You are on page 1of 36

Author’s Accepted Manuscript

Influence of NaOH concentration on microstructure


and properties of cured alkali-activated calcined
clay

Kenza El Hafid, Mohamed Hajjaji, Hassan El


Hafid
www.elsevier.com/locate/jobe

PII: S2352-7102(16)30113-9
DOI: http://dx.doi.org/10.1016/j.jobe.2017.04.012
Reference: JOBE251
To appear in: Journal of Building Engineering
Received date: 2 August 2016
Revised date: 14 April 2017
Accepted date: 20 April 2017
Cite this article as: Kenza El Hafid, Mohamed Hajjaji and Hassan El Hafid,
Influence of NaOH concentration on microstructure and properties of cured
alkali-activated calcined clay, Journal of Building Engineering,
http://dx.doi.org/10.1016/j.jobe.2017.04.012
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Influence of NaOH concentration on microstructure and properties of cured alkali-
activated calcined clay

Kenza El Hafid *, Mohamed Hajjaji, Hassan El Hafid

Laboratoire Physico-Chimie des Matériaux et Environnement, URAC20, Université Cadi Ayyad,

Faculté des Sciences Semlalia, Bd. Prince My Abdellah, B.P. 2390, 40001 Marrakech, Morocco.

Abstract

Microstructure of cured mixes composed of kaolinitic-illitic raw clay calcined at 700°C and

sodium hydroxide (4-14 M) was investigated using X-ray diffraction, Fourier-transform infrared

spectroscopy, Scanning electron microscopy and Thermal analysis. The change of the properties

(flexural strength, density, water absorption, open porosity and electrical conductivity) versus

NaOH concentration was evaluated and the possible formation of zeolite ZK-14 within the

geopolymer matrix from the amorphous clay phases (mainly metakaolinite) was examined. Sodium

carbonate was detected as a carbonation product. Due to heating, geopolymers were characterized

by dehydration and subsequent recrystallization into nepheline. The development of the fibrous-

zeolitic phase was favored by the increase of NaOH concentration and its impact on the studied

properties was somewhat advantageous. The rise of alkali concentration within 4-11.5 M enhanced

the extent of geopolymerization. The water absorption kinetic obeyed a pseudo-first order equation

and the constant was in the range (18.38–51.01) x 10-2 min-1.

1
Keywords: Calcined clay, Geopolymers, Microstructure, Physical properties, Deconvolution.

*Corresponding author: E-mail: elhafid.kenza@gmail.com (Kenza El Hafid).

2
1. Introduction

The mixture of a powdery aluminosilicate-based material and an alkaline activator has been

studied as the main chemical process giving rise to a new kind of eco-friendly binders commonly

known as geopolymers or alkali-activated cements [1-4]. They form an amorphous to semi-

crystalline 3-dimensional aluminosilicate framework structures by the combination of [SiO4]4- and

[AlO4]5- tetrahedra. The structure maintains electrical neutrality as a result of aluminium

substitution for silicon in the tetrahedral layer by the available alkalis such as Na+ [5].

Investigations have been most widely carried out using metakaolinite (dehydroxylated 1:1

clay mineral) or fly ash (dust from coal incineration) as an aluminosilicate-based material to prepare

geopolymers [6-13]. This study focused on the alkaline activation of calcined clay produced by the

thermal treatment of raw clay (naturally occurring aluminosilicate-material). The heating of clay

minerals at the temperature ranging from 600 to 850°C improved their reactivity in alkaline media.

This reactivity is associated with the removal of the structural water from the crystalline clay layers

producing amorphous or semi-amorphous products of high surface area and high chemical

reactivity [14-18].

From the standpoint of strength, durability and other properties of the final products, the

most effective alkaline activator in geopolymer synthesis is the mixture composed of NaOH and

sodium silicate hydrate (waterglass) [19-24]. The waterglass solution is needed for the

geopolymerization process and acted as binder, alkali activator and dispersant or plasticizer [25].

Moreover, it will provide additional soluble silicate (SiO2 content) and alkali (Na2O content)

necessary for the geopolymerization reaction to proceed [9]. The SiO2 content is important to

provide the silicate species to allow for the rapid exchange and oligomerization reaction between

the aluminate and silicate species from the aluminosilicate-based material, and the silicate species

3
from the waterglass [25]. On the other hand, the Na2O content is also very important for improving

the dissolution of the aluminosilicate-based material which affects greatly the rate and extent of the

polymerization and the stabilization of the final products [25,26]. Moreover, the waterglass may

generally accelerate the geopolymerization process and enhance the mechanical strength as well as

the physical properties of the end products [19-23]. The latter advantages are generally lowered by

the use of NaOH as the sole activator [20,27]. However, the industrial production of waterglass

used elevated temperatures of around 1400°C to melt silica sand and soda ash leading to energy

consumption and considerable emission of greenhouse gasses [2,28]. Hence, the use of NaOH

solution instead of a waterglass solution would be of a great benefit in reducing the cost

and the environmental impact of the construction material manufactured. Additionally, some

alternative materials such as rice husk ash or waste glass can also be used in the place of

waterglass activator in geopolymer synthesis [29-32].

Referring to the aforementioned studies dealing with the synthesis of geopolymers, to the

best of our knowledge, very little attention has been paid to the use of NaOH as the only activator in

geopolymerization process [33] and very few studies have been devoted to the polymerization of

calcined raw clays [14,15,24].

The aim of this work is to study the effect of NaOH concentration excluding the use of

waterglass on the microstructure transformations and the properties (flexural strength, density,

water absorption, open porosity and electrical conductivity) of alkali-activated calcined raw clay.

2. Materials & experimental procedures

The raw clay was obtained from the city of Safi (Morocco). Table 1 shows its chemical and

mineralogical composition. The red clay was calcined at 700°C for 2H in a furnace, operating at

ordinary atmosphere. This calcination destabilizes the structure of kaolinite and chlorite

4
(dehydroxylation) without alteration of mica/illite. The amount of the dolomite radically reduced.

Fig. 1 shows the particle size distribution of the calcined clay.

Dried samples of the calcined clay were mixed with different alkali solutions of sodium

hydroxide (4-14 M) and then worked for 10 min to give a consistent paste. The molar ratio:

H2O/SiO2 was maintained constant (2.7), whereas that of Na2O/SiO2 varied in the range (0.12-0.36).

Prismatic samples 10 х 20 х 50 mm ³ were shaped from the fresh paste by manual pressing using a

nylon (PVC) mold. The as-prepared bars were cured at 60°C for 15 hours by placing them in a

ventilated oven. After curing period, samples were removed and allowed to dry at room temperature

for 30 days. The temperature and curing time were chosen based on the author’s previous findings

[34-36].

The cured samples were characterized by X-ray diffraction analysis (XRD), Fourier

transform infrared spectroscopy (FT-IR), Scanning electron microscopy (SEM) and Thermal

analysis (DSC-TG). XRD analysis was carried out using a Philips X'Pert MPD diffractometer

operating with a copper anode (λ (Kα) = 1.5418 Å). The FT-IR spectra were recorded by a Perkin

Elmer spectrophotometer functioning in the range 4000-400 cm-1. For this purpose, 10 mg

of powdered sample was mixed with 90 mg of KBr. The deconvolution of the FT-IR spectra

was realized using the pickfit software with the regression coefficient (r2) value of 0.999 [37].

The fitting process was performed in accordance with procedures described in the literature [38-41]

combined with the self-fitting function of the software. SEM examination was conducted on gold

coated freshly fractured pieces of cured samples by a Zeiss Supra 40 VP scanning electron

microscope equipped with the X-Max 20 mm2 Silicon Drift Detector. Thermal analysis was realized

with a Setaram Setsys 24 apparatus functioning in air at the heating rate of 10°C/min.

The flexural strength (FS) of cured samples was measured using an Instron 3369 apparatus.

The used load and loading speed were 50 kN and 0.1 mm/min, respectively. Density (d) of cured

5
samples was determined by pycnometry using edible oil (d = 0.9166) as a solvent, while the open

porosity (OP) was deduced from the relation: OP = [(d– da)/d] x 100 x Va, where da and Va are the

apparent density and the apparent volume of samples respectively. For water absorption (WA)

measurements, the cured samples were placed on a water immersed sponge and periodically

weighted. The amount of absorbed water was deduced from the saturation plateau of plotted kinetic

curves. Measurements of electrical conductivity (EC) were conducted on suspensions composed of

1g of powdered samples and 10 mL of distilled water.

3. Results & discussion

3.1. Microstructural characterization

The XRD analysis of cured samples (Fig. 2) shows the presence of new zeolite-like

crystalline phase, zeolite ZK-14 and sodium carbonate. The occurrence of the zeolitic phase was

supported by the set of the absorption bands at 733, 665 and 440 cm-1 (Fig. 3) [21]. The new band at

733 cm-1 was assigned to four membered ring structures [42]. However, this band was also assigned

to symmetrical vibration of Si-O-T bonding of AlO4 and SiO4 tetrahedrons in geopolymer structures

[21]. The new band at 665 cm-1 was assigned to single and double six membered ring structures

[42]. Referring to the X-ray diffractograms of cured samples (Fig. 2), the crystalline phases of illite

and quartz detected in the initial material remained apparently unchanged. The FT-IR spectra of

Fig. 3 exhibit the prominent bands of quartz (797, 777, 696), illite (914, 3622 cm -1), and K-feldspar

(638 cm-1). The intensity of the latter bands (illite and K-feldspar) drastically diminished in the

samples with higher NaOH concentrations. The small absorption bands at 851-883 cm-1 and the

large bands at 1411-1523 cm-1 were associated to CO32- ions in the dolomite structure (Fig. 3). The

higher intensity of the latter bands in the cured samples compared to the starting material might

indicate the formation of sodium carbonate [43-45]. On the other hand, the coexistence of the band

6
at 881 cm-1 with the weak band at 851 cm-1 were taken as an indication of an increase in the amount

of Al3+ substitution in the silicate network of the geopolymer replacing the Si4+ [42]. The SEM

examination of Fig. 4 shows bundles of fibers which represent the partially developed zeolite ZK-

14 crystallites phase. The distinct particles emphasized in the circles in Fig. 5 are identified as being

sodium carbonate crystallites.

As far as the IR spectra are concerned (Fig. 3), the band at 534 cm-1 which might be

attributed to the Si-O-AlVI bonds in kaolinite (where the Al is in octahedral coordination), shifted

to higher wavenumber (557 cm-1) after calcination, then shifted slightly (559 cm-1) in the most of

geopolymer samples, suggesting that a change in the chemical bonding of the system has taken

place [21,46]. The latter band with the one at 1385 cm-1 might be associated to the metakaolinite

which gives evidence to the presence of little unreacted metakaolinite particles [21,46]. Moreover,

the band at 470 cm-1 might be linked to in-plane bending of Si–O and Al–O linkages originating

from within individual tetrahedral. After calcination, it shifted to higher wavenumber (482 cm-1),

then shifted slightly to lower wavenumbers (480 cm-1) in geopolymer samples which could be

pointed to the formation of a zeolitic phase [46]. The broad absorption band centered at 3438 cm-1

with the one at 1639 cm-1 are presumably associated to stretching and bending vibration of water,

respectively. The existence of several bonds at 3200-3600 cm-1 and around 1600-1640 cm-1

combined with a very weak bond at 3780 cm-1 suggest the presence of isolated non-interacting

surface silanol and aluminol groups (Si-OH/Al-OH) in some geopolymer samples. The increased

intensity of these bands was supposed to be related to the formation of activated products with more

surface hydroxyl groups which were hydrogen-bonded to adsorbed water (Si-OH…..H2O and

Al-OH….H2O) [39,47-51]. The presence of these hydroxyl groups was strongly dependent on

the dissolution rate of the calcined clay as an aluminosilicate source material. Furthermore, the

dissolution rate was found to be related to some factors such as the size and the specific surface area

of particles as well as the solution pH.

7
In the midwavenumber region, the position of the main band at ~1029 cm−1 in calcined clay

attributed to the Si-O-T (T= Si, Al) asymmetric stretching vibration (containing multiple

overlapping components which sum to give a broad peak) shifted to lower wavenumbers in all

geopolymer samples indicating that more drastic changes take place, affecting the Al–O and Si–O

bonds inside individual tetrahedral (Fig.3). This band shift was considered to be related to the

formation of geopolymers, eventhough it was observed also in the formation of zeolites [42].

According to references [42,52,53], this shift in the main band means that some changes were

occurring in the length and angle of the Si-O-T (T= Si, Al) bonds in the silicate network. These

changes caused either by an increase in the concentration of non-bridging oxygens (NBOs) of the

form [Si-OH, (Si-O-, Na+), or Al-OH], or by more substitution of Al into the silicate network.

Changes might also be caused by increasing compressive stress, fictive T, hydrostatic pressure,

porosity volume fraction (increases in pore size). However, the effect of a changing Si/Al ratio was

not considered. It is difficult to determine to what extent the geopolymerization reaction contributes

to this shifting of the main band, eventhough it was found that reaction with an alkali (Ca2+,...)

could lead to this peak shifting to lower wavenumbers [46]. It was reported that this shift could be

due to an obvious change in the microstructure during geopolymerization reaction leading to the

formation of new products with different microstructure from calcined clay. Similar result is found

in the XRD pattern as the newly product zeolite ZK-14 formed. This shift might be attributed also

to the partial replacement of SiO4 tetrahedron by AlO4 tetrahedron, resulting in a systematically

change in the local environment of Si-O band, leading to the formation of an aluminosilicate

network [54]. Previous study [55] had shown that the minerals (such as quartz, micas, feldspath…)

and the reactivity of the metakaolinite induce the formation of different networks within the

geopolymer material. These minerals might lead to an increase in shift for some geopolymer

samples because they may form chemical bonds with the Si-O and aluminosilicate species [56].

8
To clarify the causes behind the band shift, the spectra (Fig. 3) was deconvoluted in the

range 1200-900 cm-1. This deconvolution revealed the overlap of several fundamental bands

(Fig. 6). The small band at 1041-1046 cm−1 is associated to the (Si-O) stretching of calcined clay

[57], whereas the bands at 1194-1270 cm-1 and 1131-1149 cm-1 are linked to the symmetric

stretching mode of calcined clay [38,39,41]. The persistence of these bands means that a part of

clay still unreacted after geopolymerization. The wide new band at 1063-1082 cm-1 is attributed to

the asymmetric stretching vibrations of Si–O–Si [58]. The most notable changes after alkali

activation of the calcined clay were the new bands at 993-1003 cm-1 and 968-979 cm-1 (Fig. 6).

Indeed, the intensity of the band at ~1013 cm−1 in calcined clay drops to give rise to these new

bands which were connected with a shift of the overlapping peak towards lower frequencies. The

first new band at 993-1003 cm-1 is assigned to Si-O-T (T= tetrahedral Si, Al) asymmetric stretching

vibration in the geopolymer networks [39]. This band was found to be sensitive to connectivity of

the gel network and Si/Al ratio. In this study, the band was observed at a wavenumber associated

with the presence of predominantly Si-O-Al bonds [39,59,60]. The second new band at 968-979

cm-1 was assigned to Si-O asymmetric stretching of non-bridging oxygen (NBO) sites of the form

Si-O-Na+ type structures [39,40-42].

It can be seen from Fig. 7 that with increasing NaOH concentration, the two observed new

bands shifted to lower wavenumbers. From previous research [40], it was found that an increase in

the number of NBO should move the main band (containing the overlapping two new bands)

attributed to the Si-O-T (T= Si, Al) asymmetric stretching vibration to lower wavenumbers as well

as an increasing alkali metal oxide inclusion and/or a greater extent of silicate depolymerisation.

The similar trend was commonly observed for silicate glasses. Furthermore, it was stated that an

aluminosilicate with Al/Na = 1 (the amount of Na equals the amount of the negatively charged and

tetrahedrally coordinated Al) contains no NBO and that NBO was generated if Al/Na was less than

1. Consequently, the increased extent of alkali-activation should increase the NBO concentration of

9
the aluminosilicate (the Al is supposed to be in fourth coordination). This suggests that the observed

shift of the two new bands with the increase of NaOH concentration is likely to be related to the

occurrence of a great number of NBO due to the increased degree of fixation of Na by the

aluminosilicate gel framework as well as the Al/Na ratio which was most likely to be less than 1.

Referring to Fig. 8, the relative intensities relevant to the zeolitic phase were higher as the alkali

content increases from 4 to 11.5 M indicating that the observed shift of the two new bands is

probably caused by a specific crystallization event in the gel systems. As a matter of fact,

this might imply that the zeolitic phase preferred to form in samples with highly alkali content. This

may be a solubility effect, or it may be because high NaOH concentration favors the process

of crystallization of more polymer gels, which contributes positively to the enhancement of

mechanical strength (Fig. 11).

Considering the DSC/TG curves of Fig. 9, the endothermic peaks observed in the range

27-200°C indicates desorption of water which can be due to the elimination of free pore water in

geopolymer with almost 5.35% of mass loss [61]. This kind of dehydration might be linked to the

evacuation of residual water from the large pores in the network structure of geopolymer [62]. The

endothermic peaks observed in the range of 200-600°C indicates removal of water tightly bound to

the geopolymer structure with a mass loss of approximately 3.69% [63]. This dehydration might be

related to water loss from small inaccessible pores since the driving force for permeation is reached

at a temperature higher than that for larger more accessible pores. Furthermore, it was found that the

release of water at higher temperature could also be due to higher energy required to liberate water

chemically bound to the framework [62]. Two exothermic peaks are observed in the range 760-

980°C. The XRD diffractograms of the heated geopolymers at temperatures around 775 and 978°C

show the recrystallization of analogous systems to form nepheline (Fig. 10). This nepheline appears

at around 775°C and still clearly observable by XRD at about 978°C. This neoformation was also

revealed in view of the color change from reddish due to hematite to yellowish.

10
In the light of the above results, the alkali activation of calcined clay seems to influence the

nature of the reaction products formed and has a crucial role on them. It is likely that increasing the

NaOH concentration favors the transformation of more polymer gels toward the crystalline zeolite

ZK-14. In this connection, the dissolution of Al-Si metakaolinite (deriving from heated kaolinite)

and amorphous products of chlorite were considered to contribute mainly to the formation of

crystalline zeolite ZK-14 within the geopolymer matrix. The atmospheric carbonation of excess

NaOH migrating from the interior to the surface of the geopolymer samples led apparently to the

formation of sodium carbonate.

3.2. Mechanical & physical properties

The NaOH concentration was found to have a strong effect on the flexural strength and

its increase from 4 to 11.5 M enhances this property (Fig. 11). This increase improves also the

densification of the geopolymer matrix and reduces the water absorption (Fig. 11 and 12). The

positive effect of NaOH on these properties might be associated with the intensive development

of the zeolitic-fibrous microstructure (Fig. 4). The latter could reduce the open porosity and thus the

water absorption by gradually filling the voids and reducing the pore size (Fig. 12). As can be

deduced from the kinetic curves of Fig. 13, water absorption was a fast process since saturation of

samples was reached in less than 5 min. The rapid absorption of water might be linked to samples’

porosity. The kinetics data fitted well the equation: Ln (1-Qt/Qe) = - k t (Qt and Qe are the amounts

of absorbed water at t instant and equilibrium, respectively. k is the kinetic constant). The values of

k varied in the range (18.38–51.01) x 10-2 min-1. The variation of k is related to the microstructure

modification since T was constant. The highest concentration of 14 M reduces the density due to

samples’ expansion (Fig. 11). It causes also the decline of strength which may be attributed to the

excess of Na+ ions leading to the formation of sodium carbonate. This carbonate will probably

11
affect the geopolymerization process and result in decreased strength [21]. In contrast, the

concentrations from 4 to 11.5 M might have provided optimum alkalinity for dissolution of the

aluminosilicate source (clay), where sufficient Al3+ and Si4+ ions are released from the

aluminosilicate and participated in the geopolymerization process. Even though, the very higher

NaOH concentration of 14 M has higher dissolution ability. This was not desirable by the

polycondensation process as excess Na+ ions left in the system weakens the structure. Moreover,

it may cause the gel to set rapidly before it can be transformed to a more homogeneous structure

[21]. Therefore, the increase of NaOH concentration from 4 to 11.5 M seemed to contribute

markedly on the enhancement of the above properties and thus the extent of geopolymerization

reaction as commonly reported by Yao et al. [64] and Wang et al. [65].

The electrical conductivity increased as the NaOH concentration increased (Fig. 12) which

seems to be consistent with the increase in the amount of crystallized zeolite ZK-14 (Fig. 8). This

type of zeolite was known to undergo ion exchange of Na more readily [66]. It’s highly probable

that there was a strong relation between the crystallization of more polymer gels toward this type of

zeolite and the observed increase of EC (see the supplementary data). As a matter of fact,

conductivity reflects the degree of fixation and the magnitude of the cation exchange of Na+ by

H3O+ that takes place. The observed increase of conductivity for higher values of concentration

suggested that samples may contain excess NaOH which washes out resulting in a higher

conductivity. In contrast, for lower values of concentration, the samples may contain enough of

NaOH for the reaction to proceed sufficiently which might lead the Na-component to react nearly

completely [24]. In this case, sodium would be more strongly bound to the aluminosilicate gel

framework and its diffusion to the solution was prohibited. The increased conductivity of specimens

might also be related to the formation of sodium carbonate.

12
4. Conclusion

The alkali activation of calcined clay by sodium hydroxide as the only activator led to the formation

of polymeric product type zeolite ZK-14 and sodium carbonate. It appears that the presence of the

former product involved mainly metakaolinite and amorphous products of chlorite. The excessive

formation of sodium carbonate affects negatively the consolidation of the matrix. However, the

formation of the zeolitic phase within the geopolymer matrix seems to have a positive effect on

strength development and contributes to the increase of electrical conductivity. The geopolymers

can be transformed into stable anhydrous alkaline aluminosilicate (nepheline) by heating to high

temperatures. The increase of NaOH concentration from 4 to 11.5 M caused the sample

strengthening, densification, reduced open porosity and water absorption, likely because of its

positive influence on the extent of geopolymerization. The kinetic of water absorption was fast

and the data followed the pseudo-first-order equation. Moreover, the increase of NaOH

concentration enhances the electrical conductivity since it was associated to the increase in the

amount of Na+ ions. The improvement of the mechanical strength under the effect of alkali

concentration tended to reach a limit (about 5 MPa).

13
References

[1] B.C. McLellan, R.P. Williams, J. Lay, A. van Riessen, G.D. Corder, Costs and carbon emissions

for geopolymer pastes in comparison to ordinary portland cement, J. Clean. Prod. 19 (2011)

1080-1090.

[2] L.K. Turner, F.G. Collins, Carbon dioxide equivalent (CO2-e) emissions: A comparison between

geopolymer and OPC cement concrete, Constr. Build. Mater. 43 (2013) 125-130.

[3] P. Duxson, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, The role of inorganic polymer

technology in the development of ‘green concrete’, Cement Concrete Res. 37 (2007) 1590-1597.

[4] J.S.J. van Deventer, J.L. Provis, P. Duxson, Technical and commercial progress in the

adoption of geopolymer cement, Miner. Eng. 29 (2012) 89-104.

[5] J. Davidovits, Geopolymers and geopolymeric materials, J. Therm. Anal. 35 (1989) 429-441.

[6] V.F.F. Barbosa, K.J.D. MacKenzie, C. Thaumaturgo, Synthesis and characterisation of materials

based on inorganic polymers of alumina and silica: sodium polysialate polymers, Int. J. Inorg.

Mater. 2 (2000) 309-317.

[7] V.F.F. Barbosa, K.J.D. MacKenzie, Synthesis and thermal behaviour of potassium sialate

geopolymers, Mater. Lett. 57 (2003) 1477-1482.

[8] S.J. Lyu, T.T. Wang, T.W. Cheng, T.H. Ueng, Main factors affecting mechanical characteristics

of geopolymer revealed by experimental design and associated statistical analysis, Constr.

Build. Mater. 43 (2013) 589-597.

[9] B. Zhang, K.J.D. MacKenzie, I.W.M. Brown, Crystalline phase formation in metakaolinite

geopolymers activated with NaOH and sodium silicate, J. Mater. Sci. 44 (2009) 4668-4676.

14
[10] D. Panias, I.P. Giannopoulou, Development of inorganic polymeric materials based on fired

coal fly ash, Acta Metallurgica Slovaca 12 (2006) 321-327.

[11] M.N. Muzek, J. Zeliæ, D. Joziæ, Microstructural characteristics of geopolymers based on

alkali-activated fly ash, Chem. Biochem. Eng. 26 (2012) 89-95.

[12] K. Somna, C. Jaturapitakkul, P. Kajitvichyanukul, P. Chindaprasirt, NaOH-activated ground

fly ash geopolymer cured at ambient temperature, Fuel 90 (2011) 2118-2124.

[13] F. Škvára, L. Kopecký, J. Nìmeèek, Z. Bittnar, Microstructure of geopolymer materials based

on fly ash, Ceram-Silikaty 50 (2006) 208-215.

[14] A. Buchwald, M. Hohmann, K. Posern, E. Brendler, The suitability of thermally activated

illite/smectite clay as raw material for geopolymer binders, Appl. Clay Sci. 46 (2009) 300-304.

[15] N. Essaidi, B. Samet, S. Baklouti, S. Rossignol, Effect of calcination temperature of tunisian

clays, Ceram-Silikaty 57 (2013) 251-257.

[16] B.B. Sabir, S. Wild, J. Bai, Metakaolin and calcined clays as pozzolans for concrete: a review,

Cement Concrete Comp. 23 (2001) 441-454.

[17] A. Elimbi, H.K. Tchakoute, D. Njopwouo, Effects of calcination temperature of kaolinite

clays on the properties of geopolymer cements, Constr. Build. Mater. 25 (2011) 2805-2812.

[18] A.M. Rashad, Metakaolin as cementitious material: history, scours, production and

composition – a comprehensive overview, Constr. Build. Mater. 41 (2013) 303-318.

[19] F. Pacheco-Torgal, J. Castro-Gomes, S. Jalali, Alkali-activated binders: A review. Part 2.

About materials and binders manufacture, Constr. Build. Mater. 22 (2008) 1315–1322.

15
[20] M.L. Granizo, M.T.B. Varela, S. Martínez-Ramírez, S.A. Sika, Alkali activation of

metakaolins: parameters affecting mechanical, structural and microstructural properties, J. Mater.

Sci. 42 (2007) 2934-2943.

[21] Y.M. Liew, H. Kamarudin, A.M.M.A. Bakri, M. Luqman, I.K. Nizar, C.M. Ruzaidi,

C.Y. Heah, Processing and characterization of calcined kaolin cement powder, Constr. Build.

Mater. 30 (2012) 794-802.

[22] J.T. Kim, D.S. Seo, G.J. Kim, J.K. Lee, Effect of water glass on compressive strength of

aluminosilicate-based geopolymer, Adv. Mat. Res. 97-101 (2010) 2273-2276.

[23] A.V. Kirschner, H. Harmuth, Investigation of geopolymer binders with respect to their

application for building materials, Ceram-Silikaty 48 (2004) 117-120.

[24] M.B. Diop, M.W. Grutzeck, L. Molez, Comparing the performances of bricks made with

natural clay and clay activated by calcination and addition of sodium silicate, Appl. Clay. Sci.

54 (2011) 172-178.

[25] C.Y. Heah, H. Kamarudin, A.M.M.A. Bakri, M. Binhussain, M. Luqman, I.K. Nizar, C.M.

Ruzaidi, Y.M. Liew, Study on solids-to-liquid and alkaline activator ratios on kaolin-based

geopolymers, Constr. Build. Mater. 35 (2012) 912-922.

[26] Z. Yunsheng, S. Wei, L. Zongjin, Composition design and microstructural characterization of

calcined kaolin-based geopolymer cement, Appl. Clay Sci. 47 (2010) 271-275.

[27] H. Kamarudin, A.M.M.A. Bakri, M. Binhussain, C.M. Ruzaidi, M. Luqman, C.Y. Heah, Y.M.

Liew, Preliminary study on effect of NaOH concentration on early age compressive strength of

kaolin based green cement, 10. International Conference on Chemistry and Chemical Process,

IPCBEE, 2011, pp. 18-24.

16
[28] G. Habert, J.B. d’Espinose de Lacaillerie, N. Roussel, An environmental evaluation of

geopolymer based concrete production: reviewing current research trends, J. Clean. Prod. 19

(2011) 1229-1238.

[29] M. Torres-Carrasco, F. Puertas, Waste glass in the geopolymer preparation. Mechanical and

microstructural characterisation, J. Clean. Prod. 90 (2015) 397-408.

[30] H.K. Tchakouté, C.H. Rüscher, S. Kong, N. Ranjbar, Synthesis of sodium waterglass from

white rice husk ash as an activator to produce metakaolin-based geopolymer cements, Journal of

Building Engineering 6 (2016) 252–261. http://dx.doi.org/10.1016/j.jobe.2016.04.007.

[31] E. Kamseu, L.M. Beleuk à Moungam, M. Cannio, N. Billong, D. Chaysuwan, U.C. Melo, C.

Leonelli, Substitution of Sodium Silicate with Rice Husk Ash-NaOH solution in metakaolin

based geopolymer cement concerning reduction in global warming, J. Clean. Prod. 142 (2017)

3050–3060.

[32] H.K. Tchakouté, C.H. Rüscher, S. Kong, E. Kamseu, C. Leonelli, Thermal behavior of

metakaolin-based geopolymer cements using sodium waterglass from rice husk ash and waste

glass as alternative activators, Waste Biomass Valori. 8 (2017) 573–584.

[33] R.H. Abdul Rahim, K.A. Azizli, Z. Man, T. Rahmiati, M.F. Nuruddin, Effect of sodium

hydroxide concentration on the mechanical property of non sodium silicate fly ash based

geopolymer, J. Appl. Sci. 14 (2014) 3381-3384.

[34] M.S. Muñiz-Villarreal, A. Manzano-Ramírez, S. Sampieri-Bulbarela, J.R. Gasca-Tirado,

J.L. Reyes-Araiza, J.C. Rubio-Ávalos, J.J. Pérez-Bueno, L.M. Apatiga, A. Zaldivar-Cadena,

V. Amigó-Borrás, The effect of temperature on the geopolymerization process of a metakaolin-

based geopolymer, Mater. Lett. 65 (2011) 995-998.

17
[35] E. Arioz, Ö. Arioz, Ö.M. Koçkar, The effect of curing conditions on the properties of

geopolymer samples, Int. J. Chem. Eng. Appl. 4 (2013) 423-426.

[36] M. Bing-hui, H. Zhu, C. Xue-min, H. Yan, G. Si-yu, Effect of curing temperature on

geopolymerization of metakaolin-based geopolymers, Appl. Clay. Sci. 99 (2014) 144-148.

[37] M. Wajdyr, Fityk: a general-purpose peak fitting program, J. Appl. Crystallogr. 43 (2010)

1126-1128.

[38] P. Rovnanik, Effect of curing temperature on the development of hard structure of metakaolin

based geopolymer, Constr. Build. Mater. 24 (2010) 1176-1183.

[39] Z. Zhang, H. Wang, J.L. Provis, F. Bullen, A. Reid, Y. Zhu, Quantitative kinetic and structural

analysis of geopolymers. Part 1. The activation of metakaolin with sodium hydroxide,

Thermochim. Acta 539 (2012) 23-33.

[40] W.K.W. Lee, J.S.J. van Deventer, Use of infrared spectroscopy to study geopolymerization of

heterogeneous amorphous aluminosilicates, Langmuir 19 (2003) 8726–8734.

[41] Z. Zhang, J.L. Provis, H. Wang, F. Bullen, A. Reid, Quantitative kinetic and structural analysis

of geopolymers. Part 2. Thermodynamics of sodium silicate activation of metakaolin,

Thermochim. Acta 565 (2013) 163-171.

[42] C.A. Rees, Mechanisms and kinetics of gel formation in geopolymers (a PhD thesis), Faculty

of Engineering, Chemical and Biomolecular Engineering, University of Melbourne, 2007.

http://hdl.handle.net/11343/39579.

[43] I. Lecomte, C. Henrist, M. Liegeois, F. Maseri, A. Rulmont, R. Cloots, (Micro)-structural

comparison between geopolymers, alkali-activated slag cement and Portland cement, J. Eur.

Ceram. Soc. 26 (2006) 3789-3797. http://dx.doi.org/10.1016/j.jeurceramsoc.2005.12.021.


18
[44] S.A. Bernal, J.L. Provis, V. Rose, R. Mejía de Gutierrez, Evolution of binder structure in

sodium silicate-activated slag-metakaolin blends, Cement Concrete Comp. 33 (2011) 46–54.

[45] A.D. Hounsi, G. Lecomte-Nana, G. Djétéli, P. Blanchart, D. Alowanou, P. Kpelou, K. Napo,

G. Tchangbédji, M. Praisler, How does Na,K alkali metal concentration change the early age

structural characteristic of kaolin-based geopolymers (Review), Ceram. Int. 40 (2014) 8953-

8962.

[46] J.G.S. van Jaarsveld, J.S.J. van Deventer, G.C. Lukey, The effect of composition and

temperature on the properties of fly ash- and kaolinite-based geopolymers, Chem. Eng. J. 89

(2002) 63-73.

[47] G. Öhlmann, H. Pfeifer, R. Fricke, Catalysis and adsorption by zeolites, Stud. Surf. Sci. Catal.

65, Elsevier, Amsterdam, 1991, pp. 428 (ISBN: 9780080887500).

[48] B. Viswanathan, S. Sivasanker, A.V. Ramaswamy. Catalysis: principles and applications,

Narosa, New Delhi, 2002, pp. 135 (ISBN: 9780849324246).

[49] A. Gedeon, P. Massiani, F. Babonneau. Zeolites and related Materials: trends targets and

challenges (set): 4th International FEZA Conference, 2-6 September 2008, Paris, France, Stud.

Surf. Sci. Catal. 174 - Part A, Elsevier, Amsterdam, 2008, pp. 814 (ISBN: 9780080569864).

[50] H.G. Karge, J. Weitkamp, Characterization I: molecular sieves-science and technology,

Spinger-Verlag, Berlin (Germany), 2004, pp. 80 (ISBN: 9783540643357).

[51] E. Van Steen, L.H. Callanan, C. Claeys, Recent advances in the science and technology of

zeolites and related materials: proceedings of the 14th International Zeolite Conference, Cape

Town, South Africa, 25-30th April 2004, Stud. Surf. Sci. Catal. 154 - Part A, Elsevier,

Amsterdam (The Netherlands), 2004, pp. 153 (ISBN: 9780444518279).

19
[52] C.A. Rees, J.L. Provis, G.C. Lukey, J.S.J. van Deventer, The mechanism of geopolymer gel

formation investigated through seeded nucleation, Colloids Surf. A. Physicochem. Eng. Asp. 318

(2008) 97-105.

[53] A. Hajimohammadi, J.L. Provis, J.S.J. van Deventer, Effect of alumina release rate on the

mechanism of geopolymer gel formation, Chem. Mater. 22 (2010) 5199-5208.

[54] J. Davidovits, Geopolymer chemistry and applications: 2nd edition, Geopolymer Institute,

Saint-Quentin (France), 2008, pp. 68 (ISBN: 9782951482012).

[55] A. Autef, E. Joussein, A. Poulesquen, G. Gasgnier, S. Pronier, I. Sobrados, J. Sanz, S.

Rossignol, Influence of metakaolin purities on potassium geopolymer formulation: the existence

of several networks, J. Colloid. Interf. Sci. 408 (2013) 43-53.

[56] A. Autef, E. Joussein, G. Gasgnier, S. Pronier, I. Sobrados, J. Sanz, S. Rossignol, Role of

metakaolin dehydroxylation in geopolymer synthesis, Powder Technol. 250 (2013) 33-39.

[57] J. Madejova, FTIR techniques in clay mineral studies (Review), Vib. Spectrosc. 31 (2003)

1-10.

[58] R. Redden, N. Neithalath, Microstructure, strength, and moisture stability of alkali activated

glass powder-based binders, Cement Concrete Comp. 45 (2014) 46-56.

[59] M. Gougazeh, Geopolymers from jordanian metakaolin: influence of chemical and

mineralogical compositions on the development of mechanical properties, Jordan J. Civ. Eng. 7

(2013) 236-257.

[60] Z. Yunsheng, S. Wei, L. Zongjin, Synthesis and microstructural characterization of fully-

reacted potassium-poly(sialate-siloxo) geopolymeric cement matrix, Aci. Mater. J. 105 (2008)

156-164.
20
[61] N. Saidi, B. Samet, S. Baklouti, Effect of composition on structure and mechanical properties

of metakaolin based PSS-Geopolymer, International Journal of Material Science (IJMSCI) 3

(2013) 145-151. DOI: 10.14355/ijmsci.2013.0304.03.

[62] J. Davidovits, Geopolymer, green chemistry and sustainable development solutions:

proceedings of the world congress geopolymer, Geopolymer Institute, Saint-Quentin (France),

2005, pp. 189-193 (ISBN: 9782951482005).

[63] A. Adriano, G. Soriano, J. Duque, Characterization of water absorption and desorption

properties of natural zeolites in Ecuador, Fifth International Symposium on Energy, Puerto Rico

Energy Center-Laccei, 2013, pp. 1-9.

[64] X. Yao, Z. Zhang, H. Zhu, Y. Chen, Geopolymerization process of alkali–metakaolinite

characterized by isothermal calorimetry, Thermochim. Acta 493 (2009) 49–54.

[65] H. Wang, H. Li, F. Yan, Synthesis and mechanical properties of metakaolinite-based

geopolymer, Colloids and Surfaces A: Physicochem. Eng. Aspects 268 (2005) 1–6.

[66] Ch. Baerlocher, L.B. McCusker, D.H. Olson, Atlas of Zeolite Framework Types: 6th, Elsevier,

2007, pp. 96-97 (ISBN: 978-0-444-53064-6).

21
Fig.1. Grain size distribution of the calcined clay.

22
Fig.2. X-ray diffraction patterns of calcined clay (CC) and NaOH activated samples. Z:

Zeolite

ZK-14 (PDF2# 84-0698), S: Sodium carbonate (PDF2# 86-0307), I: Illite (PDF# 43-0685); Q:

Quartz (PDF# 05-0490).

23
Fig.3. FT-IR spectra of raw (RC), calcined clay (CC) and NaOH activated samples. I: Illite;

Q: Quartz; F: K-feldspar.

24
Fig.4. SEM micrograph and EDS spectrum relevant to the zeolitic-fibrous phase (the big circle

marks the formed fibers) developed in a sample activated by 11.5 M of NaOH solution.

25
Fig.5. SEM micrograph and EDS spectrum of sodium carbonate (S) detected in a sample activated

by 11.5 M of NaOH solution.

26
Fig.6. FTIR spectral deconvolution of the region 900-1200 cm-1 for calcined clay (a) and

NaOH activated samples. (b) 6.5 M, (c) 9 M, and (d) 11.5 M.


27
Fig.7. Effect of NaOH concentration on the position of the resolved new bands in the FTIR

deconvolution of the region 900-1200 cm-1.

28
Fig.8. Evolution of the XRD major reflection peak intensity (110) and the 665 cm-1 IR band

of Zeolite ZK-14 versus NaOH concentration.

29
Fig.9. DSC-TG curves of selected samples activated with NaOH solutions. (a) 9 M and (b)

11.5 M.

30
Fig.10. X-ray diffraction patterns of selected NaOH activated samples heated at (1): 775°C

and

(2) 978°C. (a) 9 M and (b) 11.5 M. N: Nepheline (PDF# 76-2466); Q: Quartz (PDF# 83-2471); H:

Hematite (PDF# 85-0987).

31
Fig.11. Evolution of flexural strength (FS) and density (d) of cured samples as a function of NaOH

concentration.

32
Fig.12. Variation of water absorption (WA), open porosity (OP) and electrical conductivity (EC) of

cured samples versus NaOH concentration.

33
Fig.13. Kinetics curves relevant to the water absorption of NaOH activated samples.

34
Table 1. Chemical and mineralogical compositions (wt.%) of the raw clay.

Chemical composition
Si A F C M K2 N TiO *
O2 l2O3 e2O3 aO gO O a2O 2 L.O.I
5
2 3 3. 3. 1. 9.
2 0.7
3 2.4 .7 5 5 2 3

Mineralogical composition
Illi Kaolin Quar Chlo Dol Hem K-
te ite tz rite omite atite feldspar
33 19 10 6 4 3
25
* Loss on ignition.

Research highlights

 Zeolite ZK-14 phase formed within the geopolymer matrix.

 Sodium carbonate was detected as a carbonation product.

 Heating of geopolymers resulted in the recrystallization of Nepheline.

 Properties’ change was discussed in relation to microstructure characterization.

35

You might also like