You are on page 1of 14

Accepted Manuscript

Synthesis of ceramic materials from ecofriendly geopolymer precursors

M.A. Villaquirán-Caicedo, R. Mejía de Gutiérrez

PII: S0167-577X(18)31171-6
DOI: https://doi.org/10.1016/j.matlet.2018.07.128
Reference: MLBLUE 24694

To appear in: Materials Letters

Received Date: 20 December 2017


Revised Date: 17 July 2018
Accepted Date: 28 July 2018

Please cite this article as: M.A. Villaquirán-Caicedo, R.M. de Gutiérrez, Synthesis of ceramic materials from
ecofriendly geopolymer precursors, Materials Letters (2018), doi: https://doi.org/10.1016/j.matlet.2018.07.128

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 proof before it is published in its final 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.
Synthesis of ceramic materials from ecofriendly geopolymer precursors

M.A. Villaquirán-Caicedo* & R. Mejía de Gutiérrez,

Composite Materials Group (CENM), School of Materials Engineering, Universidad


del Valle, Calle 13 # 100-00, Cali, Colombia

*Email corresponding: monica.villaquiran@correounivalle.edu.co

Abstract

Geopolymer materials based on metakaolin were produced using KOH and


ecofriendly silica sources, such as rice husk ash and silica fume. The geopolymer
produced by ecofriendly alkaline activators had mechanical properties that are
similar to the geopolymer synthesized with commercial potassium silicate. Upon
exposure of the geopolymer to high temperatures, it maintained thermal stability.
Its residual compression strength after exposure to 600 °C reached 50 MPa, and at
1200 °C, it was 110 - 150 MPa. Exposure to high temperature resulted in sintering
processes, structural rearrangement and formation of crystalline phases such as
leucite, kalsilite and mullite. Additionally, upon production of 1 m3 of ecofriendly
geopolymer paste, the global warming potential emissions were reduced by
47.45% (RHA100) and 39.4% (SF100) compared with commercial potassium
silicate.

Keywords: ceramics, geopolymer, rice husk ash, microstructure, mechanical


properties, sintering.

1. Introduction

Geopolymers are produced from aluminosilicate materials (minerals and


industrial by-products) and are synthesized in a highly alkaline medium at low
temperature (<100 °C). These materials may replace Portland cement due to their
properties, such as high strength at early ages, and low thermal and acoustic
conductivities, which are similar or even superior to Portland cement [1,2]. It has
been estimated that the production of geopolymers allows the reduction of
greenhouse gas emissions by ~70% in comparison with the production of Portland
cement, which makes geopolymers environmentally friendly [3]. However, the use
of a commercial silicate (Na2O3Si or K2O3Si) as an activator suggests that the
carbon footprint of geopolymers is almost the same as that generated during the
production of Portland cement because the production of commercial silicates
1
involves burning sodium carbonate and silica sand at temperatures above 1100 °C.
Recently, research has been focused on the production of alternative silicates to
make the geopolymerization process more environmentally sustainable. Secondary
sources of silica can be obtainable from industrial by-products, such as silica fume
[4,5], glass waste [6,7] and rice husk ash [8,9].
In this investigation, we used 100% rice husk ash (RHA) and 100% silica fume
(SF) as an ecofriendly silica-modifying source to synthesize geopolymer pastes
based on metakaolin. The thermal stability, compressive strength, and
microstructure were evaluated. Additionally, a life cycle analysis was carried out for
the geopolymer pastes.

2. Experimental Procedure

2.1 Materials
Commercial metakaolin MetaMax® (MK) from BASF was used as the precursor.
RHA was obtained from rice husk burned in an electrical furnace at 600 °C for 2 h
and subsequently grounded in a ball mill for 30 minutes. For comparison, SF
(SikaFume®) and commercial potassium silicate (KS) were also used as silica
sources. The chemical compositions of these raw materials are shown in Table 1.

Table 1. Chemical and physical properties of raw materials.

2.2 Specimen preparation and tests

Geopolymer systems were designed with molar ratios for SiO2/Al2O3 and
K2O/SiO2 of 2.5 and 0.28, respectively. The ecofriendly alkaline activator was
prepared by mixing analytical grade potassium hydroxide (KOH) with the RHA and
SF silica sources in proportions of 50 and 100%. The geopolymer pastes produced
were KS100, SF50, SF100, RHA50 and RHA100.
The components were mechanically mixed for 7 minutes and cured at 70 °C for
20 h at a relative humidity of ~90%. Subsequently, the samples were stored in a
humidity chamber at 25 °C. After 7 days, the geopolymer pastes were dried at 60
°C until a constant weight was achieved; then, they were exposed to 300, 600,
900, and 1200 °C at a heating rate of 1 °C/minute in an electric furnace. Isothermal
heating for 2 h was performed followed by natural cooling inside the oven. The
tests included a compressive strength test (ASTM C109) performed in a universal
testing machine (Instron® 3369), scanning electron microscopy (SEM) using a
JEOL JSM-6490LV and X-ray powder diffraction (XRD) using an X’Pert MRD
PANalytical diffractometer. A digital calibrator was used to measure the samples
before and after exposure to high temperatures.
Additionally, the global warming potential (GWP) index of the geopolymer pastes
was calculated as an indicator of environmental impact associated with the
production of 1 m3 of geopolymer paste. Life cycle assessment (LCA) methodology
was applied using the OpenLCA 1.4.2 software.

3. Results and Discussion

2
3.1 Morphologic and Microstructural Analysis

In the micrographs of the geopolymeric systems after exposure to 300, 600, and
900 °C (Figure 1) is distinguishable the presence of unreacted MK particles
embedded in geopolymeric gel, it is possible to observe a material with
homogeneously distributed pores. As the exposure temperature increases, a
structural densification is generally observed as a consequence of the dehydration
of the geopolymeric products formed and the structural changes generated. At 600
°C, volumetric shrinkage was 12.81% (RHA100) and 16.87% (SF100), although,
microcracks were not observed in SEM images. For all geopolymeric systems
exposed to 900 °C, the growth of microcracks was identified, which are generated
by the effect of capillary contraction during the dehydration and dehydroxylation of
the geopolymeric gel. The presence of microcracks are in a lower proportion for
RHA100 and can be attributed to a better distribution of the water loss in the
structure. These microcracks can grow and lead to a decrease in mechanical
strength at this temperature as explained later. Crack growth can be controlled
through the incorporation of particles and fibers [8].

Figure 1. SEM images of geopolymers after exposure to 300 °C, 600 °C and 900 °C
( :pores;  : microcracks).

X-ray patterns of geopolymer samples before and after exposure to high


temperatures are shown in Figure 2. The X-ray diffraction patterns of all samples at
25 °C and 600 °C showed a characteristic amorphous halo between 2θ values of
20–35°. The geopolymer gel is a potassium aluminosilicate-hydrated gel (K-A-S-
H), has a sharper peak attributed to anatase (TiO2, Inorganic Crystal Structure
Database, ICSD154604) from the original MK. Anatase was present as an impurity
and was not dissolved during alkaline activation.
After exposure to 1200 ºC, the densification of geopolymers occurred, due to
softening and viscous sintering [10,11,12] and crystalline structures had been
formed, including leucite (KAlSi2O6, Pattern diffraction File, PDF 00-038-1423),
kalsilite (KAlSiO4, American Mineralogist Crystal Structure Database, AMSCD
1874) and mullite (Al4.8O9.6Si1.2, PDF 15-776). Small leucite crystals of
approximately 1- 2 μm in size for an exposure time of 2 h at 1200 °C were
observed via SEM (Figure 3a and 3b). At 10 h of sintering, the sizes of the leucite
crystals increased (Figure 3c). The kalsilite content was notably higher in the
KS100 geopolymer; kalsilite is a transition phase during leucite hydrothermal
crystallization in Al, K and Si systems (Figure 2). The high intensity of the leucite
peak at 1200°C was due to the high content of soluble silicates in these systems.

Figure 2. X-ray diffractograms for geopolymer samples at 25 °C and after exposure to


high temperatures. (a) KS100, (b) RHA100, (c) SF100.

Figure 3. SEM images after exposure to 1200 °C at different times. Leucite crystals are
visible after attack with HF 3% wt for 45 sec. (a) KS100, 2 hours, (b) RHA100, 2 hours, (c)
RHA100, 10 hours.
3
3.2 Compressive strength

The compressive strength increased upon using an exposure temperature from


25 ºC to 1200 °C (Figure 4) due to the densification and crystallization of the
geopolymer gel. At 900 °C, all pastes exhibited a decrease in compressive
strength, and the drop was less for the RHA50 (45.74%) and RHA100 (37.86%)
pastes; however, for the reference matrix KS100, it was 75.43%. This loss of
strength at 900 °C is associated with the increased surface energy of the gel
caused by the loss of water from the surface and the appearance of small
pores in the structure of the geopolymer gel, causing the collapse of its structure
[13].
As a result of the reorganization of the gel and the formation of new crystalline
phases at 1200 °C, the compressive strength in each system increased as follows:
294.5% (KS100), 232.3% (RHA100), 151.04% (RHA50), 187.06% (SF100) and
53.40% (SF50). These results are comparable with those reported by [14] and [15]
using conventional silicates.

Figure 4. Residual compressive strengths of geopolymer pastes after exposure to high


temperatures.

3.4 Life Cycle Analysis

The GWP values for the geopolymer pastes fabricated with 100% commercial
potassium silicate, RHA and SF are reported in Table 2. As a reference, for the
geopolymer paste (KS100), the principal GWP contribution is associated with the
contribution from KOH and commercial potassium silicate. This contribution is due
to the high temperature of production of commercial potassium silicate (>1100 °C).
The CO2 contribution from the alternative activators based on RHA and SF is quite
low compared to commercial potassium silicate. For the production of 1 m3 of
geopolymer pastes, it was possible to reduce the GWP emissions by 47.45%
(RHA100) and 39.4% (SF100) by using alternative sources of silica that produce
ecofriendly geopolymers.

Table 2. Global warming potential (kg of CO2 eq) for 1 m3 of geopolymer pastes
produced.

4. Conclusions

The results obtained in this study demonstrated that using RHA and SF as a
source of silica in the development of geopolymers is feasible; hence, RHA and SF
can be used as a replacement for commercial potassium silicate in materials of
technological interest to produce a lower carbon footprint.
- Exposure of these geopolymers to high temperatures achieves compressive
strengths of up to 137.6 MPa (RHA100) and 148.5 MPa (SF100); this result is

4
primarily due to the densification and formation of crystalline products such as
leucite and mullite.
- Alkaline activators based on commercial potassium silicate provide a higher
carbon footprint (1089.027 kg of CO2 eq), whereas activators based on RHA
(emission inventory of 572.315 kg of CO2 eq) and SF (660.234 kg of CO2 eq)
are more ecofriendly.

Acknowledgements

The authors gratefully acknowledge financial support from Universidad del Valle
with project named “Bolsa concursable para la publicación de artículos 2017” (Cali,
Colombia).

5. References
[1] L. Zhan, Production of bricks from waste materials, Constr. Build. Mater. 47
(2014) 643–655.
[2] B. McLellan, R. Williams, J. Lay, A. Van Riessen, and G. Corder, Costs and
carbon emissions for geopolymer pastes in comparison to ordinary portland
cement, J. Clean. Prod. 19 (2011) 1080–1090.
[3] M. Weil, K. Dombrowski, and A. Buchwald, Geopolymers Life-cycle analysis
in Geopolymers: Structures, Processing Properties and Industrial
Applications. J. L. Provis and J. S. J. Van Deventer, Eds. Cambridge,
England, 2009.
[4] Radhakrishna, Eco-Efficient Masonry Bricks and Blocks. Elsevier, 2015.
[5] M. J. A. Mijarsh, M. A. Megat Johari, and Z. A. Ahmad, Compressive
strength of treated palm oil fuel ash based geopolymer mortar containing
calcium hydroxide, aluminum hydroxide and silica fume as mineral additives,
Cem. Concr. Compos. 60 (2015) 65–81.
[6] M. Torres-Carrasco and F. Puertas, Waste glass in the geopolymer
preparation. Mechanical and microstructural characterisation, J. Clean. Prod.
90 (2015) 397–408.
[7] A. I. Badanoiu, T. H. Al Saadi, S. Stoleriu, and G. Voicu, Preparation and
characterization of foamed geopolymers from waste glass and red mud,
Constr. Build. Mater. 84 (2015) 284–293.
[8] M.A. Villaquirán-Caicedo, R. Mejía de Gutiérrez, and N.C. Gallego, A Novel
MK-based Geopolymer Composite Activated with Rice Husk Ash and KOH :
Performance at High Temperature. Mater. Construcción. 67 326 (2017) 1–
13.
[9] C.L. Hwang and T.-P. Huynh, Investigation into the use of unground rice
husk ash to produce eco-friendly construction bricks, Constr. Build. Mater. 93
(2015) 335–341.
[10] M.A. Villaquirán-Caicedo, R. Mejía de Gutierrez, S. Sulekar, C. Davis, and J.
Nino. Thermal properties of novel binary geopolymers based on metakaolin

5
and alternative silica sources. Appl. Clay Sci.118 (2015) 276–282.
[11] P. G. He, D. Jia, T. Lin, M. Wang, and Y. Zhao. Effects of high-temperature
heat treatment on the mechanical properties of unidirectional carbon fiber
reinforced geopolymer composites. Ceram. Int. 36 4 (2010) 1447–1453.
[12] P. Duxson, G. C. Lukey, and J. S. J. Deventer. Physical evolution of Na-
geopolymer derived from metakaolin up to 1000 °C. J. Mater. Sci. 42 9
(2007) 3044–3054.
[13] D.L. Kong, J. G. Sanjayan, and K. Sagoe-crentsil, Factors affecting the
performance of metakaolin geopolymers exposed to elevated temperatures,
J. Mater. Sci. 43 (2008) 824–831.
[14] L. Yun-Ming, H. Cheng-Yong, L. Long-yua, J. Nur Ain, J. Tan Soo, and H.
Kamarudin, Formation of one-part-mixing geopolymers and geopolymer
ceramics from geopolymer powder, Constr. Build. Mater. 156 (2017) 9–18.
[15] T. Alomayri, F. U. A. Shaikh, and I. Low, Mechanical and thermal properties
of ambient cured cotton fabric-reinforced fly ash-based geopolymer
composites, Ceram. Int. 40 (2014) 14019–14028.

6
7
8
9
10
Table 1. Chemical and physical properties of raw materials.

Component
(% by weight) MK RHA SF KS
SiO2 51.52 92.33 92.33 26.38
Al2O3 44.53 0.18 0.18 --
TiO2 1.71 -- -- --
Fe2O3 0.48 0.17 0.17 --
Na2O 0.29 0.07 0.07 --
K2O 0.16 0.15 0.15 13.06
MgO 0.19 0.49 0.49 --
CaO 0.02 0.63 0.63 --
H2O -- -- -- 60.56
Loss of ignition 1.09 2.57 2.57 --
(LOI) (1000 °C)
Density (kg/m3) 2500 2140 2200
Average particle 7.8 22.84 <1
size (µm)

11
Table 2. Global warning potential (kg of CO2 eq) for 1 m3 of geopolymer pastes
produced.

Compone Kg of material used for fabrication Kg of CO2 eq. associated to


nts of 1 m3 geopolymer paste every compounds
KS100 RHA10 SF100 KS100 RHA1 SF10
0 00 0
MK 844.07 841.62
1 811.094 9 77.992 74.945 77.766
KOH 221.99 259.97 497.25 497.25 582.3
0 221.990 9 9 9 53
H2O 233.21 479.81
6 486.332 3 0.034 0.072 0.072
Silica 450.64 133.98 513.74
source 9 121.583 7 0 0.038 0.041
TOTAL 1089.0 572.31 660.2
27 5 34

12
Highlights
Ceramic materials were produced from geopolymer based on rice husk ash.
The materials developed show achieves compressive strengths of up to 137.6 MPa.
The alternative silica making the production of geopolymer more ecofriendly.

13

You might also like