and PS samples were very different. The LS10
Na/Si ratio was nearly constant at 0.08–0.12. The amount of added sample presented a very broad peak between 85 ppm and
slag led to a significant change of the Ca/Si ratio, while it led to a 106 ppm, which was similar to that of the PS50 sample. However,
slight change of the Al/Si ratio and the Na/Si ratio. the LS30 and LS50 samples showed two predominant peaks at
82 ppm to 83 ppm and at 109 ppm. The peak intensity for
3.5. Insoluble elements of an alkali-activator the LS50 sample at 82 to 83 ppm was higher than that of the
LS30 sample, while that of the LS50 sample at 109 ppm was
An ICP-OES method was used to determine the reactivity of the smaller than that of the LS30 sample. The apparent difference
alkali-activators (i.e., sodium silicate powder and liquid). The unre- among the maximum peaks indicates that the LS50 (or LS30) and
acted alkali-activator can be dissolved in distilled water, while the LS10 have different silicate structures. Accordingly, it is clear that
reacted alkali-activator cannot be dissolved in the water due to its the silicate structure of the reaction product of the AFS binder is
incorporation in the reaction products. Hence, the reacted Si and significantly affected by the amount of added slag.
Na contents of the sodium silicate can be indirectly calculated by The NMR spectra of the AFS samples were deconvoluted with
measuring the amount of dissolved alkali activator. Origin software (OriginLab Corporation) to quantify the reaction
The reactivity of the alkali-activator was calculated and products. The relative areas below the corresponding fitted peaks
calibrated by measuring the soluble Si or Na content of the alka- were calculated by using deconvolution peaks from the raw mate-
li-activator dissolved in distilled water, as shown in Eq. (1). It has rials (fly ash and slag) and the AFS samples. The NMR peaks
been assumed that the reacted Si and Na in the alkali-activator appearing at 88, 93, 99, 104 and 108 ± 1 ppm are attribut-
are not dissolved in the distilled water. A powdered AFS sample ed to silicate tetrahedra (Q4) surrounded by 0, 1, 2, 3 or 4 alu-
was added to distilled water in order to determine the insoluble minum tetrahedra, respectively [13,21,30,34]. The deconvolution
Si, Al, Ca, and Na contents by ICP-OES. After sonicating the pow- peaks were produced on the basis of these peaks associated with
dered sample for 2 h at 25 °C, the components dissolved in the dis- Q4 (nAl), where n = 1, 2, 3, and 4. Table 4 illustrates the relative
tilled water were quantified. The reactivity of the alkali-activator areas obtained from deconvolution and the peak analysis results
can be suggested as follows: via Si NMR spectroscopy.
  The fly ash-reactivity and slag-reactivity of the AFS sample can
Ew
R¼ 1  100ð%Þ ð1Þ be calculated as follows. First of all, the deconvolution peaks for the
Eo
AFS sample corresponding to the peaks at 89, 93, 98, 104,
Here Ew is the soluble Si or Na content in the AFS sample and Eo is and 109 ppm of raw fly ash are selected. Secondly, among the
the total Si or Na content in the alkali-activator. selected peaks, the peaks of which the intensity levels are higher
Table 3 shows the insoluble Si or Na element contents of the than that of the corresponding deconvolution peaks for the raw
sodium silicate powder and liquid. The insoluble Si or Na element fly ash are selected. The relative area is defined as an area below
content is defined as the alkali-activator content which reacted any deconvolution peak divided by the total area below all of the
308 N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312

Table 3
Insoluble Si or Na element content of powder and liquid sodium silicate in distilled
Slag
water (%).
Fly ash
Insoluble content LS10 LS30 LS50 PS10 PS30 PS50
Na 46.73 53.66 44.35 36.79 49.94 51.01
Si 84.99 89.40 87.79 29.20 82.45 84.22
Ala 99.42 99.45 98.92 99.89 99.06 98.79
Caa 98.97 99.44 99.46 98.41 98.56 98.65
a
Insoluble Al and Ca contents (%) means the ratio of the insoluble Al and Ca
contents into distilled water to the total Al and Ca contents in AFS sample,
respectively.
-40 -50 -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
Chemical shift (ppm)

deconvolution peaks. Afterward, the total sum of the relative areas (a) Raw materials
(A) below the selected deconvolution peaks for the AFS sample was
calculated, and the total sum of the relative areas (B) below the
selected deconvolution peaks for the raw fly ash was also calculat-
ed. The total sum of the areas (C) below the newly-appearing peaks
for the AFS samples, which do not match the deconvolution peaks
for raw fly ash, was also calculated. Accordingly, the total relative PS50

area below the peaks associated with the presence of reaction pro-
duct equals to (AB)+C. Finally, the amounts of reaction products PS30

and the reactivities of the raw materials could be calculated indi-


rectly using the method proposed by Buchwald et al. [6]. The pro- PS10
cedure for quantifying the reaction products and the reactivity is
demonstrated in Fig. 7 and the deconvolution peaks are shown in
-40 -50 -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
Figs. 8, 9 and 10.
Chemical shift (ppm)
Table 4 presents the relative area obtained from deconvolution
results and the peak analysis by Si NMR spectroscopy. The fly ash- (b) AFS samples activated by powder activator
reactivity, slag-reactivity and total reactivity were calculated by
using the relative area, obtained as follows (cf. [6]):

Sislag  U slag
Rslag ¼ ð%Þ ð2Þ LS50
Sislag

LS10
Sifly ash  U fly ash
Rfly ash ¼ ð%Þ ð3Þ
Sifly ash LS30

RTotal ¼ F c  Rfly ash þ Sc  Rslag ð%Þ ð4Þ


In these equations, Rslag and Rfly ash denote the fly ash-reactivity
and slag-reactivity in the AFS sample, and RTotal indicates the total -40 -50 -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
reactivity of the AFS sample. Sislag and Sifly ash are the silicon con- Chemical shift (ppm)
tent (%) in the raw slag and fly ash, respectively. Fc and Sc are the (c) AFS samples activated by liquid activator
raw fly ash and slag contents (%) in the AFS mixture, respectively.
29
The peaks of the AFS sample which was identical to the deconvo- Fig. 6. Si NMR spectra of the raw materials and ASF samples.
lution peaks of the raw fly ash and slag were selected. Uslag and Ufly
The reactivities are listed in Table 5. The total reactivities of the
ash are the sum of the relative areas below the selected deconvolu-
LS samples were 26.41% (LS10), 51.61% (LS30), and 61.77% (LS50);
tion peaks of which the intensity is lower than the peak intensity of
the values for the PS samples were 6.25% (PS10), 21.4% (PS30), and
the raw fly ash and the raw slag. That is, these peaks appear due to
44.98% (PS50). The total reactivity was even higher when mixed
the presence of the unreacted fly ash and slag particles remaining
the liquid activator compared to the powder activator.
in the AFS sample.

Table 4
29
Relative area obtained from deconvolution results and peak analysis of Si NMR spectroscopy (%).

Sample C-S-H Aluminosilicate (C-N-A-S-H)


Signal (ppm) 77, etc.a Q1 79 Q2 (1Al) 81 Q2 86 Q4 (4Al) 89 Q4 (3Al) 94 Q4 (2Al) 98 Q4 (1Al) 104 Q4 (0Al) 109, etc.b
Slag 100
Fly ash 0.66 3.60 10.75 24.35 60.64
PS10 0 0 0 0 0.21 4.54 9.7 23.1 62.5
PS30 2.44 0 2.41 1.04 4.62 9.53 13.42 27.16 39.38
PS50 3.45 0.55 1.03 5.3 6.29 6.10 10.26 7.79 34.07
LS10 0.40 0.35 0.53 2.55 4.81 3.77 3.5 9.86 55.6
LS30 1.15 4.27 12.89 18.55 2.66 7.21 3.41 6.53 34.24
LS50 14.58 10.65 22.98 17.64 0.96 0.64 1.48 8.10 12.76
a
The peaks at 77, 75, 72, 70, 68, 65 ppm are attributed to the presence of raw slag in AFS sample.
b
These are the peaks above 109 ppm. Most of the peaks appearing above 108 ppm were assigned to different crystalline phases of silica (Q4(0Al) signals) [13].
N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312 309

NMR spectrum PS10


By using Lorentz curve of
Origin software and the Fit
Deconvolution of Raw materials and literature study
AFS sample

Selection of the same deconvolution


peaks of AFS sample as that of
raw fly ash

Selection of the peak of which the -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
intensity is higher than that of Chemical shift (ppm)
raw fly ash
(a) PS10 AFS sample

Total relative area of the selected


peaks for AFS sample PS30
Fit
Total relative area of the selected
peaks for raw fly ash

Total relative area of newly


appearing peaks for AFS sample

-60 -70 -80 -90 -100 -110 -120 -130 -140 -150
Total relative area of the peaks
appearing due to reaction products Chemical shift (ppm)

(b) PS30 AFS sample


Using Eqs. 2, 3 and 4
Reactivity of raw material and total
reactivity
PS50

Fig. 7. Procedure for quantifying the reaction products and the reactivity. Fit

Fly ash
Fit

-60 -70 -80 -90 -100 -110 -120 -130 -140 -150
Chemical shift (ppm)

(c) PS50 AFS sample

Fig. 9. 29Si NMR spectra and deconvolution of ASF samples activated by powder
activator.

-60 -70 -80 -90 -100 -110 -120 -130 -140 -150
Chemical shift (ppm) 4. Discussions
(a) Raw fly ash
4.1. Reactivity

Slag The fly ash-reactivity, slag-reactivity and total reactivity of the


Fit AFS paste (liquid type and powder type) were calculated by means
of an NMR analysis and an ICP-OES analysis, respectively. In
Table 5, the weight loss measured by DTG, the absorption extent
of the band measured by FT-IR, and the reactivity obtained from
the NMR spectra were compared. The weight loss up to 500 °C
was related to the dehydration of the reaction product (i.e., alumi-
nosilicate gel and C-S-H gel). Likewise, the absorption band at
3450 cm1 was related to the O-H stretching band of molecular
-60 -70 -80 -90 -100 -110 -120 -130 -140 -150 water, revealing the degree of hydration. On the other hand, the
Chemical shift (ppm) reactivity was measured by using the NMR technique.
In Table 5, the reactivity, weight loss and extent of absorption
(b) Raw slag
increased as the amount of slag increased. The total reactivity by
Fig. 8. 29
Si NMR spectra and deconvolution of raw materials. NMR was twice as high in the LS30 sample as it was in the PS30
310 N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312

Table 5
Fly ash-reactivity, slag-reactivity and total reactivity of AFS samples activated by LS10
sodium silicate powder and liquid (%).
Fit
Reactivity LS10 LS30 LS50 PS10 PS30 PS50
Fly ash 23.15 38.82 60.93 4.0* 1.57 22.97
Slag 91.52 94.23 63.65 100* 87.81 88.88
Total reactivity 26.41 51.61 61.77 6.25 21.4 44.98
Weight loss by TGAa 4.59 6.67 7.54 –** 6.76 –**
Absorption extent by 51.94 64.46 68.31 23.50 59.56 63.50
FT-IRb
*
In the case of PS10 sample, the fly ash-reactivity was minus 4.0, and the slag-
reactivity was 100%. Since the Si NMR spectrum assigned to raw slag was not -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
observed in PS10 sample, all the raw slag particles seemed to react completely with Chemical shift (ppm)
alkali-activator, thus the slag-reactivity was determined to be 100%. The reasons (a) LS10 AFS sample
why the fly ash-reactivity was below zero are (1) the actual fly ash-reactivity was
very low (less than 5%), and (2) the noise of 29Si NMR spectrum occurred. Even
though there are some difficulties in measuring the 29Si NMR spectra, it is certain
that the fly ash-reactivity is very low when the amount of added slag is 10%, and the LS30
slag-reactivity is nearly 100%.
**
Fit
TGA tests of PS10 and PS50 samples were not conducted.
a
Weight loss measured from TGA analysis.
b
Absorbed band percentage (%) related to O–H stretching bend at around 3400–
3500 cm1.

sample; however, the weight loss, 6.67% of the LS30 sample by DTG
and the extent of absorption, 64.46% by FT-IR were fairly similar to
those of the PS30 sample (i.e., 6.76% and 59.56%, respectively). -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
These differences mean that the type of alkali-activator (liquid or Chemical shift (ppm)
powder) did affect the reactivity of each raw material (i.e., fly ash (b) LS30 AFS sample
and slag) measured by the NMR analysis. The LS30 and PS30 sam-
ples were similar in the degree of the hydration measured by the
DTG regardless of the type of the alkali-activator. Equally, both of
LS50
the samples were comparable in the reactivity of the alkali-activa-
tor regardless of the amount of slag as listed in Table 3. It seems Fit
that the total reactivity of AFS binder is strongly dependent upon
the mixture ratio of raw materials as well as the type of alkali-
activator, while the degree of hydration is associated with the
reactivity of alkali-activator.
Table 5 shows the fly ash-reactivity and slag-reactivity, and
total reactivity of AFS sample. The NMR peak of the PS10 sample
overlapped with the noise peaks during the measurements by
decreasing the reliability of the results so that the peak intensity -60 -70 -80 -90 -100 -110 -120 -130 -140 -150
of the PS10 sample was very low, as shown in Fig. 9(a). The slag-re- Chemical shift (ppm)
activity in all of the samples was higher than the fly ash-reactivity
(c) LS50 AFS sample
since the raw slag had a latent hydraulic property unlike fly ash,
and was primarily composed of amorphous phases. In addition, 29
Fig. 10. Si NMR spectra and deconvolution of ASF samples activated by liquid
the fly ash-reactivity was much higher when mixed with the liquid activator.
activator compared to the powder activator, while the slag-reac-
tivity was roughly equal regardless of alkali activator. Therefore,
using the liquid activator resulted in the relatively higher fly ash- [24], the Q1 and Q2 units are similar to the silicate structure of
reactivity, which can contribute to the higher compressive strength C-S-H in alkali-activated slag or OPC, and the silicate structure of
of the AFS binder compared to the results with the powder aluminosilicate gel in a geopolymer is similar to the Q3 and Q4
activator. structures. Based on this, it was found that both the amount of
reaction product and its silicate structure were remarkably varied
4.2. Reaction products depending on the amount of slag added, as listed in Table 4. The
microstructural characteristics can generally affect the durability
The results from XRD, FT-IR, EDS, and NMR analysis for the AFS of the AFS paste. It was reported that aluminosilicate gel was resis-
paste provide important new insight into the structural character- tant to acids, while C-S-H gel was vulnerable to acids [1]. Hence,
istics of the reaction products. A highly polymerized silicate struc- future research on the acid resistance of AFS paste based on the
ture with Q4 (i.e., 103 to 115 ppm) and Q3 (i.e., 95 to amounts of C-S-H gel and aluminosilicate gel needs to be
100 ppm) units was identified when the amount of slag was conducted.
10% in the LS sample with the liquid activator, whereas reaction The amounts of the aluminosilicate gel and C-S-H gel in the AFS
products with Q1 (i.e., 74 to 78 ppm) and Q2 (i.e., 83 to sample were evaluated quantitatively by using the NMR results.
88 ppm) units, but not with Q4 and Q3 units, were observed The chemical composition ratios of these reaction products were
when the amounts of slag was 30% and 50% in the LS sample. measured by using the EDS results. Fig. 4 shows the distribution
Fig. 6(c) clearly shows the differences among the maximum NMR of the atomic ratios for the reaction products in AFS samples
peaks of the LS samples. As mentioned in the previous research measured by using EDS. It was difficult to distinguish between
N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312 311

the C-S-H gel and aluminosilicate gel in spite of applying the EDS Table 6
29
point analysis results. As shown in Fig. 5, although the Ca/Si ratios Si/Al ratios and chain length of the reaction products measured from Si NMR
spectroscopy and EDS analyses.
were varied depending on the amount of added slag, a substantial
amount of Ca (i.e., Ca/Si ratio of 0.2–0.8) was incorporated into Sample Al-substituted C-S-H Ca-aluminosilicate Si/Al by EDS
the reaction products. The silicate structures of the reaction prod- Chain length Si/Al Si/Al
ucts are shown in Table 4. The C-S-H consists of Q1 (chain end PS10 – – 4.93 2.57
groups) and Q2 (middle groups in chains) units, while the alumi- PS30 – 2.87 2.89 3.62
nosilicate gel consists of Q4 (4Al), Q4 (3Al), Q4 (2Al), and Q4 (1Al) PS50 26.72 13.36 2.08 3.83
units; these units are known to be three-dimensional cross-linked LS10 21.34 12.98 2.81 4.11
LS30 19.74 5.54 2.57 3.57
sites with the different amounts of Al substitution for Si. Fig. 5
LS50 11.79 4.46 4.26 4.02
and Table 4 show that the aluminosilicate gel with Q4 (nAl) units
with n between 0 and 4 contains the certain amount of Ca.
From the fact that the geopolymer is a reaction product formed rich and calcium-deficient phases were not identified in the SEM
by polycondensation according to Davidovits [9], it can be conclud- length scale [26]. Accordingly, each of the Si/Al ratios in Fig. 4
ed that the aluminosilicate gel in the present study is likely to be a can be either for the C-S-H or Ca-based aluminosilicate.
Ca-based geopolymer, and the C-S-H gel is similar to C-S-H gel in The chain length (11.79) of the C-S-H phase for the LS50 sample
OPC or alkali-activated slag. Therefore, the Ca-based geopolymer in Table 6 was similar to the chain length (13.4) given in Buchwald
and C-S-H were only slightly different in terms of their chemical et al. [6] for the same mixture ratio. However, the Si/Al ratios (i.e.,
compositions but were quite different in terms of their silicate 4.46 and 4.26) of the Ca-based aluminosilicate and the C-S-H phase
structures; both formed simultaneously in the AFS sample. The dif- were quite different from the values (i.e., 1.19 and 2.8) of the alu-
ferent types of silicate tetrahedra in the Ca-based geopolymer and minosilicate and the C-S-H measured by Buchwald et al. [6] as the
C-S-H are quantitatively assessed in Table 4. different types of alkali-activators were used; in the present study,
The results from the DTG and FT-IR analyses indicate that the the liquid sodium silicate containing a substantial amount of sili-
reaction product in the AFS sample may have been carbonated. con was used, whereas, in the their study, the liquid sodium
In the FT-IR data, the absorption bands at 1465 cm1 related to car- hydroxide was used as an alkali-activator. The silicate ion in the
bonation were observed in all of the samples; however, this signal liquid sodium silicate condensed to form Si-O-Si (Al) bonds [8],
was very weak in the PS10 sample. There is no obvious relationship resulting in the higher Si/Al ratio in the reaction products. There-
between the degree of carbonation and the amount of added slag. fore, the Si/Al ratios of the C-S-H gel and Ca-based aluminosilicate
The LS sample with the liquid activator had the higher degree of gel were much higher than the results of Buchwald et al. [6].
carbonation than the PS sample with the powder activator due to
the larger amount of C-S-H in the LS sample, as listed in Table 4. 5. Concluding remarks
In Fig. 2(b), the PS30 sample with the powder activator showed
widely distributed peaks at a range of 500–700 °C due to the The results of an experimental study conducted to evaluate the
removal of the water absorbed by the unreacted silica gel in the reactivity and reaction products of alkali-activated, fly ash/slag
powder activator or silica gel resulting from the decomposition paste have been summarized. The following conclusions can be
of carbonate. The FT-IR results show that both LS and PS samples drawn from the results presented in this paper.
were affected by carbonation. Therefore, the high peak of the PS
30 sample at 500–700 °C is more likely to be due to the dehydra- 1. The total reactivity of the AFS paste calculated by an NMR peak
tion of absorbed water from the unreacted silica gel in the powder analysis was much higher when mixed with the liquid activator
activator rather than the dehydration of silica gel resulting from compared to the powder activator. The reactivity of the liquid
the decomposition of carbonate. activator was slightly higher than that of the powder activator.
As the amount of added slag increased, the maximum absorp- 2. The fly ash-reactivity was significantly higher when mixed with
tion band shifted toward lower wavenumbers, as shown in the liquid activator compared to the powder activator; the slag-
Fig. 3(c). The wavenumbers for the LS10, LS30 and LS50 samples reactivity was comparable regardless of the activator types. The
at their maximum intensity levels were 1023, 995, and 975 cm1, use of the liquid activator resulted in an increase in the fly ash-
while those of the PS10, PS30, and PS50 samples were 1079, reactivity and the total reactivity.
1018 and 991 cm1, respectively. Since the amount of C-S-H with 3. The amount of added slag primarily affected the amount of
Q1 and Q2 units in the AFS sample increased with the amount of reaction product and its silicate structure. As the amount of slag
added slag, the absorption band shifted toward lower wavenum- increased, the amount of C-S-H with Q1 and Q2 units increased,
bers, which was in agreement with the results of the previous whereas the amount of aluminosilicate gel with Q4 units
research [24]. In addition, when the same amount of slag was decreased. A highly polymerized silicate structure with a Q4
added in the LS and PS samples, the maximum absorption band unit was identified in all of the LS samples with the liquid acti-
in the LS sample was lower than that in the PS sample since the vator, while the reaction products with Q1 and Q2 units were
amount of C-S-H in the LS sample was higher than that in the PS identified when the amounts of added slag were 30% and 50%
sample, as listed in Table 4. in the LS samples.
The Si/Al ratios from the EDS analysis were compared with the 4. The EDS analysis and the 29Si NMR spectra showed that the alu-
Si/Al ratios obtained from Engelhard’s equation [13] and the NMR minosilicate gel with Q4 (nAl) units contained a certain amount
results in Table 6. The Si/Al ratios measured by the EDS analysis of Ca. This indicates that considering chemical composition and
were not similar to any of those of the C-S-H gel and Ca-based alu- silicate structure, the aluminosilicate gel was similar to a Ca-
minosilicate gel. Since it was difficult to distinguish between C-S-H based geopolymer. The chemical compositions of the Ca-based
and Ca-based aluminosilicate by using only a BSE image on SEM, geopolymer were nearly the same as those of C-S-H whereas
the Si/Al ratio determined by the EDS analysis has been the average their silicate structures were different.
of either the C-S-H or Ca-based aluminosilicate Si/Al ratios ran-
domly measured by the EDS analysis; although it might be possible The results of this research will help assess the effects of the
to distinguish between the C-S-H and Ca-based aluminosilicate by amounts of the reaction products and reactivity levels on the dura-
using elemental maps for calcium via SEM, the discrete calcium- bility of alkali-activated, fly ash/slag binders. Future works on the
312 N.K. Lee, H.K. Lee / Construction and Building Materials 81 (2015) 303–312

acid and chloride resistance of AFS paste will be followed in the [21] Klinowski J. Nuclear magnetic resonance studies of zeolites. Prog Nucl Magn
Reson Spectrosc 1984;16:237–309.
future on the basis of the amounts of C-S-H gel and aluminosilicate
[22] Kong DL, Sanjayan JG. Effect of elevated temperatures on geopolymer paste,
gel. mortar and concrete. Cem Concr Res 2010;40(2):334–9.
[23] Kumar S, Kumar R, Mehrotra SP. Influence of granulated blast furnace slag on
Acknowledgements the reaction, structure and properties of fly ash based geopolymer. J Mater Sci
2010;45(3):607–15.
[24] Lecomte I, Henrist C, Liegeois M, Maseri F, Rulmont A, Cloots R. (Micro)-
This research was supported by a Grant (Code 11-Technology structural comparison between geopolymers, alkali-activated slag cement and
Innovation-F04) from Construction Technology Innovation Pro- Portland cement. J Eur Ceram Soc 2006;26(16):3789–97.
[25] Lloyd RR, Provis JL, van Deventer JS. Microscopy and microanalysis of inorganic
gram (CTIP) funded by Ministry of Land, Transportation and Mar- polymer cements. 2: the gel binder. J Mater Sci 2009;44(2):620–31.
itime Affairs (MLTM) of Korean government. [26] Lloyd RR, Provis JL, van Deventer JS. Microscopy and microanalysis of inorganic
polymer cements. 1: remnant fly ash particles. J Mater Sci 2009;44(2):608–19.
[27] Lee NK, Lee HK. Setting and mechanical properties of alkali-activated fly ash/
References slag concrete manufactured at room temperature. Constr Build Mater
2013;47:1201–9.
[1] Allahverdi A, Skvara F. Nitric acid attack on hardened paste of geopolymeric [28] Lee NK, Kim HK, Park IS, Lee HK. Alkali-activated, cementless, controlled low-
cements. Part 1. Ceramics 2001;45(3):81–8. strength materials (CLSM) utilizing industrial by-products. Constr Build Mater
[2] Bernal SA, de Gutierrez RM, Provis JL, Rose V. Effect of silicate modulus and 2013;49:738–46.
metakaolin incorporation on the carbonation of alkali silicate-activated slags. [29] Lee NK, Jang JG, Lee HK. Shrinkage characteristics of alkali-activated fly ash/
Cem Concr Res 2010;40(6):898–907. slag paste and mortar at early ages. Cem Concr Compos 2014;53:239–48.
[3] Bernal SA, Provis JL, Rose V, Mejía de Gutierrez R. Evolution of binder structure [30] Lippmaa E, Mägi M, Samoson A, Tarmak M, Engelhardt G. Investigation of the
in sodium silicate-activated slag-metakaolin blends. Cement Concr Compos structure of zeolites by solid-state high-resolution silicon-29NMR
2011;33(1):46–54. spectroscopy. J Am Chem Soc 1981;103(17):4992–6.
[4] Bernal SA, Rodríguez ED, de Gutiérrez RM, Gordillo M, Provis JL. Mechanical [31] Luo X, Xu J, Bai E, Li W. Systematic study on the basic characteristics of alkali-
and thermal characterisation of geopolymers based on silicate-activated activated slag-fly ash cementitious material system. Constr Build Mater
metakaolin/slag blends. J Mater Sci 2011;46(16):5477–86. 2012;29:482–6.
[5] Bernal SA, Provis JL, Walkley B, San Nicolas R, Gehman JD, Brice DG, et al. Gel [32] Mollah MYA, Lu F, Cocke DL. An X-ray diffraction (XRD) and Fourier transform
nanostructure in alkali-activated binders based on slag and fly ash, and effects infrared spectroscopic (FT-IR) characterization of the speciation of arsenic (V)
of accelerated carbonation. Cem Concr Res 2013;53:127–44. in Portland cement type-V. Sci Total Environ 1998;224(1):57–68.
[6] Buchwald A, Hilbig H, Kaps C. Alkali-activated metakaolin-slag blends— [33] Mostafa NY, El-Hemaly SAS, Al-Wakeel EI, El-Korashy SA, Brown PW.
performance and structure in dependence of their composition. J Mater Sci Characterization and evaluation of the hydraulic activity of water-cooled
2007;42(9):3024–32. slag and air-cooled slag. Cem Concr Res 2001;31(6):899–904.
[7] Clayden NJ, Esposito S, Aronne A, Pernice P. Solid state 27 Al NMR and FTIR [34] Neuhoff PS, Stebbins JF, Bird DK. Si–Al disorder and solid solutions in analcime,
study of lanthanum aluminosilicate glasses. J Non-Cryst Solids chabazite, and wairakite. Am Mineral 2003;88(2–3):410–23.
1999;258(1):11–9. [35] Palomo A, Glasser FP. Chemically-bonded cementitious materials based on
[8] Criado M, Fernández-Jiménez A, Palomo A, Sobrados I, Sanz J. Effect of the metakaolin. Br Ceram Trans J 1992;91(4):107–12.
SiO2/Na2O ratio on the alkali activation of fly ash. Part II: 29Si MAS-NMR [36] Peng GF, Huang ZS. Change in microstructure of hardened cement paste
Survey. Microporous Mesoporous Mater 2008;109(1):525–34. subjected to elevated temperatures. Constr Build Mater 2008;22(4):593–9.
[9] Davidovits J. Chemistry of geopolymeric systems, terminology. In: [37] Puertas F, Fernández-Jiménez A. Mineralogical and microstructural
Geopolymer, 99; (1999). p. 9–40. characterisation of alkali-activated fly ash/slag pastes. Cement Concr
[10] Davidovits J. Geopolymers: inorganic polymeric new materials. J Mater Educ Compos 2003;25(3):287–92.
1994;16:91–138. [38] Puligilla S, Mondal P. Role of slag in microstructural development and
[11] Duxson P, Provis JL, Lukey GC, Van Deventer JS. The role of inorganic polymer hardening of fly ash-slag geopolymer. Cem Concr Res 2013;43:70–80.
technology in the development of ‘green concrete’. Cem Concr Res [39] Thiery M, Villain G, Dangla P, Platret G. Investigation of the carbonation front
2007;37(12):1590–7. shape on cementitious materials: effects of the chemical kinetics. Cem Concr
[12] El-Didamony H, Amer AA, El-Sokkary TM, Abd-El-Aziz H. Effect of substitution Res 2007;37(7):1047–58.
of granulated slag by air-cooled slag on the properties of alkali activated slag. [40] Villain G, Thiery M, Platret G. Measurement methods of carbonation profiles in
Ceram Int 2013;39(1):171–81. concrete: thermogravimetry, chemical analysis and gammadensimetry. Cem
[13] Engelhardt G, Michel D. High-Resolution Solid-State NMR of Silicates and Concr Res 2007;37(8):1182–92.
Zeolites. New York: Wiley; 1987. [41] van Jaarsveld JGS, Van Deventer JSJ, Schwartzman A. The potential use of
[14] Fernández-Jiménez A, Palomo A. Characterisation of fly ashes. Potential geopolymeric materials to immobilise toxic metals: Part II. Material and
reactivity as alkaline cements. Fuel 2003;82(18):2259–65. leaching characteristics. Miner Eng 1999;12(1):75–91.
[15] Fernández-Jimenez A, De La Torre AG, Palomo A, López-Olmo G, Alonso MM, [42] van Jaarsveld JGS, Van Deventer JSJ, Lukey GC. The characterisation of source
Aranda MAG. Quantitative determination of phases in the alkali activation of materials in fly ash-based geopolymers. Mater Lett 2003;57(7):1272–80.
fly ash. Part I. Potential ash reactivity. Fuel 2006;85(5):625–34. [43] Xu H, van Deventer JSJ. The geopolymerisation of aluminosilicate minerals. Int
[16] Gruskovnjak A, Lothenbach B, Holzer L, Figi R, Winnefeld F. Hydration of alkali- J Miner Process 2000;59(3):247–66.
activated slag: comparison with ordinary Portland cement. Adv. Cem. Res. [44] Yip CK, Lukey GC, van Deventer JSJ. The coexistence of geopolymeric gel and
2006;18(3):119–28. calcium silicate hydrate at the early stage of alkaline activation. Cem Concr Res
[17] Haha MB, Le Saout G, Winnefeld F, Lothenbach B. Influence of activator type on 2005;35(9):1688–97.
hydration kinetics, hydrate assemblage and microstructural development of [45] Yip CK, Van Deventer JSJ. Microanalysis of calcium silicate hydrate gel formed
alkali activated blast-furnace slags. Cem Concr Res 2011;41(3):301–10. within a geopolymeric binder. J Mater Sci 2003;38(18):3851–60.
[18] Hou X, Kirkpatrick RJ, Struble LJ, Monteiro PJ. Structural investigations of alkali [46] Yu P, Kirkpatrick RJ, Poe B, McMillan PF, Cong X. Structure of calcium silicate
silicate gels. J Am Ceram Soc 2005;88(4):943–9. hydrate (C-S-H): near-, mid-, and far-infrared spectroscopy. J Am Ceram Soc
[19] Ismail I, Bernal SA, Provis JL, San Nicolas R, Hamdan S, van Deventer JS. 1999;82(3):742–8.
Modification of phase evolution in alkali-activated blast furnace slag by the [47] Zhao FQ, Ni W, Wang HJ, Liu HJ. Activated fly ash/slag blended cement. Resour
incorporation of fly ash. Cement Concr Compos 2014;45:125–35. Conserv Recycl 2007;52(2):303–13.
[20] Jang JG, Lee NK, Lee HK. Fresh and hardened properties of alkali-activated fly
ash/slag pastes with superplasticizers. Constr Build Mater 2014;50:169–76.

You might also like