Sunteți pe pagina 1din 10

Journal of Cleaner Production xxx (2015) 1e10

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Development of greener alkali-activated cement: utilisation of sodium


carbonate for activating slag and y ash mixtures
Ahmed F. Abdalqader*, Fei Jin, Abir Al-Tabbaa
Department of Engineering, University of Cambridge, Trumpington Road, Cambridge CB2 1PZ, United Kingdom

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 31 July 2015
Received in revised form
25 November 2015
Accepted 6 December 2015
Available online xxx

Alkali activated y ash/slag (AAFS), a newly evolved type of alkali-activated cements (AACs), is here
studied with the aim of developing a more sustainable alternative to Portland cement (PC), known for its
adverse environmental impact. In this study, sodium carbonate (Na2CO3) was used as the alkali activator
for the y ash (FA)/slag blends. The effects of different factors on the strength, reaction rate, hydration
products and microstructure were examined; these factors include the activator dosage, FA/slag ratio,
and curing regime. It was found that increasing the Na2CO3 dosage signicantly increased the
compressive strength. The inclusion of up to 25 wt% y ash marginally decreased the compressive
strength up to 28 days while the inclusion beyond 25 wt% can lead to a remarkable reduction in strength,
particularly for water-cured specimens. Sealed curing in general was found to be benecial to the
strength development of AAFS paste especially at a 50/50 ratio of FA/slag. Both the activator dosage and
FA/slag ratio were found to have notable inuences on reaction rate and reaction products and microstructure. Increasing the activator dosage accelerates the kinetics of the reaction while increasing the FA/
slag ratio slows the reaction rates. The main binding phase is C-(N)-A-S-H with varying Ca/Si ratios
ranging between 0.6 and 1.0 depending mainly on FA/slag ratio. The results indicated the possibility of
production of greener cementing materials by utilizing appropriate ratios of FA/slag, dosage of sodium
carbonate, and curing regimes.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Fly ash
Slag
Sodium carbonate
Alkali-activated cements
Microstructure

1. Introduction
Carbon dioxide (CO2) emissions associated with Portland
cement (PC) production account for 5e10% of global man-made CO2
emissions (Scrivener and Kirkpatrick, 2008), a fact that puts an
onus on the cement industry to nd a more environmentally
benign alternative. Alkali activated cements (AACs), which have
been known about since the 1960s, are one of the candidates for
green concrete (Provis, 2013). AACs are cementitious materials
formed as a result of the dissolution of aluminosilicate precursors
from natural or industrial waste materials in an alkaline environment. These materials, when mixed with alkaline activators, set
and harden, producing a binder with good mechanical and durability properties. Most recent estimates indicated that 55%e75%
reduction of CO2 emissions compared to PC can be achieved by
using AACs in concrete (Yang et al., 2013).

* Corresponding author.
E-mail address: aalqader@gmail.com (A.F. Abdalqader).

The use of by-product materials in cement signicantly decreases the negative impacts on the environment. Firstly it is not
necessary to quarry virgin material to manufacture these byproduct materials. Secondly, it reduces the land demand for landlling. Finally, it saves energy (Higgins et al., 2011; Nehdi, 2011).
Meanwhile, these by-products are increasingly being generated on
a planetary scale (Nehdi and Mindess, 1999). Therefore, not only do
these materials improve the sustainability of cement but they also
compensate for the shortage in supply due to the increasingly high
demand of construction materials (Sarkar and Roumain, 2003).
Ground granulated blast-furnace slag (GGBS) and y ash (FA) are
the most prevalent industrial by-products used in the productions
of AACs. GGBS is a by-product from the manufacture of iron.
Globally, the annual production of GGBS is estimated to be
270e320 Mt (USGS, 2013a). While the production of 1 t of PC needs
5000 MJ of energy, with a corresponding CO2 emissions of 1 t, the
production of 1 t GGBS consumes only 1300 MJ of energy and emits
only 0.07 t CO2 to the atmosphere (Cabeza et al., 2013). FA is a byproduct material collected among other coal combustion products
from the burning of coal in power station furnaces. The worldwide

http://dx.doi.org/10.1016/j.jclepro.2015.12.010
0959-6526/ 2015 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

production of coal combustion products was about 777 Mt in 2010;


around half of them are FA (Heidrich et al., 2013).
Alkali activated slag (AAS) and alkali activated y ash (AAF) have
been studied extensively by various researchers as reported in
nez and Palomo, 2005; Horpibulsuk
the literature (Fern
andez-Jime
et al., 2014; Lloyd et al., 2008; Neramitkornburi et al., 2015;
Phetchuay et al., 2014; Rashad, 2014, 2013; Srinivasan and
Sivakumar, 2013; Suksiripattanapong et al., 2015; Wang et al.,
1994). Given the limited global resources of the individual byproducts, combining them would provide a much larger resource
and potentially counterbalance the disadvantages of each activation process (Abdalqader and Al-Tabbaa, 2015, 2014a, 2014b;
Puertas et al., 2000).
Like other AACs, the nature of the raw materials along with the
type and dosage of the activators have signicant effects on AAFS
properties (Nazari and Sanjayan, 2015; Shi et al., 2006). Other
important factors include the FA to GGBS ratio (Meja et al., 2015;
Puertas and Fernandez-Jimenez, 2003; Puertas et al., 2000) and
the curing condition (Provis and van Deventer, 2014). The strength
development of the alkali activated FA/slag (AAFS) system depends
mostly on the activator type and dosage as well as FA/slag ratio. The
most widely used activators are NaOH, waterglass, and a combination of NaOH and waterglass. These activators, however, are a
source of concern because they are the most expensive component
in the system, are known to shrink and harden more rapidly than is
desirable, are caustic in nature, and additionally are the primary
source of greenhouse gas (GHG) emissions in the production of AAC
concrete (Deventer et al., 2010; Habert et al., 2011). These reasons
have prevented the widespread use of AACs in real applications.
Other activators can play a vital role in overcoming these shortcomings. Among them is sodium carbonate (Na2CO3), which is
proposed here as an alternative activator in order to develop
greener AACs. Na2CO3 is a naturally occurring mineral and can be
obtained from trona and sodium carbonate-rich brines (USGS,
2013b). It can also be manufactured from one of several chemical
processes such as Solvay process (Shi et al., 2006). The worldwide
reserves of sodium carbonate are more than 24 billion t (USGS,
2013b). Na2CO3 is ~2e3 times cheaper than NaOH or sodium silicate (Kostick, 2013). In addition to the low cost of this activator
compared to conventional ones, it is safer to handle and has been
reported to yield lower drying shrinkage. Thus, the use of Na2CO3 as
an activator can contribute to the development of more sustainable
AACs.
The use of Na2CO3 as an activator is less extensively studied by
researchers as observed from the literature although it has been
shown that buildings made of Na2CO3-activated binders have
acceptable strength and excellent in-situ performance (Xu et al.,
2008). Compared to other conventional activators, Na2CO3 yields
a lower early age strength due to its lower pH, but it can demonstrate higher strength at later ages than NaOH resulting from the
effect of CO2
3 ions (Wang et al., 1994), which cause the formation of
carbonated compounds that improves mechanical strengths
ndez-Jime
nez et al., 1999). Li and Sun (2000) used Na2CO3
(Ferna
with and without NaOH to activate slag alone and a combination of
slag and FA. The compressive strength of 10% Na2CO3-activated slag
developed from 0 MPa at 3 days to 60 MPa at 28 days. The reaction
mechanism of the activation of aluminosilicate materials by Na2CO3
is yet to be fully understood. Recently, Bernal et al. (2014) examined
the role of Na2CO3 in the activation mechanism of AAS. Based on
their analytical results, they proposed that the reaction takes place
in three different stages starting with the dissolution of the slag and
the formation of gaylussite and zeolite A in the rst day, continuing
with an extended induction period of 4e6 days with the conversion
of gaylussite to CaCO3 and the formation of hydrotalcite, followed
by the precipitation of C-A-S-H gel. The joint activation of GGBS and

FA by Na2CO3, which has recently attracted more interest due to its


advantages in infrastructure applications (Gao et al., 2015), is still to
be examined.
The present paper aims to advance the knowledge of the
properties and reaction mechanisms of AACs based on Na2CO3. The
mechanical and hydration properties of the Na2CO3-based AAFS
system are the main focus of this study. The effect of Na2CO3
dosage, FA/GGBS ratio, and curing conditions are here investigated
by employing various techniques such as isothermal calorimetry, Xray diffraction (XRD), Fourier Transformed Infrared Spectroscopy
(FTIR), thermogravimetric analysis (TGA) and electron microscopy
(EM).

2. Materials and methods


2.1. Materials and sample preparation
The GGBS used was supplied by Hanson cement, UK, and has
basicity (Kb (CaO MgO)/(SiO2 Al2O3)) and hydration modulus
(HM (CaO MgO Al2O3)/SiO2) values of ~1.0 and ~1.60,
respectively. The GGBS was mainly amorphous, with merwinite
(Ca3Mg(SiO4)2) identied as the only crystalline phase present. FA
was obtained from Cemex, Rugby, UK and was classied to meet the
requirements of the British standard for use with PC (BS 3892: Part
1). The chemical compositions of the GGBS and FA are shown in
Table 1. Sodium carbonate, with a purity of >99%, was supplied in
powder form by Fisher scientic, UK.
The powder mixtures were prepared from GGBS and FA, and
activated by Na2CO3 solutions. The content of Na2CO3 was 5% or 10%
by weight of the powder mixtures. The Na2CO3 solutions were
prepared by dissolving the as-received Na2CO3 powder in predetermined deionized water until complete dissolution. All mixes
had a w/b ratio of 0.31. Table 2 shows the mix constituents and their
nomenclatures.
For the preparation of the paste samples, all the dry materials
(GGBS, FA) were mixed by hand in the bowl followed by 5 min dry
mixing in a bench-top mixer to which the Na2CO3 solution was then
added. The mixer was stopped after 3 min of slow mixing to collect
any unmixed solids scraped from the sides of the mixing bowl and
the paddle into the bowl. Then 2 more minutes of slow mixing and
5 min of fast mixing were applied. For each mix, the freshly mixed
cement paste was placed into 40  40  40 mm steel cubic moulds
in three layers, and in between each layers the mixture was tapped
with a spatula for at least 25 times in two directions to remove air
voids. The samples were demoulded after 2 days of curing and then
cured in two regimes until the testing age: 1) water tank at a
temperature of 20  C 2  C and 2) sealed in a cling lm at a
temperature of 20  C 2  C.

Table 1
Chemical composition and physical characteristics of the materials used (based on
the suppliers' datasheets).
Component

GGBS

FA

CaO%
SiO2%
Al2O3%
Fe2O3%
MgO%
SO3%
K2O%
Na2O%
SSA (m2/kg)
L.O.I

39.24
36.79
11.51
0.42
8.10
1.03
0.63
0.37
545
0.63

6.8 3.6
49.3 6.2
24.1 0.4
9.7 1.3
1.1 0.2
3.3 1.3
3.5 0.3
1.2 0.1
2600
3.9a

Obtained from (Liska and Al-Tabbaa, 2009).

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

Table 2
The mix proportions used in this study.
Mix ID

GGBS%

FA%

Na2CO3%

G100F0N5 (-Sa)
G75F25N5 (-S)
G50F50N5 (-S)
G100F0N10 (-S)
G75F25N10 (-S)
G50F50N10 (-S)

100
75
50
100
75
50

0
25
50
0
25
50

5
5
5
10
10
10

For sealed curing (-S) is added for each mix ID.

2.2. Tests
The compressive strength testing was carried out using Controls
Advantest 9 with a maximum capacity of 250 kN and with a loading
rate of 2400 N/s. Three identical cubes were tested at different
periods of ageing, namely 3, 7, 28, 56, and 90 days. The cube
strength reported herein was an average of three specimens at each
period of ageing. Immediately after the compressive strength test
was performed at 28-day age, selected samples for microstructural
analyses were immersed in acetone for three days in order to stop
any further hydration. Then the samples were ltered to remove
the acetone followed by vacuum drying in a desiccator and oven
drying at 60  C for at least 24 h. Thereafter, parts of each sample
were crushed and ground in an agate mortar until passing 75 mm
sieve and the sealed in plastic vials for XRD, FTIR, and TGA analysis.
Isothermal calorimetry experiments were conducted using a
TAM Air Isothermal calorimeter, at a base temperature of
20 0.02  C. Fresh paste was mixed externally, weighed into an
ampoule, and immediately placed in the calorimeter, and the heat
ow was recorded for the rst 140 h of reaction except for the
sample G100F0N5 where the data was recorded for 72 h as it was
done as a preliminary sample before deciding to extend the
observation period to 140 h. All values of the heat release rate were
normalised by the total weight of the paste.
Powder XRD was employed to identify the crystalline phases in
the sample. The powders were placed on glass microscope slides
onto which acetone was dripped. After the acetone evaporated,
the sample was afxed to the slide and placed in the Siemens
D500 X-ray diffractometer with a CuKa source operating at 40 kV
and 40 mA, emitting radiation at a wavelength of 1.5405 . The
scanning regions were between 2q values of 5e60 , at a rate of
0.02 /step. TGA experiments (Perkin Elmer STA6000, Simultaneous Thermal Analyser) were conducted using 20 2 mg of
powder under static air in an open alumina crucible heated at
10  C/min over the range of 40e1000  C. Temperature and weight
change, DW, (mg) were recorded every 0.1 s. FTIR spectra of the
samples were taken using Perkin Elmer FTIR Spectrometer Spectrum 100 Optica. Spectra were collected in transmittance mode
from 4000 to 600 cm1 at a resolution of 1 cm1. Additionally,
backscattered electron microscopy (BSE) and energy dispersive Xray analysis (EDX) were carried out on the 28-day samples using
FEI Nova NanoSEM FEG at 15 kV accelerating voltage and a
working distance of 5 mm. The samples were impregnated in
epoxy resin prior to polishing down to 0.25 mm and thereafter
coated with carbon.

Fig. 1. The compressive strength of mixes cured under water.

strength in the rst 7 days and then the strength gradually


increased with the curing time. Increasing the activator dosage
from 5% to 10% increased the strength of all the mixes. This indicates that the alkali (Na2O) content plays a vital role in the
development of mechanical strength. The improvement can be
attributed to the higher content of Na2O leading to an increase in
the pH of the system. With increased pH, the dissolution of reactive
species (SiO2, Al2O3) increases. This arises mainly from the slag but
also partly from y ash. The increased dissolution in turn increases
the extent of the reaction and the strength. Compressive strengths
of >60 MPa can be achieved after 28 days at room temperature by
using only Na2CO3 as the alkali source; this is in agreement with Li
and Sun (2000) who found that the mechanical properties of
Na2CO3-activated slag improved with increasing the dosage of
Na2CO3 up to 10%, at which the compressive strength effectively
developed from 0 MPa at 3 days to 60 MPa at 28 days at 0.38 solution to binder ratio (Li and Sun, 2000). However, in our study, the
compressive strengths after 3 days were considerably higher even
when including FA in the system.
The inclusion of 25% FA (circle shape in Figs. 1 and 2) gave
comparable or even higher strength than mixes containing only
slag, after 90 days of curing. This may indicate that FA reaction
started to contribute to the formation of the strength-giving phases

3. Results and discussion


3.1. Compressive strength
The compressive strengths of the mixes cured under water and
in sealed curing are shown in Figs. 1 and 2, respectively. It is shown
that under both curing conditions, the mixes gained substantial

Fig. 2. The compressive strength of mixes in sealed curing.

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

at late ages more than in the early stages due to its low reactivity.
The long-term improvement in mechanical strength by the inclusion of FA was also reported in the literature (Chi and Huang, 2013;
Guerrieri and Sanjayan, 2010; Ismail et al., 2013b; Puertas et al.,
2000). However, increasing the content of FA to 50% (triangle
shape) decreased the strength signicantly particularly for samples
cured under water at all ages and in sealed curing at early ages. For
samples with FA/slag ratio of 50/50 and cured under water, the
strength was very low due to the slow reaction of such system with
Na2CO3. A possible explanation for this might be that the leaching
of alkali ions in water delayed the reaction kinetics and decreased
the extensive reaction of the precursors. The decrease in mechanical strength due to the increased content of FA in AAS systems,
activated by different activators, has been observed in the literature
by many researchers (Escalante Garca et al., 2006; Kim and Kim,
2012; Kumar et al., 2010; Li and Sun, 2000; Puertas et al., 2003;
Shen et al., 2011; Shi and Day, 1999; Wang et al., 1994). For
example, Li and Sun (2000) showed that inclusion of more than 10%
of FA led to a drop in the compressive strength for mixtures activated by 5% Na2CO3 combined with 5% NaOH. On the contrary,
some researchers found either insignicant change or positive
change on the strengths by adding up to 50 wt.% FA (Chen et al.,
2015; Chi and Huang, 2013; Collins and Sanjayan, 1999; Guerrieri
and Sanjayan, 2010; Shi and Day, 1999). The discrepancies in
these studies correspond to the differences in the activator type and
concentration, the aluminosilicate materials' physical and chemical
properties, and the curing conditions.
Fig. 3 shows the 28-day compressive strength of all mixes in
both curing conditions. The main feature is that the sealed curing
led to higher compressive strength than curing under water. The
most notable improvement was observed for mixes including 50%
FA. Samples G50F50N5-S and G50F50N10-S had ~4 times and 5
times higher strength than G50F50N5 and G50F50N10, respectively. The possible explanation for the improved strength by sealed
curing is that it prevents the ions from leaching into the water, thus
enhancing the reaction rate and leading to more strength-giving
products. Comparison between these types of curing has not previously been reported in the literature. However, it was suggested
that sealed curing of AAFS can yield better strength over underwater curing (Lloyd, 2008).
3.2. Isothermal calorimeter
The heat release curves, normalized by the mass of starting
materials, of mixes containing 0% and 25% y ash at both activator

Fig. 3. 28-day compressive strengths of all mixes in both curing conditions.

Fig. 4. Heat release rate (top) and cumulative heat release (bottom) of mixes.

dosages are shown in Fig. 4. There was an initial pre-induction


period in the rst hour, associated with the partial dissolution of
mainly the slag and marginally of the FA. At 5% Na2CO3 the addition
of FA did not cause a signicant change on the initial dissolution.
However the peak of this period decreased with adding 25% FA at
10% Na2CO3 due to the reduction of the slag content and the low
reactivity of FA. This period was then followed by a long induction
period where little heat evolution occurred. This was also observed
nez et al.,
in the literature (Bernal et al., 2014; Fern
andez-Jime
1998); however, both of these studies reported different times
nez et al. (1998) or
which could be earlier such as in Fern
andez-Jime
later as in Bernal et al. (2014). These discrepancies could be
attributed to the differences in both physical and chemical properties of the slags used. In our study, the slag is ner than that in
Bernal et al. (2014).
It is clear that increasing the activator dosage shortened this
period while increasing the FA content delayed this period. On one
hand, increasing the activator concentration led to higher pore
solution pH, thereby increasing the dissolution and reaction rates:
on the other hand, introducing FA to the system diluted the reactive
species and hence decreased both the dissolution and reaction rate.
After the induction period, another high intensity heat evolution
process between 20e50 h and 30e60 h in neat slag binders activated by 10% Na2CO3 and 5% Na2CO3, respectively, was identied.
This peak corresponds to the acceleration and deceleration processes when the precipitation of voluminous reaction products
occurs, thereby releasing a signicant heat of reaction. These

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

observations are common for mixes containing 25% FA. The difference between the two systems was the time at which the second
peak appeared. The intensity of this peak and the starting time
depend strongly on the properties of the slag, the activator dosage,
and the mineral additives (Abdalqader et al., 2015).

3.3. X-ray diffraction


Figs. 5e6 present the crystalline products of selected mixes at
different curing times and conditions. It is found that the peak intensity is dependent on FA/slag ratio and the time of curing. In
mixes with only slag (Fig. 5a), the main reaction product is shown
near 29.5 2q, resembling the diffraction pattern of a disordered CS-H with a riversideite type structure, and/or calcite. C-S-H is
generally considered to be poorly crystalline but its crystallinity in
alkali-activated slag has already been reported by (Shi et al., 2006).
However, calcite occurrence is possible due to the recarbonation of
Ca from absorbing CO2 from CO2
3 ions as reported by other researchers (Sakulich, 2009; Xu et al., 2008) along with other calcium
carbonate polymorphs such as vaterite and aragonite (Bernal et al.,
2014).
Gaylussite (Na2Ca(CO3)2$5H2O) was detected as one of the main
crystalline phases. Gaylussite is known to form as a natural evaporite in alkali lake waters (Mees et al., 1998). The formation of such
phases implies that at early ages there is a preferential reaction
2
between the dissolved CO2
released from the partial
3 and the Ca
dissolution of the slag. However, this phase is unstable and it

Fig. 6. Effect of activator dosage and curing condition on XRD for some mixes at 28
days.

transforms to other phases with curing time as its peaks diminished with time. A low intensity peak at 11.7 2q is also observed,
assigned to hydrotalcite (Mg6Al2(CO3) (OH)16$4(H2O), which has
also previously been identied in other alkali-activated slag systems (Haha et al., 2011; Puertas and Fernandez-Jimenez, 2003).
This phase is very stable and its content increased with time as
shown in Fig. 5 (see increased peaks' intensities at 11.7 2q and 23
2q).
In the mixes with 25 wt.% FA (Fig. 5b), a poorly ordered C-S-H
type gel is also observed alongside calcite, hydrotalcite, and gaylussite. However, the peak intensity of hydrotalcite phase is much
smaller in this mix than the neat slag binders. The crystalline phase
quartz, from the unreacted y ash, was also identied. The peak of
gaylussite decreased with time and almost disappeared after 180
days indicating the transformation of these phases into other
phases. Additionally, the peak of quartz between 28 days and 180
days also decreased which may correspond to the activation of y
ash.
Fig. 6 shows the effect of activator content and curing condition
on the hydration products formed. In general, same hydration
products were detected in the mixes containing 25% FA regardless
of activator dosage and curing condition. However, it was observed
that increasing the Na2CO3 dosage from 5% to 10% facilitated the
formation of gaylussite due to the higher availability of Na and
CO2
3 ions. Furthermore, sealed curing also increased the intensity
of this phase, which can be attributed to its prevention of these ions
from leaching into the water as mentioned above. This agrees well
with the strength results. The same trend was observed for the
formation of calcite. Further evidence of enhanced hydration by
sealed curing was manifested by the reduction in the quartz peak
compared to the water curing, which indicates the increased
dissolution of FA and hence increased amounts of hydration
products.
3.4. Thermogravimetric analysis

Fig. 5. XRD for (a) G100F0N10 and (b) G75F25N10 at different ages.

Fig. 7 showed the TG curves of G100F0N10 at different ageing


periods. Other samples at different ageing periods are not presented here because they have the same distinctive features. The
comparison between different samples after 28 days of curing is
depicted in Fig. 8. Figs. 7e8 show that there were mainly four
humps in the DTG curves, in which the rst one was at 85e105  C
and is attributed to the removal of free evaporable water which is
present in the pores of the geopolymer gel products, either C-(A)-S-

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

Fig. 7. TGA analysis of mix G100F0N10 at different ages.

H type or N-A-S-H (zeolite-like) gels (Ismail et al., 2013a). The peak


between 300  C and 400  C is due to the decomposition of
hydrotalcite (Wang et al., 1994). The third peak observed at
500e600  C could be due to either the dehydration of thomsonite
ldv
(Fo
ari, 2011), M-S-H gel (Jin et al., 2014), or the decomposition of
poorly crystallined phase of calcium carbonate (Ismail et al., 2013b;
Sakulich, 2009). The temperature range of 600e800  C is the
decomposition range of various carbonate-containing phases
including hydrotalcite, magnesium carbonate, and calcium carbonate (Jin et al., 2014). These results support the XRD results of the
crystalline hydration products. The diminishing of the peak at
500e600  C indicates that this phase was transformed with
extended curing to other phases. For instance, the low crystalline
calcium carbonate (e.g. vaterite) phases may slowly convert to
calcite, which is more stable (Bernal et al., 2014).
As observed in the XRD analysis the main reaction products
were identical in all mixes regardless of the curing regime used or
the inclusion of 25% FA. However, the reaction degree was dependent on these factors as indicated from the different mass losses of
the same phases for different mixes. Sealed curing was found to
enhance the reaction degree as shown by the increased mass loss
from C-S-H decomposition from 7.44% for G100F0N10 to 8.41% for
G100F0N10-S and from 6.25% for G75F25N10 to 8.19% for
G75F25N10-S and the total mass loss from 16.24% for G100F0N10 to
17.23% for G100F0N10-S and from 16.25% for G75F25N10 to 17.12%

Fig. 8. TGA analysis of different mixes at 28 days.

for G75F25N10-S. Increasing the activator dosage had a similar


effect on the reaction degree as indicated by the increased mass loss
from C-S-H decomposition from 5.68% for G75F25N5 to 6.25% for
G75F25N10 and the total mass loss from 12.4% to 16.25% for
G75F25N5 and G75F25N10, respectively.
3.5. FTIR analysis
The FTIR spectra of unreacted precursor materials used in this
study and some samples at 28-day age are presented in Fig. 9.
Although the measurement were performed between 600 and
4000 cm1, only up to 2000 cm1 were presented in Fig. 10 as there
is no distinctive feature beyond this wavelength. For the raw materials, the band between 1100 and 900 cm1 is associated with the
asymmetric stretching vibration mode of Si-O-T bonds
(T tetrahedral Al, Si). This band position changed from 900 cm1
in the GGBS to 1010 cm1 in the FA, due to their different glass
structures. The shoulder at 875 cm1 identied in the GGBS can be
attributed to the asymmetric stretching of AlO4 groups in the glass
phases. Additionally, the band at 1450 cm1 identied in the GGBS
is assigned to the symmetric stretching mode of the O-C-O bonds of
CO2
3 groups associated with supercial weathering of the GGBS
during storage.
After activation, all samples show identical bands, suggesting a
very similar nature of hydration products irrespective of the activator dosage and FA content. Major bands systems were detected at

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

the v3 (SieO) vibration in the IR spectra of C-S-H gels is at


950e994 cm1 (Garcia-Lodeiro et al., 2011; Puertas and FernandezJimenez, 2003). The position of this band is consistent with both the
C-(A)-S-H structure formed by the activation of slag in alkaline
media (Garcia-Lodeiro et al., 2011; Puertas and Fernandez-Jimenez,
2003), and the N-A-S-H gels formed in geopolymer systems derived
from FA (Rees et al., 2007).
In samples with 25 wt% FA, the Si-O-T band shifted slightly towards higher wavenumbers compared to the plain slag samples.
This is possibly due to the activation of y ash, which increases the
formation of more polymerized gel. This is consistent with ndings
reported by (Ismail et al., 2014).
3.6. Scanning electron microscopy

Fig. 9. FTIR spectra of unreacted precursors and some mixes at 28 days.

approximately 1650, 1410, 950, and 870 cm1. The structure of


molecular water in the alkali activated FA/slag system is characterized by the OeH stretching band, from 3200 to 3700 cm1, while
the bending of the chemically bonded HeOeH is located at
1650 cm1 (Ismail et al., 2013a). The existence of bands at 1410 and
870 cm1 suggests the presence of CO2
3 , which can be attributed to
the presence of calcite or hydrotalcite as detected by both XRD and
TGA. The major band, at ~950 cm1, corresponds to the SieO
asymmetric stretching bands in Q2 units. It is well known that

Backscattered electron microscopy (BSEM) images of selected


28-day cured samples at 2500 magnication are presented in
Fig. 10. These samples represent different FA/slag ratios and activator dosages. The mix G50F50N10-S was chosen because it was
difcult to prepare the same mix cured under water. Both unreacted slag (angular light gray particles) and unreacted FA (spherical
light gray particles) are embedded within the binding matrix.
Cracks are observed in some samples, which are likely to have
occurred during sample preparation.
The main binding phase (continuous gray regions) in the images
corresponds to C-A-S-H gel, as previously identied via XRD, TGA
and FTIR. In Fig. 10a, it is likely that slag with small particle size had
been fully hydrated while the larger particles were not completely

Fig. 10. BSEM images at 28 days of mixes (a) G100F0N10, (b) G75F25N5, (c) G75F25N10, (d) G50F50N10-S.

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

hydrated yet with clear hydration rims around. Likewise, Fig. 10bed
shows the same trend for slag particles. Hydration rims also were
formed around the slag and FA particles, particularly for mixes
activated by 10% Na2CO3. These reaction products grew gradually
from the outer area into the inner part of the unreacted particles. FA
particles have been embedded in the matrix (Fig. 10bed) so, when
unreacted, they can play a space-lling role. This may explain the
unchanged strength in samples containing up to 25% FA. However,
when the content of FA increased (Fig. 10d), this affected the reaction mechanism and led to strength reduction due to the low
reactivity of FA.
To determine the elemental composition of the hydration
products, EDX was performed on at least 20 points selected on the
gels and some of the slag grains on samples cured for 28 days at a
magnication of 2500. The data are presented here in a ternary
composition graph of CaeSieAl (Fig. 11a) and in a Al/Si vs Mg/Si
graph (Fig. 11b). These plots clearly show the variation in gel
composition across the different AAFS blends as a function mainly
of FA/slag ratio and to a less degree of activator dosage and curing
conditions. The main binding phase in neat slag binders is C-A-S-H
gel (dashed circle in Fig. 11a), in agreement with what have been
reported in the literature (Bernal et al., 2014; Deir et al., 2014;
Escalante-Garca et al., 2003; Ismail et al., 2013a; Puertas and
Fernandez-Jimenez, 2003). In samples containing 25% FA, a binding phase of both C-A-S-H and C-(N)-A-S-H is likely to be formed,
while increasing the FA content to 50% led to the formation of C(N)-A-S-H as the major binding-phase (dotted circle in Fig. 11a).
Again, this trend has been observed in other studies (Ismail et al.,
2014; Marjanovic et al., 2015; Puertas and Fernandez-Jimenez,
2003; Puertas et al., 2000; Puligilla and Mondal, 2015; Zhao et al.,
2007) using various alkali activators. It is suggested that the use of a
higher activator concentration enhances the incorporation of Al
into the gel structure when introducing FA (Lloyd et al., 2010). The
main modication induced in the current study mixtures is the
incorporation of carbonate-bearing products such as gaylussite and
calcite.
In Fig. 11b, the approximate linear correlation of Mg/Si with Al/Si
indicates the existence of other phases rich in aluminum and
magnesium such as hydrotalcite-like phases, which was also
detected by XRD and TGA and previous studies (Bernal et al., 2014;
Escalante-Garca et al., 2003; Ismail et al., 2014; Jin et al., 2014;
Puertas and Fernandez-Jimenez, 2003). The presence of a positive
x-axis intercept reveals the level of incorporation of Al in the C-S-H
(Table 3).
The average chemical compositions of the gels in some samples
at 28 days are shown in Table 3. The Ca/Si ratio decreased with
increasing FA due to the reduction of Ca2 ions resulted from the
partial replacement of GGBS. Al/Si values varied between 0.38 and
0.43, which is very high for either a pure chain-structured C-A-S-H
phase (Al/Si < 0.2 (Hunnicutt, 2013)) or considerable degree of
crosslinking (0.19  Al/Si  0.26) (Myers et al., 2013). Therefore, it
may correspond to the presence of additional Al-rich products
intermixed with Al-substituted C-S-H gel (Bernal et al., 2014). The
range of Na/Si in the investigated samples was from 0.14 to 0.53 as
shown in Table 3. The role of Na in the structure of the reaction
products is to balance the negative framework charge induced by
the incorporation of Al (Duxson et al., 2006; Puertas et al., 2000) as
Na Al3 Si4 (Hong and Glasser, 2002). The curing conditions
did not substantially affect the elemental composition of pastes.
3.7. Proposed reaction mechanism
Combining the results of mineralogical and microstructural
analysis, this study clearly showed that the reaction mechanism of
Na2CO3-activated FA/slag is dependent on the activator dosage,

(a)

(b)

Fig. 11. EDX analysis of some mixes at 28 days (a) ternary plot of CaeAleSi, (b) atomic
ratios Al/Si vs Mg/Si.

FA/slag ratio and curing condition. By adapting the reaction process


of Na2CO3-activated slag proposed in (Bernal et al., 2014), the reaction mechanism of Na2CO3-activated FA/slag can be derived from
the current study, taking into account the different slags used in
both studies. Firstly, an initial dissolution of the slag and y ash, to a
lesser degree, takes place, accompanied by heat release. Then the
reaction goes through an induction period starting from 10 h and
extends up to 40 h depending on the FA/slag ratio and the activator
dosage. Compared to Bernal et al. (2014), it is suggested that the
heat release and the length of this induction period are dependent
on the reactivity of the slag. During these two stages, an increase in
the concentration of dissolved OH-, Si, and Al occurs and different
reaction products such as calcite, gaylussite, and hydrotalcite are
formed (Jin et al., 2015; Moseson, 2011). Also the precipitation of
C-A-S-H or C-(N)-A-S-H gel starts. The reaction stoichiometry is
dominated by GGBS as it is more reactive than FA. In the last stage
(2e3 days onwards), the precipitation of C-A-S-H or C-(N)-A-S-H
2
gel increases. The consumption of CO2
3 causes reduction of CO3
concentration leading to an increase in Ca2 concentration, thereby
promoting further precipitation of the gel and reducing the

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

Table 3
Calculated parameters from EDS results at 28 days.
Sample

Ca/Si

Al/Si

Mg/Si

Na/Si

Mg/Al (calculated from Fig. 11b)

Al substitution

Slag
G100F0N10
G75F25N5
G75F25N10
G75F25N10-S
G50F50N10-S

1.19
1.07
0.95
0.83
0.71
0.63

0.40
0.40
0.44
0.43
0.38
0.41

0.36
0.25
0.35
0.30
0.27
0.18

0.09
0.14
0.42
0.53
0.31
0.33

0.7
1.90
1.34
1.94
1.28
0.40

e
0.27
0.18
0.28
0.18
NA

NA not applicable.

porosity. Additionally, Mg2 ions continue to release and react with


dissolved Al to form hydrotalcite.
4. Conclusion
The aim of this paper was to investigate the properties of
Na2CO3-activated y ash/slag pastes in terms of mechanical
strength, reaction kinetics, reaction products, and microstructure.
The key inuencing factors on the properties of this binder were
scrutinized: activator dosage, y ash/slag ratio, and curing conditions. The main ndings of this paper are summarized as follows:
1. Sodium carbonate can be used to effectively activate mixtures of
FA and GGBS. The optimum mixture (10% sodium
carbonate 25% FA 75% slag) can yield strength up to
60 MPa at 28 days and up to 80 MPa at 90 days.
2. The most important factors affecting the strength development
were the activator dosage and FA/slag ratio. Increasing the
activator dosage led to an increase in the strengths at all ages
while increasing FA content by more than 25 wt% resulted in a
signicant drop in strength when cured under water and a slight
reduction in sealed curing.
3. With the exception of mixes containing 50% FA, the water and
sealed curing made no signicant difference to the strength
development.
4. The remarkable inuence of activator dosage and FA/slag ratio
was also manifested in the reaction kinetics and the microstructure of the cement paste.
5. The main reaction product was found to be a poorly crystalline
calcium silicate hydrate rich in Al and includes Na into its
structure. The uptake of Al increased with increasing FA contents. Secondary products due to the role of carbonate were
formed, such as hydrotalcite, calcite, and gaylussite.
Acknowledgements
The nancial support of the PhD scholarship for the rst author
from the Yousef Jameel Foundation and Cambridge Overseas Trust
are gratefully acknowledged.
References
Abdalqader, A., Al-Tabbaa, A., 2015. Sustainable binder based on sodium carbonateactivated y ash/slag and reactive magnesia. In: 2015 International Concrete
Sustainability Conference. National Ready Mixed Concrete Association,
pp. 1e15.
Abdalqader, A., Al-Tabbaa, A., 2014a. Hydration and mechanical properties of
reactive magnesia and sodium carbonate-activated y ash/slag paste blends. In:
The RILEM PRO092 Proceeding of the Second International Conference on Advances in Chemically-activated Materials (CAM'2014-China), June 1e3, 2014,
Changsha, China. RILEM Publications S.A.R.L., Changsha, China, pp. 269e280.
Abdalqader, A., Al-Tabbaa, A., 2014b. Factors affecting the properties of Na2CO3activated y ash/slag paste. In: The 34th Annual Cement and Concrete Science
Conference. University of Shefeld, Shefeld, UK, pp. 329e332.
Abdalqader, A., Jin, F., Al-Tabbaa, A., 2015. Characterisation of reactive magnesia and
sodium carbonate-activated y ash/slag paste blends. Constr. Build. Mater. 93,
506e513. http://dx.doi.org/10.1016/j.conbuildmat.2015.06.015.

Bernal, S., Provis, J.L., Myers, R.J., San Nicolas, R., van Deventer, J.S.J., 2014. Role of
carbonates in the chemical evolution of sodium carbonate-activated slag
binders. Mater. Struct. 1e13. http://dx.doi.org/10.1617/s11527-014-0412-6.
, L., Morera, J.M., Bartol, E., Ine
s Ferna
ndez, A.,
Cabeza, L.F., Barreneche, C., Miro
2013. Low carbon and low embodied energy materials in buildings: a review.
Renew. Sustain. Energy Rev. 23, 536e542. http://dx.doi.org/10.1016/
j.rser.2013.03.017.
Chen, C.-T., Nguyen, H.-A., Chang, T.-P., Yang, T.-R., Nguyen, T.-D., 2015. Performance
and microstructural examination on composition of hardened paste with nocement SFC binder. Constr. Build. Mater. 76, 264e272. http://dx.doi.org/
10.1016/j.conbuildmat.2014.11.032.
Chi, M., Huang, R., 2013. Binding mechanism and properties of alkali-activated y
ash/slag mortars. Constr. Build. Mater. 40, 291e298. http://dx.doi.org/10.1016/
j.conbuildmat.2012.11.003.
Collins, F., Sanjayan, J., 1999. Effects of ultra-ne materials on workability and
strength of concrete containing alkali-activated slag as the binder. Cem. Concr.
Res. 29, 459e462. http://dx.doi.org/10.1016/S0008-8846(98)00237-3.
Deir, E., Gebregziabiher, B.S., Peethamparan, S., 2014. Inuence of starting material
on the early age hydration kinetics, microstructure and composition of binding
gel in alkali activated binder systems. Cem. Concr. Compos 48, 108e117. http://
dx.doi.org/10.1016/j.cemconcomp.2013.11.010.
Deventer, J.S.J., Provis, J.L., Duxson, P., Brice, D.G., 2010. Chemical research and
climate change as drivers in the commercial adoption of alkali activated materials. Waste Biomass Valoriz. 1, 145e155. http://dx.doi.org/10.1007/s12649010-9015-9.
nez, A., Provis, J.L., Lukey, G.C., Palomo, A., Deventer, J.S.J.,
Duxson, P., Fern
andez-Jime
2006. Geopolymer technology: the current state of the art. J. Mater. Sci. 42,
2917e2933. http://dx.doi.org/10.1007/s10853-006-0637-z.
Escalante Garca, J.I., Campos-Venegas, K., Gorokhovsky, A., Fern
andez, A., 2006.
Cementitious composites of pulverised fuel ash and blast furnace slag activated
by sodium silicate: effect of Na2O concentration and modulus. Adv. Appl. Ceram.
105, 201e208. http://dx.doi.org/10.1179/174367606X120151.
Escalante-Garca, J.I., Fuentes, A.F., Gorokhovsky, A., Fraire-Luna, P.E., MendozaSuarez, G., 2003. Hydration products and reactivity of blast-furnace slag activated by various alkalis. J. Am. Ceram. Soc. 86, 2148e2153. http://dx.doi.org/
10.1111/j.1151-2916.2003.tb03623.x.
ndez-Jime
nez, A., Palomo, A., 2005. Composition and microstructure of alkali
Ferna
activated y ash binder: effect of the activator. Cem. Concr. Res. 35, 1984e1992.
http://dx.doi.org/10.1016/j.cemconres.2005.03.003.
ndez-Jime
nez, A., Palomo, J.G., Puertas, F., 1999. Alkali-activated slag mortars
Ferna
mechanical strength behaviour. Cem. Concr. Res. 29, 1313e1321.
ndez-Jime
nez, A., Puertas, F., Arteaga, A., 1998. Determination of kinetic
Ferna
equations of alkaline activation of blast furnace slag by means of calorimetric
data. J. Therm. Anal. 52, 945e955.
ldv
Fo
ari, M., 2011. Handbook of the Thermogravimetric System of Minerals
and its Use in Geological Practice, 213th ed. Central European Geology.
Geological Institute of Hungary, Budapest. http://dx.doi.org/10.1556/
CEuGeol.56.2013.4.6.
Gao, X., Yu, Q.L., Brouwers, H.J.H., 2015. Reaction kinetics, gel character and strength
of ambient temperature cured alkali activated slagey ash blends. Constr. Build.
Mater. 80, 105e115. http://dx.doi.org/10.1016/j.conbuildmat.2015.01.065.
ndez-Jime
nez, a., Macphee, D.E., 2011. CompatGarcia-Lodeiro, I., Palomo, a., Ferna
ibility studies between N-A-S-H and C-A-S-H gels. Study in the ternary diagram
Na2OeCaOeAl2O3eSiO2eH2O. Cem. Concr. Res. 41, 923e931. http://dx.doi.org/
10.1016/j.cemconres.2011.05.006.
Guerrieri, M., Sanjayan, J.G., 2010. Behavior of combined y ash/slag-based geopolymers when exposed to high temperatures. Fire Mater. 34, 163e175. http://
dx.doi.org/10.1002/fam.
Habert, G., d'Espinose de Lacaillerie, J.B., Roussel, N., 2011. An environmental
evaluation of geopolymer based concrete production: reviewing current
research trends. J. Clean. Prod. 19, 1229e1238. http://dx.doi.org/10.1016/
j.jclepro.2011.03.012.
Haha, M. Ben, Lothenbach, B., Le Saout, G., Winnefeld, F., 2011. Inuence of slag
chemistry on the hydration of alkali-activated blast-furnace slag d part I: effect
of MgO. Cem. Concr. Res. 41, 955e963. http://dx.doi.org/10.1016/
j.cemconres.2011.05.002.
Heidrich, C., Feuerborn, H., Weir, A., 2013. Coal combustion products: a global
perspective. In: World of Coal Ash (WOCA) Conference. Lexington, p. 17.
Higgins, D., Sear, L., King, D., Price, B., Barnes, R., Clear, C., 2011. Cementitious Materials: the Effect of GGBS, Fly Ash, Silica Fume and Limestone Fines on the
Properties of Concrete. Surrey.

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

10

A.F. Abdalqader et al. / Journal of Cleaner Production xxx (2015) 1e10

Hong, S.Y., Glasser, F.P., 2002. Alkali sorption by C-S-H and C-A-S-H gels: part II. Role
of alumina. Cem. Concr. Res. 32, 1101e1111. http://dx.doi.org/10.1016/S00088846(02)00753-6.
Horpibulsuk, S., Munsrakest, V., Udomchai, A., Chinkulkijniwat, A., Arulrajah, A.,
2014. Strength of sustainable non-bearing masonry units manufactured from
calcium carbide residue and y ash. Constr. Build. Mater. 71, 210e215. http://
dx.doi.org/10.1016/j.conbuildmat.2014.08.033.
Hunnicutt, W., 2013. Characterization of Calcium-Silicate-Hydrate and Calcium-Alumino-Silicate-Hydrate. University of Illinois at Urbana-Champaign (Master thesis).
Ismail, I., Bernal, S. a., Provis, J.L., Hamdan, S., Deventer, J.S.J., 2013a. Microstructural
changes in alkali activated y ash/slag geopolymers with sulfate exposure.
Mater. Struct. 46, 361e373. http://dx.doi.org/10.1617/s11527-012-9906-2.
Ismail, I., Bernal, S. a., Provis, J.L., San Nicolas, R., Hamdan, S., van Deventer, J.S.J.,
2014. Modication of phase evolution in alkali-activated blast furnace slag by
the incorporation of y ash. Cem. Concr. Compos. 45, 125e135. http://
dx.doi.org/10.1016/j.cemconcomp.2013.09.006.
Ismail, I., Bernal, S., Provis, J., San Nicolas, R., Brice, D.G., Kilcullen, A.R., Hamdan, S.,
van Deventer, J.S.J., 2013b. Inuence of y ash on the water and chloride
permeability of alkali-activated slag mortars and concretes. Constr. Build. Mater.
48, 1187e1201. http://dx.doi.org/10.1016/j.conbuildmat.2013.07.106.
Jin, F., Gu, K., Al-Tabbaa, A., 2015. Strength and hydration properties of reactive
MgO-activated ground granulated blastfurnace slag paste. Cem. Concr. Compos.
57, 8e16. http://dx.doi.org/10.1016/j.cemconcomp.2014.10.007.
Jin, F., Gu, K., Al-Tabbaa, A., 2014. Strength and drying shrinkage of reactive MgO
modied alkali-activated slag paste. Constr. Build. Mater. 51, 395e404. http://
dx.doi.org/10.1016/j.conbuildmat.2013.10.081.
Kim, H., Kim, Y., 2012. Characteristics of the geopolymer using y ash and blast
furnace slag with alkaline activators. In: 4th International Conference on
Chemical, Biological and Environmental Engineering. IACSIT Press, Singapore,
pp. 154e159. http://dx.doi.org/10.7763/IPCBEE.
Kostick, D.S., 2013. Soda Ash: Statistics and Information. U.S. Geological Survey.
Kumar, S., Kumar, R., Mehrotra, S.P., 2010. Inuence of granulated blast furnace slag
on the reaction, structure and properties of y ash based geopolymer. J. Mater.
Sci. 45, 607e615. http://dx.doi.org/10.1007/s10853-009-3934-5.
Li, Y., Sun, Y., 2000. Preliminary study on combined-alkali-slag paste materials. Cem.
Concr. Res. 30, 963e966.
Liska, M., Al-Tabbaa, A., 2009. Ultra-green construction: reactive magnesia masonry
products. Proc. ICE Waste Resour. Manag 162, 185e196. http://dx.doi.org/
10.1680/warm.2009.162.4.185.
Lloyd, R.R., 2008. The Durability of Inorganic Polymer Cements (PhD thesis). Department of Chemical and Biomolecular Engineering, The University of Melbourne.
Lloyd, R.R., Provis, J.L., Deventer, J.S.J., 2008. Microscopy and microanalysis of
inorganic polymer cements. 1: remnant y ash particles. J. Mater. Sci. 44,
608e619. http://dx.doi.org/10.1007/s10853-008-3077-0.
Lloyd, R.R., Provis, J.L., Van Deventer, J.S.J., 2010. Pore solution composition and
alkali diffusion in inorganic polymer cement. Cem. Concr. Res. 40, 1386e1392.
http://dx.doi.org/10.1016/j.cemconres.2010.04.008.
Marjanovi
c, N., Komljenovi
c, M., Bas
carevi
c, Z., Nikoli
c, V., Petrovi
c, R., 2015.
Physicalemechanical and microstructural properties of alkali-activated y
asheblast furnace slag blends. Ceram. Int. 41, 1421e1435. http://dx.doi.org/
10.1016/j.ceramint.2014.09.075.
Mees, F., Reyes, E., Keppens, E., 1998. Stable isotope chemistry of gaylussite and
nahcolite from the deposits of the Crater lake at Malha, northern Sudan. Chem.
Geol. 146, 87e98. http://dx.doi.org/10.1016/S0009-2541(98)00006-0.
rrez, R., Gallego, N., 2015. Preparation and
Meja, J.M., Rodrguez, E., Meja de Gutie
characterization of a hybrid alkaline binder based on a y ash with no commercial
value. J. Clean. Prod.104, 346e352. http://dx.doi.org/10.1016/j.jclepro.2015.05.044.
Moseson, A.J., 2011. Design and Implementation of Alkali Activated Cement for
Sustainable Development (PhD thesis). Drexel University.
Myers, R.J., Bernal, S., San Nicolas, R., Provis, J.L., 2013. Generalized structural
description of calcium-sodium aluminosilicate hydrate gels: the cross-linked
substituted tobermorite model. Langmuir 29, 5294e5306. http://dx.doi.org/
10.1021/la4000473.
Nazari, A., Sanjayan, J.G., 2015. Synthesis of geopolymer from industrial wastes.
J. Clean. Prod. 99, 297e304. http://dx.doi.org/10.1016/j.jclepro.2015.03.003.
Nehdi, M., 2011. Ternary and quaternary cement for sustainable development.
Concr. Int. 23, 36e44.
Nehdi, M., Mindess, S., 1999. A quantitative approach to predicting the performance
of blended cements. Concr. Sci. Eng. 1, 205e214.

Neramitkornburi, A., Horpibulsuk, S., Shen, S., Arulrajah, A., Miri Disfani, M., 2015.
Engineering properties of lightweight cellular cemented clay-y ash material.
Soils Found. 55, 471e483. http://dx.doi.org/10.1016/j.sandf.2015.02.020.
Phetchuay, C., Horpibulsuk, S., Suksiripattanapong, C., Chinkulkijniwat, A.,
Arulrajah, A., Disfani, M.M., 2014. Calcium carbide residue: alkaline activator for
clay-y ash geopolymer. Constr. Build. Mater. 69, 285e294. http://dx.doi.org/
10.1016/j.conbuildmat.2014.07.018.
Provis, J.L., 2013. Geopolymers and other alkali activated materials: why, how, and
what? Mater. Struct. 47, 11e25. http://dx.doi.org/10.1617/s11527-013-0211-5.
Provis, J.L., van Deventer, J.S.J. (Eds.), 2014. Alkali Activated Materials, RILEM Stateof-the-Art Reports. Springer Netherlands, Dordrecht. http://dx.doi.org/10.1007/
978-94-007-7672-2.
ndez-Jime
nez, a., Va
zquez, T., 2003. Mechanical and
Puertas, F., Amat, T., Ferna
durable behaviour of alkaline cement mortars reinforced with polypropylene
bres. Cem. Concr. Res. 33, 2031e2036. http://dx.doi.org/10.1016/S00088846(03)00222-9.
Puertas, F., Fernandez-Jimenez, A., 2003. Mineralogical and microstructural characterisation of alkali-activated y ash/slag pastes. Cem. Concr. Compos. 25,
287e292.
Puertas, F., Martinez-Ramirez, S., Vazquez, T., Alonso, S., 2000. Alkali-activated y
ash/slag cement strength behaviour and hydration products. Cem. Concr. Res.
30, 1625e1632.
Puligilla, S., Mondal, P., 2015. Co-existence of aluminosilicate and calcium silicate
gel characterized through selective dissolution and FTIR spectral subtraction.
Cem. Concr. Res. 70, 39e49. http://dx.doi.org/10.1016/j.cemconres.2015.01.006.
Rashad, A.M., 2014. A comprehensive overview about the inuence of different
admixtures and additives on the properties of alkali-activated y ash. Mater.
Des. 53, 1005e1025. http://dx.doi.org/10.1016/j.matdes.2013.07.074.
Rashad, A.M., 2013. A comprehensive overview about the inuence of different
additives on the properties of alkali-activated slag e a guide for civil
engineer. Constr. Build. Mater. 47, 29e55. http://dx.doi.org/10.1016/
j.conbuildmat.2013.04.011.
Rees, C.A., Provis, J.L., Lukey, G.C., Deventer, J.S.J. Van, 2007. Attenuated total
reectance Fourier transform infrared analysis of y ash geopolymer gel aging.
Langmuir 23, 8170e8179.
Sakulich, A., 2009. Characterization of Environmentally-friendly Alkali
Activated Slag Cements and Ancient Building Materials (PhD thesis). Drexel
University.
Sarkar, S., Roumain, J.C., 2003. Role of cement science in sustainable development.
In: International Symposium Dedicated to Professor Fred Glasser. University of
Aberdeen. Thomas Telford, Dundee, pp. 45e57.
Scrivener, K.L., Kirkpatrick, R.J., 2008. Innovation in use and research on cementitious material. Cem. Concr. Res. 38, 128e136.
Shen, W., Wang, Y., Zhang, T., Zhou, M., Li, J., Cui, X., 2011. Magnesia modication of
alkali-activated slag y ash cement. J. Wuhan. Univ. Technol. Sci. Ed. 26,
121e125. http://dx.doi.org/10.1007/s11595-011-0182-8.
Shi, C., Day, R.L., 1999. Early strength development and hydration of alkali-activated
blast furnace slag/y ash blends. Adv. Cem. Res. 11, 189e196. http://dx.doi.org/
10.1680/adcr.1999.11.4.189.
Shi, C., Roy, D., Krivenko, P., 2006. Alkali-activated Cements and Concretes. Taylor &
Francis, Oxon.
Srinivasan, K., Sivakumar, a., 2013. Geopolymer binders: a need for future concrete
construction. ISRN Polym. Sci. 1e8. http://dx.doi.org/10.1155/2013/509185.
Suksiripattanapong, C., Horpibulsuk, S., Chanprasert, P., Sukmak, P., Arulrajah, A.,
2015. Compressive strength development in y ash geopolymer masonry units
manufactured from water treatment sludge. Constr. Build. Mater. 82, 20e30.
http://dx.doi.org/10.1016/j.conbuildmat.2015.02.040.
U.S. Geological Survey, 2013a. Iron and Steel Slag, Mineral Commodity Summaries.
U.S. Geological Survey, 2013b. Mineral Commodity Summaries: Soda Ash.
Wang, S., Scrivener, K.L., Pratt, P.L., 1994. Factors affecting the strength of alkaliactivated slag. Cem. Concr. Res. 24, 1033e1043.
Xu, H., Provis, J.L., Deventer, J.S.J. Van, Krivenko, P.V., 2008. Characterization of aged
slag concretes. ACI Mater. J. 102, 131e139.
Yang, K.-H., Song, J.-K., Song, K.-I., 2013. Assessment of CO2 reduction of alkaliactivated concrete. J. Clean. Prod. 39, 265e272. http://dx.doi.org/10.1016/
j.jclepro.2012.08.001.
Zhao, F.-Q., Ni, W., Wang, H.-J., Liu, H.-J., 2007. Activated y ash/slag blended
cement. Resour. Conserv. Recycl 52, 303e313. http://dx.doi.org/10.1016/
j.resconrec.2007.04.002.

Please cite this article in press as: Abdalqader, A.F., et al., Development of greener alkali-activated cement: utilisation of sodium carbonate for
activating slag and y ash mixtures, Journal of Cleaner Production (2015), http://dx.doi.org/10.1016/j.jclepro.2015.12.010

S-ar putea să vă placă și