Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Minerals, Metallurgy and Materials

Volume 26, Number 5, May 2019, Page 555


https://doi.org/10.1007/s12613-019-1764-2

Replacement of alkali silicate solution with silica fume in


metakaolin-based geopolymers

Ahmed Nmiri1,2), Myriam Duc3), Noureddine Hamdi1), Oumaya Yazoghli-Marzouk4),


and Ezzeddine Srasra1)
1) National Center for Research in Materials Sciences, Technopôle Borj Cedria (CNRSM), Soliman 8027, Tunisia
2) Department of Chemistry, Faculty of Sciences of Tunis, University of Tunis El-Manar, Tunis 2092, Tunisia
3) French Institute of Science and Technology for Transport, Development and Networks (IFSTTAR), Marne-la-Vallée 77447, France
4) Center for Studies and Expertise on Risks, Environment, Mobility and Development (CEREMA), Autun 71404, France
(Received: 31 May 2018; revised: 29 October 2018; accepted: 30 October 2018)

Abstract: A metakaolin (Mk)-based geopolymer cement from Tunisian Mk mixed with different amounts of silica fume (SiO2/Al2O3 molar
ratio varying between 3.61 and 4.09) and sodium hydroxide (10 M) and without any alkali silicate solution, is developed in this work. After
the samples were cured at room temperature under air for 28 d, they were analyzed by X-ray diffraction (XRD), Fourier transform infrared
(FTIR) spectroscopy, environmental scanning electron microscopy, mercury intrusion porosimetry, 27Al and 29Si nuclear magnetic resonance
(NMR) spectroscopy, and compression testing to establish the relationship between microstructure and compressive strength. The XRD,
FTIR, and 27Al and 29Si NMR analyses showed that the use of silica fume instead of alkali silicate solutions was feasible for manufacturing
geopolymer cement. The Mk-based geopolymer with a silica fume content of 6wt% (compared with those with 2% and 10%), corresponding
to an SiO2/Al2O3 molar ratio of 3.84, resulted in the highest compressive strength, which was explained on the basis of its high compactness
with the smallest porosity. Silica fume improved the compressive strength by filling interstitial voids of the microstructure because of its fine
particle size. In addition, an increase in the SiO2/Al2O3 molar ratio, which is controlled by the addition of silica fume, to 4.09 led to a geopo-
lymer with low compressive strength, accompanied by microstructures with high porosity. This high porosity, which is responsible for
weaknesses in the specimen, is related to the amount of unreacted silica fume.

Keywords: metakaolin; silica fume; geopolymer cement; compressive strength

1. Introduction the production of building materials [3]. However, the pro-


duction of geopolymer cement is more expensive than the
The ordinary cement manufacturing process has disastr- production of Portland cement because it requires large
ous effects on the atmosphere and contributes to global amounts of alkali silicates, which may not be available in
warming through the emission of huge amounts of green- large volumes, as mentioned by Davidovits [3]. Therefore,
house gases. Indeed, the conversion of limestone to hydrau- the development of geopolymer cements requires the identi-
lic calcium aluminate in the presence of clay requires high fication of new sources of alkaline silicates. In this project,
calcination temperatures (approximately 1450°C), contrary another source of silica—silica fume—was tested to pro-
to the maximum 800°C needed to fire the aluminosili- duce a sustainable geopolymer material with acceptable cost
cate-rich materials (such as kaolinitic clay) required to and no negative effects on the environment by maintaining
produce geopolymer cement. With good mechanical, the Na2O/SiO2 molar ratio below 0.625 (European Direc-
fire-retardant [1], and insulating properties [2], geopolymer tives 67/548/EEC and 91/155/EEC) [4].
cement also ensures huge reductions in CO2 emissions and Among the reactive aluminosilicate precursors mixed
is used for the storage of waste, as thermal insulation, and in with alkali hydroxide solutions, with or without silicate,
Corresponding author: Ahmed Nmiri E-mail: nmiriahmed@gmail.com
© University of Science and Technology Beijing and Springer-Verlag GmbH Germany, part of Springer Nature 2019
556 Int. J. Miner. Metall. Mater., Vol. 26, No. 5, May 2019

fly-ash, blast-furnace slags, and metakaolinite are the most strength of geopolymer mortars. However, a further increase
widespread [5–6]. According to Dupuy et al. [7], heated of the silica fume concentration caused a decrease in com-
materials such as metakaolinite are chosen because heating pressive strength because of the accompanying increase in
the raw material generally promotes the amorphization porosity. Latella et al. [20] also studied the mechanical
process to release silicium and aluminum when they contact properties of a Mk-based geopolymer manufactured with
alkaline solutions. Their degree of reactivity for further NaOH, silica fume, and Mk. They demonstrated that poros-
production of aluminosiliceous geopolymer materials is thus ity is the factor controlling the mechanical properties of the
increased. The temperature of kaolin calcination required to geopolymers and that compressive strength is more sensitive
produce metakaolin (Mk) varies from one author to another, to microstructure than to chemical composition.
ranging from 700°C for 12 h to 800 or 850°C for 2 h [8–10], Rostami and Behfarnia [21] examined the effect of using
which affects Mk reactivity. silica fume on the permeability of alkali-activated slag con-
According to some authors [11–13], the highest com- crete by substituting three levels of silica fume (5wt%,
pressive strength of Mk-based geopolymers is obtained 10wt%, and 15wt%) for slag. The results showed that in-
when the Na2SiO3/NaOH mass ratio is approximately 0.20 creasing the content of silica fume increases the compres-
and the NaOH solution concentration is approximately 10 M. sive strength and reduces the permeability of the alka-
Silva et al. [14] revealed that the properties of geopoly- li-activated slag concrete.
mer systems can be drastically affected by minor changes in Finally, the third component of the mixture, water, plays
the available SiO2 and Al2O3 concentrations during synthe- an important role in geopolymer synthesis. Water is essential
sis. The best resistance of Mk-based geopolymers was ob- to achieve good workability (fluidity) of the geopolymer
tained when the SiO2/Al2O3 molar ratio was between 3.0 cement paste. A high liquid/solid (L/S) ratio improves the
and 3.8. Moreover, the addition of silica fume, which has a workability of a fresh geopolymer mixture but generally re-
high percentage of SiO2, can improve the geopolymerization duces its compressive strength. Numerous reports [22–23]
process. Also, this granular stack creates a very dense mi- have revealed that a high L/S ratio and high NaOH concen-
crostructure, which results in a compact geopolymer con- tration can accelerate the dissolution of raw materials and
crete with good mechanical characteristics, as demonstrated the hydrolysis of Si and Al compounds; however, when the
by Okoye et al. [15], who studied the effect of silica fume OH− concentration is sufficiently high (e.g., 12 M), it can
introduced in fly-ash-based geopolymer concrete with also hinder polycondensation. In fact, larger amounts of
NaOH (14 M) and sodium silicate as alkali activators. Con- OH− attack the surface of the solid phase, leading to greater
cretes with silica fume have a reduced porosity [16] and a amounts of Si and Al species released into the liquid phase,
low permeability, which leads to substantial resistance to which in turn results in a higher compressive strength of the
aggressive attacks under acidic or sulfate-rich solutions, un- geopolymer concrete.
der carbonation, or in alkali silicate reactions [17]. In this research, we studied the effect of substituting
Although silica fume improves mechanical strength, its 2wt%, 6wt%, and 10wt% of Mk with silica fume in a mix-
quantity in a geopolymer must be optimized. To elucidate ture of Mk, silica fume, and sodium hydroxide (without any
the effects of silica fume on the properties of fly-ash-based alkali silicate solution, as usually used). After evaluating the
geopolymer, Duan et al. [18] partially replaced fly-ash with effect of silica fume on the mechanical properties of
silica fume at levels ranging from 0 to 30wt% in 10% in- Mk-based geopolymer cement, we introduce microstructural
crements. Their results revealed that increasing the silica and mineralogical considerations to better explain the ob-
fume content enhances the compressive strength, optimizes served behaviors.
the microstructure, and improves the thermal resistance of
the geopolymer. Uysal et al. [19] investigated the effect of 2. Experimental
adding silica fume on the physical properties of Mk-based
2.1. Materials
geopolymer composites. Their results revealed that the addi-
tion of silica fume yielded an improvement in mechanical Mk was obtained after natural kaolin from Tabarka (Tu-
strength and abrasion resistance. Dutta et al. [16] studied the nisia) was crushed to 106 μm and calcined at 700°C for 2 h.
effect of varying the quantity of silica fume (between 2.5% The chemical composition and particle size distribution
and 5%) incorporated into fly-ash geopolymer on its poros- curve of Mk are given in Table 1 and Fig. 1, respectively.
ity and compressive strength. The addition of silica fume in SiO2 and Al2O3 are the main components, with contents of
concentrations as high as 5% improved the compressive 63.76wt% and 30.93wt%, respectively. Small amounts of
A. Nmiri et al., Replacement of alkali silicate solution with silica fume in metakaolin-based geopolymers 557

iron and titanium are also present. The granulometric curves has a unimodal distribution, with a single homogeneous
(class distribution and cumulative distribution) of Mk, as population of particles smaller than 124 µm (the average
given by a particle size analyzer, are shown in Fig. 1. Mk particle diameter is approximately 15 µm).
Table 1. Chemical composition of metakaolin wt%
SiO2 Al2O3 Fe2O3 CaO MgO Na2O K2O TiO2 SO3 P2O5 LOI
63.76 30.93 1.43 0.11 0.36 0.1 0.88 1.61 0.1 0.12 0.58
Note: LOI⎯Loss on ignition.

Silica fume (Condensil S95DM for concrete applications)


was obtained from Sika France. Its chemical composition
(provided by the supplier) is given in Table 2.
Sodium hydroxide solution was obtained by dissolving
dry NaOH powder from Sigma-Aldrich (99% purity) in dis-
tilled water as well as sodium silicate (Na2SiO3, SiO2/Na2O
molar ratio of 2) solution supplied by Fisher Chemicals. Af-
terwards, NaOH solution was used directly or mixed with
the sodium silicate solution using a mechanical mixer at
ambient laboratory temperature and was protected from air
Fig. 1. Particle size distribution on metakaolin powder: class for 24 h before use to ensure that the activator component
distribution and cumulative distribution. was mixed uniformly.

Table 2. Silica fume characteristics from supplier


SiO2 / Apparent density / Real SiC / Free C / Total S / Na2O / BET / Oxide Cl /
wt% (g⋅cm–3) density wt% % wt% wt% (m2⋅g–1) content* / wt% wt%
95 (>92) 0.4–0.45 2.24 1.5 (<3) 1.5 (<3) 0.1 (<0.2) 0.6 (<1) 23 (20–26) 1 (<1.5) 0.06 (<0.2)
Note: *CaO, MgO, Al2O3, and Fe2O3.

2.2. Sample preparation NaOH solution (10 M, 23.6wt% Na2O, 76.4wt% H2O, and d
Mk was dry-mixed with silica fume until a homogeneous = 1.33 g/cm3, where d is the apparent density) used as an acti-
powder was obtained. The resulting blend was then me- vator. For comparison, Mk-based geopolymers with NaOH (10
chanically mixed with activators for 4 min. The fresh geo- M) (without silica fume addition) and with a mixture of NaOH
polymer paste was then rapidly poured into cylindrical solution and alkali silicate Na2SiO3 solution (d = 1.04 g/cm3)
molds with a 1:2 diameter-to-length ratio [24–25]. All of the were prepared with the ratios given in Table 3. In the mix-
samples were vibrated for 2 min on a vibration table to re- ture B-G, NaOH solution was mixed with Na2SiO3 with a
move trapped air bubbles. Three mixture formulations con- Na2SiO3/NaOH mass ratio of 0.2 to prepare a liquid alkali
taining different amounts of silica fume were prepared, with activator; the mixture was prepared 24 h prior to use.

Table 3. Composition of tested metakaolin-based geopolymer mixtures and the corresponding calculated molar ratios (Three sam-
ples were manufactured for each formulation)
Geopolymer Liquid/solid Metakaolin / Silica fume / SiO2/Al2O3 Na2O/SiO2 Na2O/Al2O3 H2O/Na2O
Activators
specimen mass ratio wt% wt% molar ratio molar ratio molar ratio molar ratio
A-G NaOH 0.97 100 0 3.39 0.36 1.22 11.18
B-G NaOH + Na2SiO3 1.11 100 0 3.54 0.35 1.25 13.05
SF1-G NaOH 1.10 98 2 3.61 0.38 1.38 11.36
SF2-G NaOH 1.10 94 6 3.84 0.37 1.44 11.35
SF3-G NaOH 1.10 90 10 4.09 0.37 1.51 11.34

The samples were left at ambient temperature for 24 h electron microscopy (Env. SEM) and mercury intrusion po-
before demolding. Compressive strength was tested after 28 d. rosimetry (MIP). Other fragments for X-ray diffraction
The fragments were then used for environmental scanning (XRD), Fourier transform infrared (FTIR) spectroscopy, and
558 Int. J. Miner. Metall. Mater., Vol. 26, No. 5, May 2019

thermal analysis were further finely crushed in an agate XRD with a Bruker D8 Advance diffractometer equipped
mortar to pass through an 80-µm sieve. with a copper anode (λCukα1 = 0.154 nm). The XRD patterns
were recorded in a 2θ span of 5° to 80° at a 2θ scanning rate
2.3. Testing methods
of 2°/min.
The compressive strength of all samples was measured at Open porosity measurements were carried out on
20°C after the samples were cured for 28 d at ambient la- freeze-dried samples using a mercury porosimeter (Auto-
boratory temperature. A compressive strength testing appa- pore IV, Micromeritics, France). The structural characteris-
ratus (EZ50 from AMETEK-LLOYD) was used, with the tics of the geopolymers (morphology and particle arrange-
displacement rate fixed at 5 mm/min for all of the tests [26]. ment) were studied on fresh fractured surfaces from geopo-
Compressive strength tests were performed using cylinders lymer specimens using Env. SEM (Quanta 400, FEI, France).
with a 1:2 diameter-to-length ratio [20,24,27]. Three cylind- The 27Al and 29Si NMR spectra were acquired using a
ers of each geopolymer were tested, enabling the calculation 600 MHz spectrometer with Bruker UltraShield supercon-
of average experimental values. ducting magnet. The sample was placed in a rotor rotating at
Thermal analysis (i.e., differential thermal analy- the magic angle and at a speed as high as 15 kHz.
sis–thermogravimetric analysis (DTA–TGA)) using a Netzsch
STA-409 E apparatus was carried out on hardened paste 3. Results and discussion
from ground geopolymers. Thermal analyses were per-
3.1. Mineralogical characterization of geopolymers
formed on a mass of 180.8 mg of powder under flowing air.
The functional groups present in the synthesized geopo- 3.1.1. Thermal analysis
lymers were identified by FTIR spectroscopy (from 4000 to The thermal resistance of SF2-geopolymer was studied
450 cm−1) using a Perkin Elmer spectrometer 180. by thermal analysis. Thermal behavior of SF2-G is shown in
Qualitative mineralogical analyses were performed by Fig. 2.

Fig. 2. Thermal analysis on SF2-G.

The DTA curve shows a large endothermic peak over the


range 25–300°C, with intense mass loss from the sample indi-
cated on the TGA curve. This peak is attributed to the evapora-
tion of free water [28]. The sample mass thereafter remained
relatively constant to temperatures in excess of 1200°C.
The exothermic peak at 898.2°C may correspond to the
transformation of the geopolymer to nepheline and carne-
gieite. These crystalline phases soften and partially melt at
temperatures above 1000°C, characterized by an endother-
mic peak centered at 1160.6°C [29]. The exothermic peak at
966.9°C corresponds to the recrystallization of unpolyme-
rized Mk to spinel [30].
3.1.2. XRD analysis Fig. 3. XRD patterns on metakaolin (Mk) and the geopoly-
Fig. 3 compares the XRD pattern of Mk to that of the mer specimens: A-G, B-G, and SF2-G. I⎯Illite; Q⎯Quartz;
manufactured geopolymer specimens cured at room S⎯Sodium aluminosilicate.
A. Nmiri et al., Replacement of alkali silicate solution with silica fume in metakaolin-based geopolymers 559

temperature for 28 d. The synthesized geopolymers without and/or aluminosilicate structures in the Mk, combined with
added silica fume (labeled as A-G) or with the the substitution of some Si by Al. This depolymerization is
NaOH/Na2SiO3 mixture (labeled as B-G) were faced to the followed by the formation of a tetrahedral framework cha-
geopolymer using 10 M NaOH as activator with silica fume. racteristic of a geopolymer system. That is, the shift of the
Because the XRD patterns of the three tested SF-geopolymers band at 1093 cm−1 indicates dissolution of the Mk amorph-
with increasing amount of silica fume were similar, the ous phase in the alkaline-activating solutions and the forma-
SF2-geopolymer pattern shown in Fig. 3 was chosen as a tion of a new product as an amorphous aluminosilicate gel
common representation. phase [33–34].
The XRD pattern of Mk consists of a broad peak cen-
tered around 22° (2θ), which indicates the amorphous cha-
racter of Mk. The sharp peaks at 20.85, 26.93, and 50.36°
(2θ) are attributed to quartz, and those at 8.70 and 3.85° (2θ)
are attributed to illite.
The alkaline activation of a dry mixture of Mk with silica
fume caused little change in the XRD pattern of Mk: the large
peak, characterizing the amorphous character of Mk, was dis-
placed from 24° < 2θ < 33° to 28° < 2θ < 40°. This indicates
the partial dissolution of the Mk amorphous particles and
the formation of a new amorphous aluminosilicate phase attri-
buted to the geopolymer systems [22,31]. This broad peak was
present for each formulation and was uniform, irrespective of
the silica fume content (from 0 to 10wt%). The formation of Fig. 4. FTIR spectra of metakaolin (Mk) and the geopolymer
specimens: A-G, B-G, and SF2-G (cured at room temperature
this new amorphous aluminosilicate phase was also clearly
for 28 d).
recognized by FTIR analysis, as discussed below.
In the presence of silica fume, a new small peak was specif- The appearance of new bands at 724 cm−1, attributed to
ically observed at 34.26° (2θ) on the SF-geopolymer specimen the symmetric stretching vibrations of Si–O–Si and
and was attributed to a likely sodium aluminosilicate. The inten- Al–O–Si [35–36] is associated with the formation of
sity of quartz peaks at 0.333, 0.181, and 0.424 nm and the illite amorphous aluminosilicate materials. The band located at
peak at 0.994 nm remained unchanged before and after activa- approximately 470 cm−1 is attributed to bending vibrations
tion, indicating that they did not participate in the reactions. of Si–O–Si and O–Si–O bonds [35–36], indicating that
3.1.3. FTIR analysis quartz is almost inert in the geopolymerization reaction. By
To confirm geopolymer formation, FTIR spectra were contrast, the bands in the regions of 1600 and 3450 cm−1,
recorded; the results are shown in Fig. 4. As with the XRD which are attributed to bending vibrations of H–O–H and
analysis, no difference was observed in the FTIR spectra for stretching vibrations of –OH, respectively, remain at their
the three FS-geopolymer formulations (therefore, only the positions observed in the spectrum of Mk.
spectrum recorded for the SF2-geopolymer is shown in Fig. 4). 3.1.4. NMR measurements
Notably, geopolymers, with or without silica fume and To confirm the formation of a geopolymer with a
whatever the alkaline activator, exhibit similar spectra. The three-dimensional structure, aluminum and silicon atom speci-
band at approximately 1405 cm−1 is assigned to the vibration ation was accurately established through 27Al and 29Si magic
of the C–O bond of sodium carbonate, stemming from the angle spinning (MAS) NMR. The spectra of the Mk-based
reaction between NaOH and atmospheric CO2, which in- SF2-geopolymer matrix synthesized with 10 M NaOH and
duces efflorescence inside materials. cured at room temperature for 28 d are presented in Fig. 5(a)
A comparison of the Mk spectrum with that after alkaline (for 27Al MAS NMR) and Fig. 5(b) (for 29Si MAS NMR).
activation, with or without silica fume, shows a shift of the The other SF-geopolymer matrixes behaved similarly.
vibrational band initially at 1093 cm−1 toward a lower wa- The 27Al MAS NMR spectrum shows that the Mk-based
venumber (approximately 1004 cm−1) in the spectra of the SF2-geopolymer presents a main peak at 60.96 ppm and a
geopolymers. Such behavior is due to the symme- small peak at 3.48 ppm. By contrast, the 29Si MAS NMR
tric-to-asymmetric stretching mode of T–O–Si (T = Si or Al) spectrum reveals a wide peak centered between −85.27 and
bonds and the depolymerization [32] of original silicate −86.82 ppm with a shoulder at −91.13 ppm. To interpret the
560 Int. J. Miner. Metall. Mater., Vol. 26, No. 5, May 2019

NMR results and in the absence of direct experimental to Al(VI) or Al(IV) respectively, whereas Al(V) should re-
measurements on Mk, it is important to remember that kao- sonate at 30 ppm. After geopolymerization, the geopolymer
linitic clay has essentially aluminum atoms with a coordina- matrixes usually contain chemical shifts of not only
tion number equal to VI. Thermal treatment of the kaolinitic AlQ4(4Si) (Al linked to four Si atoms via four oxygens) but
clay leads to the formation of IV- and V-coordinate alumi- also of SiQ4(mAl) (Si linked to m Al atoms via four oxy-
num atoms [37]. Thus, the 27A1 MAS NMR spectra of Mk gens), with 0 ≤ m ≤ 4 [3]. In the Mk-based geopolymer, the
usually present three signals centered at 0, 30, and 60 ppm. main chemical shift of aluminum at approximately 55–60
Many studies [3,37] have assigned the signal at 0 or 60 ppm ppm corresponds to AlQ4(4Si) [3].

Fig. 5. (a) 27Al MAS NMR and (b) 29Si MAS NMR spectra of a metakaolin-based SF2-geopolymer. δ⎯Chemical shift.

The experimental results in Fig. 5 are consistent with a SiO2/Al2O3 molar ratio of 3.84. It reaches 35.00 MPa,
such behavior. The SF2-geopolymer contains AlQ4(4Si) whereas the lowest compressive strength is 14.47 MPa,
coordination characterized by the peak centered at 60.96 which corresponds to the geopolymer mixture with 10% si-
ppm (>60 ppm), and trace amounts of Al(VI) characterized lica fume (SF3-G). In this case, the SiO2/Al2O3 ratio is 4.09.
by the small peak centered at 3.48 ppm (the position of this An intermediate compressive strength was obtained for
peak did not increase, as expected on the basis of the litera- SF1-G with 2% silica fume and SiO2/Al2O3 = 3.61. These
ture). The presence of a clear peak near 3.48 ppm means that results demonstrate the existence of an optimum silica fume
unreacted Mk may be present (perhaps because of insufficient content of approximately 6%. In parallel, the compressive
concentration of NaOH set at 10 M). Singh et al. [38] have strengths of the Mk-based geopolymers without added silica
reported a similar result. fume and with NaOH (in A-G) and NaOH/Na2SiO3 (in B-G)
For the 29Si MAS NMR, the spectra of kaolinitic clay and as activators with SiO2/Al2O3 = 3.39 and 3.54, respectively,
Mk consist of a single resonance peak at −91.5 ppm as- reached 33.75 MPa and 20.42 MPa, respectively. Replacing
signed to the SiQ3(3Si) coordination [37,39–40]. By contrast, a fraction of the Mk (requiring energy for production) with
the Mk-based geopolymer usually presents a broad peak silica fume (an industrial byproduct) gives almost the same
centered at −94.5 ppm assigned to the SiQ4(mAl) coordina- mechanical performance.
tion [3]. The experimental results differ from those reported
in the literature. Whereas the shoulder observed at −91.13
ppm could be associated with unreacted Mk (as observed in
the 27Al MAS NMR spectrum), the peak associated with the
SiQ4(3Al) coordination was positioned at −85.27 ppm [3].
3.2. Physical characterization of SF-geopolymers
3.2.1. Compressive strength
Compressive strength has been used by many re-
searchers [14,41] as a tool for confirming the success of
geopolymerization. The compressive strengths of the tested
geopolymer specimens are shown in Fig. 6 (values are av-
erage measurements calculated from three replicates). Fig. 6. Compressive strength of geopolymer specimens: A-G,
The specimen with the highest compressive strength after B-G, SF1-G, SF2-G, and SF3-G (cured at room temperature for
curing for 28 d (SF2-G) contains 6wt% silica fume and with 28 d).
A. Nmiri et al., Replacement of alkali silicate solution with silica fume in metakaolin-based geopolymers 561

The addition of Na2SiO3 to NaOH in the B-G specimen hardened geopolymer. Generally, a low SiO2/Al2O3 ratio
failed to give higher compressive strength than that meas- (below 3.4) leads to low-strength products accompanied by
ured on the A-G or SF2-G specimens. In fact, the addition a microstructure with more Na–Al–Si grains rather than
of sodium silicate to the NaOH solution increases the vis- amorphous Na–Al–Si, which is associated with geopolymers.
cosity of the total solution and reduces the flow of the By contrast, an increase in the SiO2/Al2O3 molar ratio to 3.4
geopolymer [42]. Thus, the reaction is incomplete and the or 3.8 is strongly related to the improvement in compressive
dissolution of aluminosilicate is insufficient. The paste har- strength [14,44]. Such results agree with those of Bernal
dens quickly, meaning that part of the Mk particles remain et al. [27], who studied the effect of granulated blast-furnace
unreacted and the polycondensation process is slow, requir- slag addition on the compressive strength of Mk-based
ing the addition of water to increase geopolymer paste flo- geopolymer. They tested SiO2/Al2O3 molar ratios of 3.0, 3.4,
wage. Therefore, the contact between the activation solution 3.8, and 4.0 with NaOH/Na2SiO3 as an alkaline activator
and the Mk particles is limited because of the large volume and found that formulations with SiO2/Al2O3 ratios between
of liquid used, which weakens the compressive strength of 3.4 and 4.0 exhibited the highest compressive strengths.
the geopolymer. A high L/S ratio improves the workability 3.2.2. SEM observation
of a fresh geopolymer mixture, but reduces its compressive The Env. SEM technique was used to study variations in
strength [22,34]. According to Zhang et al. [43], as little the texture (particle organization) and porosity (density) of
water as possible should be added to ensure high mechanical geopolymer materials as the amount of silica fume added to
strength and excellent durability. a geopolymer mixture was varied. Fig. 7 shows the SEM
These results also confirm the importance of the SiO2/Al2O3 micrographs of the geopolymer specimens cured at room
molar ratio, which affects the compressive strength of the temperature for 28 d.

Fig. 7. SEM micrographs of geopo-


lymer test specimens cured at room
temperature for 28 d: (a) A-G; (b) B-G;
(c) SF1-G; (d) SF2-G; (e) SF3-G.

The morphology and particle size of the A-G geopolymer 0.2 µm), helps reduce the volume of voids, which directly
varies after the addition of sodium silicate to NaOH, as ob- affect geopolymer strength. The density and porosity in
served for the B-G geopolymer. The A-G fractured surface SF-geopolymers, however, cannot be quantified from Env.
reveals a denser matrix with fine particles, whereas the B-G SEM images, nor can the presence or absence of unreacted
material develops a heterogeneous matrix with larger par- Mk and silica particles [45]. No significant differences in
ticles and higher porosity. This difference contributes to the microstructures between the three geopolymers (SF1-G,
decrease in compressive strength (from 33.75 MPa for A-G SF2-G, and SF3-G) were clearly observed.
to 20.42 MPa for B-G). 3.2.3. MIP analysis
In the case of the SF-geopolymers, condensation appears To obtain a more quantitative image of the microstructure,
to be higher. Silica fume, because of its small size (around MIP was used to complement the Env. SEM observations and
562 Int. J. Miner. Metall. Mater., Vol. 26, No. 5, May 2019

to explain the effect of silica fume on the mechanical proper- the total volume intruded. The physical properties of geo-
ties of Mk-based geopolymers. Under the assumption of a polymer calculated from MIP analysis are given in Table 4.
regular pore geometry and pore interconnection, we measured Additionally, Fig. 8 shows the pore size density function
porosity by the volume of mercury intruded into freeze-dried (corresponding to the differential calculation of the intruded
hardened specimens. Each incremental pressure corresponds volume of mercury (dV) versus the logarithm of pore di-
to a pore diameter and a total porosity is determined from ameter (dlgD)) versus the pore diameter in the geopolymers.

Table 4. Physical proprieties of geopolymers determined through mercury intrusion porosimetry


Geopolymer Total intrusion volume / Pore diameter of the main pore Calculated apparent particle density /
Total porosity / %
specimen (mL⋅g–1) family / nm (g⋅cm–3)
A-G 0.20 174 2.20 30.6
B-G 0.25 445–500 2.24 36.2
SF1-G 0.18 195 2.15 28.6
SF2-G 0.21 220 2.22 31.6
SF3-G 0.24 250 2.16 33.7

at 250 nm, which is linked to the amount of unreacted ma-


terial in the specimens. Furthermore, particle morphology
and the arrangement and porosity are key microstructural
parameters that govern the mechanical properties of the
geopolymers, irrespective of precursor type or mixture
composition, as noted by Latella et al. [20]. The develop-
ment of compressive strength is related to the evolution of
pore structure [46]. Here the increase in pore diameter of the
B-G and SF3-G geopolymers (measured as 445–500 and
250 nm, respectively) and their total porosity (evaluated as
Fig. 8. MIP curves of A-G, B-G, SF1-G, SF2-G, and SF3-G 36.2 and 33.7%, respectively) are correlated to the decrease
geopolymer specimens.
in compressive strength compared to those of the SF2-G and
The mercury intrusion curves of all geopolymers are A-G geopolymers. The lower mechanical strength of the
characterized by unimodal behavior with a major population SF1-G geopolymer, even though it displayed the lowest po-
of macropores (defined by the International Union of Pure rosity, must be explained by other means.
and Applied Chemistry as pores larger than 50 nm). How-
ever, the position of the peak characterizing the most com- 4. Conclusions
mon pore size within the pore families shifts with the type
of geopolymer mixture. As observed by Env. SEM, the The aim of this work was to study the feasibility of re-
geopolymer B-G presents a pore family with the largest pore placing the alkali silicate solution usually added to
size, centered at 445–500 nm. In comparison, the A-G and Mk-based geopolymer paste with silica fume. The proper-
SF-geopolymers display mainly smaller pores, centered ties of SF-geopolymers with 2, 6, or 10wt% of silica fume
between 174 and 250 nm. activated by NaOH were compared with those of geopoly-
The addition of silica fume affects the SiO2/Al2O3 molar mers without added silica fume (A-G geopolymer) or acti-
ratios, which strongly influence the mechanical properties of vated with NaOH mixed with Na2SiO3 (B-G geopolymer).
the geopolymers. The mercury intrusion curves of samples The FTIR spectra of the SF-geopolymers revealed a band
of SF1-G, SF2-G, and SF3-G with corresponding shift from 1093 cm−1 (in the spectrum of Mk) to ~1004 cm−1
SiO2/Al2O3 molar ratios of 3.61, 3.84, and 4.09, respectively, (in the spectra of geopolymers). A displacement of the broad
show that a substantial change in pore size accompanied the peak from 24° < 2θ < 33° to 28° < 2θ < 40° was also ob-
change in the SiO2/Al2O3 molar ratio. As shown in Fig. 8, served in the XRD patterns. Moreover, the 29Si and 27Al
the mercury intrusion curve of SF3-G, which exhibited the MAS NMR spectra of the SF-geopolymers showed chemi-
lowest compressive strengths, presents a pore family with cal shifts at approximately −85.27 ppm and 60.96 ppm, cor-
the highest pore size among the SF-geopolymers, centered responding to SiQ4(3Al) and AlQ4(4Si) coordination, re-
A. Nmiri et al., Replacement of alkali silicate solution with silica fume in metakaolin-based geopolymers 563

spectively. These latter three results indicated the formation on physical-thermal properties of clay, Thermochim. Acta,
of geopolymer material. 666(2018), p. 148.
The mechanical properties of the geopolymers were cor- [10] V. Medri, S. Fabbri, J. Dedecek, Z. Sobalik, Z. Tvaruzkova,
and A. Vaccari, Role of the morphology and the dehydrox-
related with their microstructural characteristics to explain
ylation of metakaolins on geopolymerization, Appl. Clay Sci.,
the various behaviors observed in the specimens. The mor- 50(2010), No. 4, p. 538.
phology (observed by Env. SEM) and porosity (given by [11] Y.M. Liew, H. Kamarudin, A.M. Mustafa Al Bakri, M.
MIP) are the main microstructural characteristics determin- Luqman, I. Khairul Nizar, C.M. Ruzaidi, and C.Y. Heah,
ing the compressive strength of the SF-geopolymers, along Processing and characterization of calcined kaolin cement
with the SiO2/Al2O3 molar ratio, porosity level, and pore powder, Constr. Build. Mater., 30(2012), p. 794.
diameter. The highest compressive strength (35 MPa) was [12] Y.M. Liew, H. Kamarudin, A.M. Mustafa Al Bakri, M. Bin-
attained with the addition of 6wt% silica fume (the hussain, M. Luqman, I. Khairul Nizar, C.M. Ruzaidi, and
C.Y. Heah, Influence of solids-to-liquid and activator ratios
SiO2/Al2O3 molar ratio was 3.84), Finally, on the basis of
on calcined kaolin cement powder, Phys. Procedia, 22(2011),
observations of the physical, microstructural, and thermal p. 312.
properties, the silica fume geopolymers were found to be [13] C.Y. Heah, H. Kamarudin, A.M. Mustafa Al Bakri, M.
much more sustainable than sodium silicate geopolymers. Bnhussain, M. Luqman, I. Khairul Nizar, C.M. Ruzaidi, and
Y.M. Liew, Kaolin-based geopolymers with various NaOH
Acknowledgement concentrations, Int. J. Miner. Metall. Mater., 20(2013), No. 3,
p. 313.
The authors are pleased to acknowledge the Tunisian [14] P. De Silva, K. Sagoe-Crenstil, and V. Sirivivatnanon, Kinet-
ics of geopolymerization: Role of Al2O3 and SiO2, Cem.
Ministry of Higher Education and Scientific Research for its
Concr. Res., 37(2007), No. 4, p. 512.
help in financing internships in France. [15] F.N. Okoye, J. Durgaprasad, and N.B. Singh, Effect of silica
fume on the mechanical properties of fly ash
References based-geopolymer concrete, Ceram. Int., 42(2015), No. 2, p.
3000.
[1] M. Mustafa, A. Bakri, H. Mohammed, H. Kamarudin, I.K. [16] D. Dutta, S. Thokchom, P. Ghosh, and S. Ghosh, Effect of si-
Niza, and Y. Zarina, Review on fly ash-based geopolymer lica fume additions on porosity of fly ash geopolymers, J.
concrete without Portland Cement, J. Eng. Technol. Res., Eng. Appl. Sci., 5(2010), No. 10, p. 74.
3(2011), No. 1, p. 1. [17] C.S. Poon, S.C. Kou, and L. Lam, Compressive strength ,
[2] M.A. Villaquirán-caicedo, R.M. de Gutiérrez, S. Sulekar, C. chloride diffusivity and pore structure of high performance
Davis, and J.C. Nino, Thermal properties of novel binary metakaolin and silica fume concrete, Constr. Build. Mater.,
geopolymers based on metakaolin and alternative silica 20(2006), No. 10, p. 858.
sources, Appl. Clay Sci., 118(2015), p. 276. [18] P. Duan, C.J. Yan, and W. Zhou, Compressive strength and
[3] J. Davidovits, Properties of geopolymer cements, [in] First microstructure of fly ash based geopolymer blended with si-
International Conference on Alkaline Cements and Con- lica fume under thermal cycle, Cem. Concr. Compos.,
cretes, Kiev, 1994, p. 131. 78(2017), p. 108.
[4] H.T. Huynh, New generation geopolymers, [in] 2nd French [19] M. Uysal, M.M. Al-mashhadani, Y. Aygörmez, and O. Can-
Seminar on Geopolymers, Clermont-Ferrand, 2013. polat, Effect of using colemanite waste and silica fume as
[5] Y.M. Liew, C.Y. Heah, A.B. Mohd Mustafa, and H. Kama- partial replacement on the performance of metakaolin-based
rudin, Structure and properties of clay-based geopolymer ce- geopolymer mortars, Constr. Build. Mater., 176(2018), p.
ments: A review, Prog. Mater. Sci., 83(2016), p. 595. 271.
[6] P. Zhang, Y.X. Zheng, K.J. Wang, and J.P. Zhang, A review [20] B.A. Latella, D.S. Perera, D. Durce, E.G. Mehrtens, and J.
on properties of fresh and hardened geopolymer mortar, Davis, Mechanical properties of metakaolin-based geopoly-
Composites Part B, 152(2018), p. 79. mers with molar ratios of Si/Al ≈ 2 and Na/Al ≈ 1, J. Mater.
[7] C. Dupuy, A. Gharzouni, N. Texier-Mandoki, X. Bourbon, Sci., 43(2008), No. 8, p. 2693.
and S. Rossignol, Thermal resistance of argillite-based alka- [21] M. Rostami and K. Behfarnia, The effect of silica fume on
li-activated materials. Part 1: Effect of calcination processes durability of alkali activated slag concrete, Constr. Build.
and alkali cation, Mater. Chem. Phys., 217(2018), p. 323. Mater., 134(2017), p. 262.
[8] M.I. Khan, H.U. Khan, K. Azizli, S. Sufian, Z. Man, A.A. [22] D. Panias, I.P. Giannopoulou, and T. Perraki, Effect of syn-
Siyal, N. Muhammad, and M.F. ur Rehman, The pyrolysis thesis parameters on the mechanical properties of fly
kinetics of the conversion of Malaysian kaolin to metakaolin, ash-based geopolymers, Colloids Surf. A, 301(2007), No. 1-3,
Appl. Clay Sci., 146(2017), p. 152. p. 246.
[9] J.S. Geng and Q. Sun, Effects of high temperature treatment [23] Z.H. Zhang, X. Yao, H.J. Zheng, and C. Yue, Role of water
564 Int. J. Miner. Metall. Mater., Vol. 26, No. 5, May 2019

in the synthesis of calcined kaolin-based geopolymer, Appl. polymer, Open Civ. Eng. J., 7(2013), p. 84.
Clay Sci., 43(2009), No. 2, p. 218. [35] W.K.W. Lee and J.S.J. van Deventer, Structural reorganisa-
[24] J.G.S. van Jaarsveld and J.S.J. van Deventer, Effect of the al- tion of class F fly ash in alkaline silicate solutions, Colloids
kali metal activator on the properties of fly ash based geopo- Surf. A, 211(2002), No. 1, p. 49.
lymers, Ind. Eng. Chem. Res., 38(1999), No. 10, p. 3932. [36] T. Bakharev, Geopolymeric materials prepared using Class F
[25] P.N. Lemougna, A.B. Madi, E. Kamseu, U.C. Melo, M.P. fly ash and elevated temperature curing, Cem. Concr. Res.,
Delplancke, and H. Rahier, Influence of the processing tem- 35(2005), No. 6, p. 1224.
perature on the compressive strength of Na activated lateritic [37] J. Rocha and J. Klinowski, 29Si and 27Al mag-
soil for building applications, Constr. Build. Mater., ic-angle-spinning NMR studies of the thermal transformation
65(2014), p. 60. of kaolinite, Phys. Chem. Miner., 17(1990), No. 2, p. 179.
[26] A.M.M. Al Bakri, H. Kamarudin, M. Bnhussain, J. Liyana, [38] P.S. Singh, M. Trigg, I. Burgar, and T. Bastow, Geopolymer
and C.M. Ruzaidi Ghazali, Nano geopolymer for sustainable formation processes at room temperature studied by 29Si and
27
concrete using fly ash synthesized by high energy ball mil- Al MAS-NMR, Mater. Sci. Eng. A, 396(2005), No. 1-2, p.
ling, Appl. Mech. Mater., 313-314(2013), p. 169. 392.
[27] S.A. Bernal, E.D. Rodríguez, R. Mejía de Gutiérrez, M. Gor- [39] F. Škvára, L. Kopecký, J. Němeče k, and Z. Bittnar, Micro-
dillo, and J.L. Provis, Mechanical and thermal characterisa- structure of geopolymer materials based on fly ash, Ceram.
tion of geopolymers based on silicate-activated metakao- Silik., 50(2006), No. 4, p. 208.
lin/slag blends, J. Mater. Sci., 46(2011), No. 16, p. 5477. [40] A. Bourlon, Physico-chimie et Rhéologie de Géopolymères
[28] T. Revathi, R. Jeyalakshmi, and N. P. Rajamane, Study on Frais Pour la Cimentation des Puits Pétroliers [Dissertation],
the role of n-SiO2 incorporation in thermo-mechanical and Pierre et Marie Curie University, Paris, 2010.
microstructural properties of ambient cured FA-GGBS geo- [41] M.A. Soleimani, R. Naghizadeh, A.R. Mirhabibi, and F. Go-
polymer matrix, Appl. Surf. Sci., 449(2018), p. 322. lestanifard, Effect of calcination temperature of the kaolin
[29] C. Kuenzel, L.M. Grover, L. Vandeperre, A.R. Boccaccini, and molar Na2O/SiO2 activator ratio on physical and micro-
and C.R. Cheeseman, Production of nepheline/quartz ceram- structural properties of metakaolin based geopolymers, Iran.
ics from geopolymer mortars, J. Eur. Ceram. Soc., 33(2013), J. Mater. Sci. Eng., 9(2012), No. 4, p. 43.
No. 2, p. 251. [42] P. Chindaprasirt, T. Chareerat, and V. Sirivivatnanon, Wor-
[30] G. Kakali, T. Perraki, S. Tsivilis, and E. Badogiannis, Ther- kability and strength of coarse high calcium fly ash geopoly-
mal treatment of kaolin: The effect of mineralogy on the mer, Cem. Concr. Compos., 29(2007), No. 3, p. 224.
pozzolanic activity, Appl. Clay Sci., 20(2001), No. 1-2, p. 73. [43] Y.S. Zhang, W. Sun, and Z.J. Li, Composition design and
[31] S. Ahmari, X. Ren, V. Toufigh, and L.Y. Zhang, Production microstructural characterization of calcined kaolin-based
of geopolymeric binder from blended waste concrete powder geopolymer cement, Appl. Clay Sci., 47(2010), No. 3-4, p.
and fly ash, Constr. Build. Mater., 35(2012), p. 718. 271.
[32] W.K.W. Lee and J.S.J. van Deventer, Use of infrared spec- [44] C.K. Ma, A.Z. Awang, and W. Omar, Structural and material
troscopy to study geopolymerization of heterogeneous performance of geopolymer concrete: A review, Constr.
amorphous aluminosilicates, Langmuir, 19(2003), No. 21, p. Build. Mater., 186(2018), p. 90.
8726. [45] T. da S. Rocha, D.P. Dias, F.C.C. França, R.R. de S. Guerra,
[33] Q. Wan, F. Rao, S.X. Song, D.F. Cholico-González, and N.L. and L.R. da C. de O. Marques, Metakaolin-based geopolymer
Ortiz, Combination formation in the reinforcement of meta- mortars with different alkaline activators (Na+ and K+),
kaolin geopolymers with quartz sand, Cem. Concr. Compos., Constr. Build. Mater., 178(2018), p. 453.
80(2017), p. 115. [46] P. Duan, Z.G. Shui, W. Chen, and C.H. Shen, Effects of me-
[34] K. Gao, K.L. Lin, D.Y. Wang, H.S. Shiu, C.L. Hwang, and takaolin , silica fume and slag on pore structure, interfacial
T.W. Cheng, Effects of nano-SiO2 on setting time and com- transition zone and compressive strength of concrete, Constr.
pressive strength of alkali- activated metakaolin-based geo- Build. Mater., 44(2013), p. 1.

You might also like