Curing Temp Kaolin

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Physics Procedia 22 (2011) 286 291

2011 International Conference on Physics Science and Technology (ICPST 2011)

The Effect of Curing Temperature on Physical and Chemical Properties of Geopolymers


A.M. Mustafa Al Bakria a*, H. Kamarudina, M. BinHussainb, I.Khairul Nizarc, Y. Zarinaa, A.R. Rafizaa
Centre of Excellence Geopolymer System Research, School of Material Engineering, Universiti Malaysia Perlis (UniMAP), 01000, P.O.Box 77, D/A Pejabat Pos Besar, Kangar, Perlis, Malaysia b King Abdul Aziz Science & Technology (KACST), P.O. Box 6086, Riyadh 11442, Kingdom of Saudi Arabia c School of Environmental Engineering, Universiti Malaysia Perlis (UniMAP), 01000, P.O. Box 77, D/A Pejabat Pos Besar, Kangar, Perlis, Malaysia
a

Abstract Fly ashbased geopolymers required heat to increase the geopolymerization process in order to obtain higher compressive strength. As such, geopolymer samples were prepared using different curing temperatures (room temperature, 50C, 60C, 70C, 80C), in which sodium silicate and NaOH were used as alkaline activators. The samples were cured for 24 hours in the oven and tested on the seventh day. The result revealed that the maximum compressive strength (67.04 MPa) was obtained at a temperature of 60C. However when the geopolymers sample cured at temperature more than 60C, the compressive strength decreased. From the FTIR spectra, the higher content of Si on sample cured at 60C also contributed to higher compressive strength. Moreover, SEM analysis showed a denser matrix as well as less unreacted fly ash of the sample cured at 60C compared to other temperatures.

by Elsevier B.V. Selection and/or peer-review under responsibility of Garry Lee. organizer] 2011 2011Published Published by Elsevier Ltd. Selection and/or peer-review under responsibility of [name
Keywords: fly ash; geopolymer; curing temperature;compressive strength,;FTIR; SEM

1. Introduction Geopolymers have been introduced to replace the utilization of ordinary Portland cement (OPC) in the concrete industry, where fly ash-based geopolymers have been known to have a high potential for

* Corresponding author. Tel.: +6012-5055020; fax: +604-9798178. E-mail address: mustafa_albakri@unimap.edu. my

1875-3892 2011 Published by Elsevier B.V. Selection and/or peer-review under responsibility of Garry Lee. doi:10.1016/j.phpro.2011.11.045

A.M. Mustafa Al Bakri et al. / Physics Procedia 22 (2011) 286 291

287

replacing the application of OPC. This material was introduced due to the extensive consumption of energy during the manufacturing of OPC, which releases a large amount of greenhouse gases into the atmosphere [1]. It has been reported that 13,500 million tons are produced from this process worldwide, accounting for approximately 7% of the greenhouse gases annually produced [1-2]. Since fly ash is the source of silica and alumina required for geopolymers and is excessively available worldwide, the used of fly ash will reduce problems with dumping sites for these waste materials. The silica and alumina in the fly ash are activated by an alkaline activator that consists of sodium silicate and sodium hydroxide to produce aluminosilicate gel that acts as a binder. Previous research has shown that geopolymers possess excellent mechanical properties, fire resistance, and acid resistance [2-5]. The geopolymer curing process is important for determining the existence of water during alkaliactivated fly ash matrix hardening [6] because water is a key element in alkaline activation reactions and curing regime humidity has an obvious effect on the structural and mechanical properties of alkaliactivated fly ash pastes, mortars, and concretes [6-8]. In addition, Palomo et al. [9] reported that the curing temperature was a reaction accelerator in fly ash-based geopolymers. Heat curing was also required to produce a fast geopolymerization process to achieve an acceptable strength within very short periods [10-13]. J. Temuujin et al. [14] reported that using curing temperatures between 40C and 80C for 4 to 48 hours is one of the important conditions for the synthesis of geopolymers. Hardjito et al. [15] mentioned that a higher curing temperature does not necessarily ensure that the compressive strength of the product will be higher. The current works objective is to investigate the effect of different curing temperatures on physical and chemical properties of a fly ash-based geopolymer. 2. Materials and Experimental Details 2.1. Raw Materials Fly ash was obtained from Manjung power station, Lumut, Perak, Malaysia. The chemical composition of the fly ash is tabulated in Table 1, which was obtained from an XRF analysis. According to the analysis, the fly ash was equivalent to ASTM class F. The sodium silicate solution was supplied by South Pacific Chemicals Industries Sdn. Bhd. (SPCI), Malaysia. The chemical compositions were comprised of 30.1% SiO2, 9.4% Na2O, and 60.5% H2O with a SiO2/Na2O modulus of 3.2, specific gravity at 20C = 1.4mg/cc, and viscosity at 20C = 400cP. Sodium hydroxide (NaOH) pellets with 99% purity were used after being diluted using distilled water to obtain 12 M concentration [9, 16]. The combination of sodium silicate and NaOH solution was used as an alkaline activator to induce the pozzolanic properties of fly ash.
Table 1. Chemical composition of fly ash Chemical Composition SiO2 Al2O3 Fe2O3 TiO2 CaO MgO Na2O K2O P2O5 SO3 MnO Percentage (%) 52.11 % 23.59 % 7.39 % 0.88 % 2.61 % 0.78 % 0.42 % 0.80 % 1.31 % 0.49 % 0.03

288

A.M. Mustafa Al Bakri et al. / Physics Procedia 22 (2011) 286 291

2.2. Sample Preparation The alkaline activator was prepared by mixing a sodium silicate and NaOH solution with a concentration of 12 M. The fly ash was then mixed with the alkaline activator in the mixer. The fresh geopolymer was placed in the 50x50x50 mm mold and vibrated for 20s on the vibrating table. Samples were subsequently cured at different curing temperatures; room temperature, 50C, 60C, 70C, and 80C for 24 hours. The average of three samples was prepared for each curing temperature. The details of the mix design are shown in Table 2.
Table 2. Mix design details Mixture No. 1 2 3 4 5 Fly ash/alkaline activator(FA/AA) 2.0 2.0 2.0 2.0 2.0 Sodium silicate/NaOH 2.5 2.5 2.5 2.5 2.5 Fly Ash(g) 562.5 562.5 562.5 562.5 562.5 Sodium silicate (g) 200.9 200.9 200.9 200.9 200.9 NaOH (g) 80.4 80.4 80.4 80.4 80.4 Curing Temperature Room temperature 50C 60C 70C 80C

2.3. Testing The hardened geopolymer samples were analyzed according to three criteria: mechanical strength (compressive strength), mineralogically, and microstructurally studied by Fourier Transform Infrared Spectroscopy (FTIR), and Scanning Electron Microscope (SEM). Compressive strength was tested on the seventh day. 3. Results and Discussion 3.1. Compressive Strength The compressive strength of fly ash-based geopolymers on the seventh day with different curing temperatures is shown in Fig. 1. As clearly shown in the graph, the samples cured at 60C produced the maximum compressive strength (67.04MPa) compared to others. Meanwhile, the minimum compressive strength was obtained from room temperature (22.9 MPa). Based on this result, when the curing temperature was increased, the compressive strength also increased. However, the value of compressive strength tended to decrease when cured at a temperature greater than 60C. Hardjito et al. [17] also found that increasing the curing temperature beyond 60C did not substantially increase the compressive strength. Hence, some researchers have suggested that the curing temperature for fly ash and kaolinite geopolymers should be 60C, especially for the study of geopolymer paste with 50mm cube samples [1819]. Chindaprasirt et al. [1] suggested that the small specimen with a high surface-to-volume ratio is more susceptible to the heat of curing and to the loss of moisture compared to that of the large specimen, which could lead to a reduction in strength for curing at high temperatures. Moreover, when the curing temperature is high, the sample experiences a substantial loss of moisture; the geopolymer reaction requires the presence of moisture in order to develop good strength [1]. 3.2. Fourier Transform Infrared Spectroscopy (FTIR) Fig. 2 shows the IR spectra of the geopolymer cured at different temperatures. The FTIR spectra of band Si-O-Si of geopolymer was cured at room temperature (958 cm-1), 50C (964 cm-1), 60C (972cm-1), 70C (970 cm-1), and 80C (962 cm-1). All these values signify that the vitreous component of the fly ash

A.M. Mustafa Al Bakri et al. / Physics Procedia 22 (2011) 286 291

289

reacts with the alkaline activator to produce a new product: alkaline aluminosilicate gel [22]. Comparing the shift of the band, samples cured at 60C produced the maximum band (972cm-1), indicating that the reaction product contained more Si than others contributing to the maximum compressive strength [2425]. The broad band appearing in all IR spectra 3500cm-1 and 1600cm-1 responded to stretching (-OH) and bending (H-O-H) vibrations of bound water molecules, which are surface absorbed or entrapped in the large cavities of the polymeric framework [19, 22, 26]. The higher values of these bands indicate a higher degree of water molecule absorption in their mass. p

Fig. 1: The compressive strength of different curing temperatures

*RT = room temperature Fig. 2: FTIR spectra of fly ash and geopolymer with different curing temperatures

3.3. Scanning Electron Microscopy (SEM) The microstructure of fly ash and geopolymer paste for different curing temperatures was observed using SEM. Fig. 3 (a) shows the microstructure of the original fly ash before being activated with the

290

A.M. Mustafa Al Bakri et al. / Physics Procedia 22 (2011) 286 291

alkaline activator. As seen in the figure, the fly ash consists of spherical particles of different sizes, where some particles may contain other particles of a smaller sizes in their interior [27]. The surface texture of fly ash appears to be smooth and dense to highly porous [28]. The surface of the fly ash regularly includes the existence of some quartz particles or some vitreous unshaped fragments [22]. Fig. 3 (b) (f) shows the microstructure morphology of geopolymers at different curing temperatures (room temperature, 50C, 60C, 70C, 80C). All samples that had been cured showed that the materials are heterogeneous; partially reacted and unreacted fly ash existed on the dense gel-like matrix geopolymer. The considerable amount of unreacted or not totally consumed spheres of fly ash indicated a moderate degree of reaction in the system. Fig. 3 (d) showed less unreacted fly ash and a denser matrix compared to others, which leads to higher compressive strength. Meanwhile, the microstructure of other samples are less dense as the existence of microcracks and pores contributed to lower compressive strength.

Fig. 3(a-f): Microstructure of fly ash and geopolymer with different curing temperatures

4. Conclusion Based on the data from this study, it can be concluded that the optimum curing temperature for fly ash-based geopolymer is 60C, which produced the maximum compressive strength (67.04 MPa). When the geopolymer cured at high temperature, the samples do not had enough moisture in order to develop better strength. Moreover cured at high temperature also cause the samples to crack. This result was supported by FTIR analysis and microstructure analysis using SEM.

A.M. Mustafa Al Bakri et al. / Physics Procedia 22 (2011) 286 291

291

Acknowledgement This study was funded by King Abdul Aziz City Science and Technology (KACST). We would like to extend our appreciation to the Green Concrete @UniMAP and the School of Material Engineering, Universiti Malaysia Perlis (UniMAP). References
[1] Chindaprasirt P, Chareerat T, and Sirivivatnanon VJ. Cem and Conc Composites, Vol. 29; 2007, p. 224-229. [2] Malhotra VM. ACI Concrete International, Vol. 24 (7) ; 2002. [3] Davidovits J, Davidovics M. Ceramic Engineering Science Proceeding, Vol. 9; 1988, p. 842-853. [4] Palomo A, Macias A, Blanco MT, Puertas F. In: Proceedings of the 9th International Congress on the Chemistry of Cement; 1992, p. 505-511. [5] Van Jaarsveld JGS, van Deventer JSJ, Lukey GC, Chemical Engineering Journal, Vol. 89; 2002, p. 63-73. [6] Kovalchuk G, Fernandez-Jimenez A, Palomo A, Fuel, Vol. 86; 2007, p. 315-322. [7] Krivenko PV, Kovalchuk GYu. Dundee ; 2002, p. 123-132. [8] Criado M, Palomo A, Fernandez-Jimenez A. Fuel, Vol. 84; 2005, p. 2048-2054. [9] Palomo A, Grutzeck MW, Blanco MT. Cement and Concrete Research, Vol. 29(8); 1999, p. 1323-1329. [10] Mustafa Al Bakri AM, Kamarudin H, Khairul Nizar I, Bnhussain M, Zarina Y, Rafiza AR. Advanced Materials Research, Vols. 341-342; 2012, p. 189-193. [11] Hardjito D, Rangan BV. Research Report GC 1, Faculty of Engineering, Curtin University of Technology, Perth, Australia ; 2005. [12] Hardjito D, Wallah SE, Sumajouw DMJ, Rangan BV, Australian Journal of Structural Engineering, No. 6; 2005, p. 1-9. [13] Matthew R, Brian OC. Journal of Material Chem., No. 13; 2003, p. 1161-1165. [14] Temuujin J, Williams RP, Van Riessen A. Journal of Materials Processing Technology, Vol. 29; 2009, p. 5276-5280. [15] Hardjito D, Chung Cheak C, Ho Lee Ing C. Modern Applied Science, Vol. 2, No. 4, July; 2008. [16] Mustafa Al Bakri AM, Kamarudin H, Bnhussain M, Khairul Nizar I, Rafiza AR, Zarina Y. Journal of Engineering and Technology Research, Vol. 3(2); 2011) , p. 44-49. [17] Hardjito D, Wallah SE, Sumajouw DMJ, Rangan BV. ACI Materials Journal, Vol. 101, No. 6, Nov-Dec; 2004. [18] Rangan BV, In: Low-Calcium, Fly-Ash-Based Geopolymer Concrete, Concrete Construction Engineering Handbook, Taylor and Francais Group, LLC; 2008 , p. 1-9. [19] Swanepoel JC, Strydom CA. Applied Geochemistry, Vol. 17(8); 2002, p. 1143-1148. [20] Panias D, Giannopoulou IP, Perraki T. Colloids and Surfaces A: Physicochemical. Eng. Aspects, Vol. 301; 2007, p. 246254. [21] Steenbruggen G, Hollman GG. Journal Geochemical Exploration, Vol. 62; 1998, p. 305-309. [22] Fernandez-Jimenez A, Palomo A. Cement and Concrete Research, Vol. 35; 2005, p. 1984-1992. [23] Bakharev T. Cement and Concrete Research, Vol. 35 ; 2005, p. 1224-1232. [24] Fernandez-Jimenez A, Palomo A. Microporous and Mesoporous Mat (in press) ; 2005. [25] Mozgawa W, Sitarz M, Rokita M. Journal of Molecular Structure, Vols. 511-512; 1999, p. 251-257. [26] Panias D, Giannopoulou IP, Perraki T. Colloids and Surfaces A: Physicochemical, Eng. Aspects, Vol. 301; 2007, p. 246254. [27] Fernandez-Jimenez A, Palomo A, Criado M. Cement and Concrete Research, Vol. 35; 2005, p. 1204-1209. [28] Fernandez-Jimenez A, Palomo A. Fuel, Vol. 82 ;2003, p. 2259-2265.

You might also like