Sie sind auf Seite 1von 15

Construction and Building Materials 41 (2013) 751765

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


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

Review

Alkali-activated metakaolin: A short guide for civil Engineer An overview


Alaa M. Rashad
Building Materials Research and Quality Control Institute, Housing & Building National Research Center, HBRC, Cairo, Egypt

h i g h l i g h t s
" AAMK system has better acid,

g r a p h i c a l a b s t r a c t

" "

"

"

seawater attack, sodium sulfate resistance than PC. AAMK system has very good heat resistant up to 12001400 C. Polypropylene and short bers increase AAMK exural, strength and impact energy. Replacing 10% MK with FA or lime or 20% with calcite gave higher strength. MK and blended MK with steel slag can be used as repair materials.

Flexural strength of MK-based geopolymer at different NaOH concentrations [32]

Effect of slag replacement percentages on compressive strength [90]

a r t i c l e

i n f o

a b s t r a c t
The development of new binders, as an alternative to Portland cement (PC), by alkaline activation, is a current researchers interest. Alkali-activated metakaolin (AAMK), belongs to prospective materials in the eld of Civil Engineering. This paper presents a comprehensive overview of the previous works carried out on the use of MK in alkali activation. 2013 Elsevier Ltd. All rights reserved.

Article history: Received 10 October 2012 Received in revised form 24 November 2012 Accepted 19 December 2012 Available online 4 February 2013 Keywords: Metakaolin alkali activation Durability Blended alkali-activated metakaolin

Contents 1. 2. 3. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hydration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nature, concentration of activator, SiO2/Al2O3 ratio and solid/liquid ratio. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Nature and concentration of activator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. SiO2/Al2O3 ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3. Solid/liquid ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Effect of curing condition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Effect of ultrasound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Effect of MK fineness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Resistance of aggressive solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 752 752 753 753 754 754 755 755 755 755

4. 5. 6. 7.

Mobile: +20 1228527302; fax: +20 233351564, 233367179.


E-mail address: alaarashad@yahoo.com 0950-0618/$ - see front matter 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.conbuildmat.2012.12.030

752

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

8. 9. 10. 11. 12. 13. 14. 15.

Resistance of elevated temperature and fire . . Fibers effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . MK blended with FA . . . . . . . . . . . . . . . . . . . . . MK blended with slag . . . . . . . . . . . . . . . . . . . . MK blended with calcium hydroxide . . . . . . . . MK blended with other materials . . . . . . . . . . . Special applications of MK based geopolymer . Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

756 757 758 758 760 760 762 763 763

1. Introduction Concrete is one of most extensively used construction materials in the world. Each year, the concrete industry produces approximately 12 billion tonnes of concrete and uses about 1.6 billion tonnes of PC worldwide [1]. Indeed, with the manufacture of 1 tonne of cement approximately 0.94 tonnes of CO2 are launched into the atmosphere [2]. The cement industry accounts for 58% of worldwide CO2 emission [3]. Not only CO2 releases from cement manufacture but also SO3 and NOx which can cause the greenhouse effect and acid rain [4,5]. These cause serious environmental impact. To reduce the environmental impact of cement industries, MK and other cementitious materials are used to replace part of cement or as a source of new cementless materials. MK reacts chemically with hydrating cement to form modied paste microstructure. In addition, to its positive environmental impact, MK improves concrete mechanical properties and durability. The term of MK pozzolan refers to a silecious material which, in the presence of water, will react chemically with calcium hydroxide to form cementitious compounds. On the same line, to reduce the environmental impact resulting from cement industries, AAMK (future cement) is recently used. Investigations in the eld of alkali activation had an exponential increased after the research results of the French author Davidovits [6] that developed and patented binders obtained from alkali activation of MK. Alkali activation of MK and other materials can be classied this new kind of binders as the third generation cement after lime and PC. The alkali activation of MK is a way to reduce carbon dioxide lunched into the atmosphere. The alkali activation of MK yields strong [710] and durable cementitious materials that harden at temperatures under 100 C [7,9,10]. The composition, structure and properties of the reaction product obtained in alkali activation of MK are directly impacted by the specic surface and composition of the initial kaolin and the activator type and its concentration. However, in this investigation, the author conducted a literature review focused on AAMK. Hydration, natural of activators, curing condition, resistance of aggressive solutions and re resistance of AAMK were reviewed. In addition, a review on blended MK with other materials, in alkali activation system was presented. 2. Hydration The dissolutionreprecipitation reactions occurring during the leaching of MK with alkaline hydroxide or silicate solutions were relatively well described in terms of the conditions under which certain zeolitic products will form [11,12]. Granizo and Blanco [13] studied the reaction of MK with NaOH solutions. They reported that the alkaline activation of MK to yield a cementitious material was an exothermic process involving three steps: an initial and very fast process of dissolution, followed by an induction period in which the heat exchanged rate decreased and nally an exothermic step of reaction in which cementitious materials precipitated and after which the heat exchanged rate decreased. Based

on calorimetric results, they also reported that the induction period was lengthened as the NaOH solution concentration and liquid percentage increased. The induction period was shortened as the temperature increased. The total heat increased as the liquid percentage and the NaOH concentration increased. Granizo et al. [14] reported that the reaction product of MK activation with sodium silicate + NaOH solutions is an amorphous hydrated sodium aluminosilicate. Some authors [9,15] have found the product of MK activation with NaOH solutions is NASH gel with good mechanical properties. Granizo et al. [16] studied different concentrations of NaOH (from 12 to 18 M) to activate two different types of MK with variable solution/solid ratios. They concluded that the material obtained after MK alkaline activation is mainly an amorphous sodium aluminosilicate. The total heat released increased as Na2O concentration increased as well as the neness of MK increased. The insoluble residues (IR) of the samples made from coarse MK are higher than those of ne MK. IR decreased exponentially as Na2O in solution increased. So, reaction degree increased as NaOH concentration increased. Davidovits [17] described the alkali activation of MK using a polymerization model similar to that proposed to describe the formation of zeolites [18] or zeolite precursors from alkali alumunosilicate solutions. Madani et al. [12] reported that the MK activation involved a dissolution step followed by a step of polycondensation that could be assigned to those described for zeolites which form when kaolinites or metakaolinites are attacked by NaOH solution. Rahier et al. [19] reported that when the activator is a NaOH and waterglass mixture, the material formed is amorphous and cementitious, but its structure and composition are different from the product formed if NaOH is used alone. The amorphous NASH gel had thus similar chemical composition as natural zeolitic materials but without the extensive crystalline zeolitic structure [20]. Zhang et al. [21] studied the effect of the NaOH content and the presence of sodium silicate activators on the formation of crystalline phases from MK-based geopolymers. They reported that geopolymers activated with NaOH alone with Si/Na ratios of 4/4 or less formed the crystalline zeolite NaA (Na96Al96Si96O384.216H2O), but at ratios > 4/4 nanosized crystals of another zeolite (Na6[A1SiO4]6.4H2O) were formed. The Si/Na ratio of 4/4 produced a product of greatest crystallinity. The addition of sodium silicate in addition to NaOH signicantly reduced crystallite formation. However, Li et al. [22] reported that it could be condently stated regarding the nanostructure of NASH gel is: the NASH gel structure was that of a charge-balance aluminosilicate, which was inuenced by the Si/Al ratio and the alkali cations presented. In the structure, Al trended to be surrounded by four Si neighbors in a 4-coodinated geopolymer framework. The charge-balanced alkali metal would not associate the Al atom, but rather would associate with one or more negatively-charged oxygen atoms surrounding the aluminum. As in the case for CSH gel, NASH gels were difcult to characterize with XRD due to their amorphous or nanocrystalline nature. Valuable information about gel nanostructure and composition could be nished using techniques of FTIR, SEM or TEM [23].

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

753

Yunsheng et al. [24] studied the hydration process of MK geopolymer activated with potassium silicate solution, KOH. They reported that at the early stage of hydration the MK particles pack loosely together resulting in the existence of many large voids. As hydration progressed, sponge like gel products were gradually produced and precipitated on the surface of these particles and extended outwards. As a result, these voids were fully lled. At the later stage, the MK particles were covered by a thick gel layer, and the microstructure of geopolymeric paste became very compacted. Sum et al. [25] and Zhang et al. [26] studied MK geopolymer activated with potassium solution. They concluded that at early stage, the MK particles pack loosely together resulting in the existence of many large voids. As hydration preceded, gel products gradually precipitated on the surface of these particles and extended outsides; as a result, these voids were fully lled. At later stage, the MK particles were wrapped by a thick gel layer, and the microstructure of geopolymer paste became very denser. Alonso and Palomo [27] believed that when MK activation was carried out with highly concentration of alkaline solutions (NaOH) in the presence of calcium hydroxide, the main reaction product was a sodium aluminosilicate similar to that obtained when MK was activated in the absence of calcium hydroxide and the formation of CSH gel as product was observed as a secondary reaction product. This system was inuenced by parameters such as curing temperature, alkali concentration, initial solids content, etc. When the activator concentration increased, a delay in polymer formation was produced, whereas temperature accelerated its formation. The solids ratio did not inuence the rate of aluminosilicate formation. In another study, Alonso and Palomo [28] did the calorimetric study of alkaline activation of calcium hydroxideMK solid mixtures. A series of MK and calcium hydroxide mixtures were activated in 1:1 proportion, with different NaOH concentrations (5, 10, 12, 15 and 18 M) at 45 C during 24 h. The activation steps and reaction products were examined through isothermal conduction calorimetry and the reaction products characterizations were carried out by means of chemical analysis and instrumental techniques (XRD, FTIR and nuclear magnetic resonance NMR). Based on the investigation, they concluded that: (a) when NaOH concentration was 5 M or lower in the MK activation in calcium hydroxide presence, the main product formed was CSH gel; (b) when the activation was 10 M or higher the main reaction product was the alkaline aluminosilicate with polymeric character similar to the one obtained in the MK activation with the same activator but in calcium hydroxide absence. As secondary reaction product (between 20% and 30%), CSH gel was obtained; (c) a threshold OH concentration existed above which, the alkaline polymer was mainly formed, and beneath which, the CSH gel was the main reaction product; (d) a high hydroxyl group concentration impeded the calcium hydroxide dissolution, Ca2+ concentration was small and the dissolved silicates were xed as sodium aluminosilicate, when OH was not so high, Ca2+ in solution increased and it precipitated as hydrated calcium silicate.

Granizo et al. [15] investigated MK and mixtures of (MK + calcium hydroxide) with a 1:1 ratio activated with 5 and 12 M NaOH solutions cured at 45 C. They concluded that CSH gel was produced readily with a low NaOH concentration in the presence of calcium hydroxide. The main product formed both with and without calcium hydroxide was the same network structure, with the general approximate formula: Si2Al2Na2H4O10. The rate of alkali material formation was very low in the presence of calcium hydroxide. The reaction rate of alkali material formation was very low in the presence of calcium hydroxide. Table 1 summarizes the hydration product versus activator type. 3. Nature, concentration of activator, SiO2/Al2O3 ratio and solid/ liquid ratio 3.1. Nature and concentration of activator Alkali hydroxides and waterglass or a combination of them have been studied for AAMK cements. Waterglass-activated cements often give much higher strength than alkali hydroxide-activated cements. However, high curing temperature and high concentration of alkalis are also required to achieve high strength from the activation of MK [2931]. Wang et al. [32] studied MK geopolymer activated with NaOH. They reported that mechanical properties of the geopolymers were greatly dependent on the concentration of NaOH solution. Flexural strength (Fig. 1), compressive strength and apparent density of the geopolymer increased along with increasing of NaOH concentration within 414 mol/l. This attributed to the enhanced dissolution of the metakaolinite particulates and hence the speeded condensation of the monomer in the presence of NaOH with higher concentration. Granizo et al. [14] supported the idea that the alkali activation of MK using solution containing sodium silicate and NaOH results in the production of materials exhibiting higher mechanical strength compared to the activation with only NaOH. Moreover, the exural strength increased when the concentration of Na increased. Pinto [33] studied AAMK and found that mechanical strength increased when using a 12 M concentration of NaOH activator and calcium hydroxide percentage from 0% to 20%. However, for a concentration of 15 M it was noticed that the calcium hydroxide percentage did not inuence strength. Also the alkali activation of MK was studied and reported that the used of an alkaline activator with waterglass caused an increase in mechanical strength, from 30 to 60 MPa in compression and from 5 to 7 MPa in exural strength. Duxson et al. [34] reported that the concentration of soluble silicon affected the distribution of porosity in MK-based geopolymers, i.e. low concentrations result in the formation of dense gel, while high concentrations result in reduced gel skeletal densities. They also reported that the compressive strength of MK based geopolymers increased linearly by approximately 400% from Si/ Al = 1.15 to Si/Al = 1.9, where it obtained its maximum value, before decreasing again at the highest Si/Al ratio of 2.15. Higher

Table 1 Hydration product versus activator type. Author Granizo et al. [14] Granizo et al. [9] Granizo et al. [15] Madani et al. [12] Zhang et al. [21] Alonso and Palomo [27] Alonso and Palomo [28] Granizo et al. [15] Activator Sodium silicate + NaOH NaOH NaOH NaOH NaOH NaOH NaOH NaOH Hydration product Hydrated sodium aluminosilicate NASH gel NASH gel Zeolite Zeolite Sodium aluminosilicate Alkaline aluminosilicate CSH Notes

MK + lime

754

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

Fig. 1. Flexural strength of MK-based geopolymer at different NaOH concentrations [32].

strength was recorded when the ratios of SiO2/Al2O3 and Na2/Al2O3 were 3.03.8 and about 1, respectively.

3.2. SiO2/Al2O3 ratio Pinto [33] studied AAMK based mixtures and found that some mixtures with calcium hydroxide and an atomic ratio of SiO2/ Al2O3 = 5.1 led to higher compressive strength performance. Other authors [35] working with high SiO2/Al2O3 molar ratio MK. They found out that hydration products developed during geopolymerization have lower SiO2/Al2O3 molar ratio than in the original precursor material. De Silva et al. [36] studied the early-stage reaction kinetics of MK/sodium silicate/NaOH geopolymer system. The setting time and early strength development characteristics of mixtures containing varying SiO2/Al2O3 ratios, cures at 40 C for up to 72 h were studied. They concluded that (a) increasing SiO3/ Al2O3 ratio led to longer setting times; (b) increasing SiO3/Al2O3 molar ratios up to 3.43.8 was responsible for the high-strength gained observed at later stages; (c) increasing in Al led to lower strengths with increasing NaAlSi grained rather than amorphous NaAlSi containing geopolymers. Rowles and OConnor [37] studied the alkali-activation of MK, noticing that the mechanical strength was higher for a molar ratio Si/Al/Na of 2.5:1:1.3, Provis and van Deventer [38] used potassium, sodium and mixed (1:1) sodium/potassium silicate solutions that prepared by dissolving amorphous silica in KOH solution with H2O/M2O = 11 (M = K and/ or Na), giving solutions with composition SiO2/M2O = 0.02.0. These solutions were then mixed with MK. They investigated the effects of sample SiO2/Al2O3 ratio, Na/(Na + K) ratio and reaction temperature. The results obtained indicated that the initial gel phase formed during geopolymerisation was later transformed to a second, more-ordered gel phase, and provided detailed information of the rst gel phase during the rst 3 h of reaction. Increasing the SiO2/Al2O3 ratio generally decreased the initial rate reaction. Zhang et al. [39] studied the effect of SiO2/Al2O3, K2O/Al2O3 and H2O/K2O on the compressive strength of K-PSS geopolymer. They prepared K-PSS geopolymer with different SiO2/Al2O3, K2O/Al2O3 and H2O/K2O ratios by activation the mixture of MK, silica fume (SF) and NaAlO2 with potassium silicate (KOH) solution. The specimens were cured at temperature of 20 C and 95% RH. They concluded that the SiO2/Al2O3 had a signicant impact on the compressive strength. The highest compressive strength was obtained when SiO2/Al2O3 = 4.5, K2O/Al2O3 = 0.8 and H2O/K2O = 5.0. Yunsheng et al. [40] studied the inuence of three key parameters of SiO2/Al2O3, M2O/Al2O3 and H2O/M2O on the synthesis of MKbased geopolymer cement. MK was activated with NaOH and sodium silicate solution with the mole ratio of SiO2/Na2O of 3.2

and a solid content of 37%. The silicon content was increased by the addition of SF with 95% or higher SiO2 content to compensate for the shortage of silicon in MK. The mole ratios were varied: 5.5 6 SiO2/Al2O3 6 6.5, 0.8 6 Na2O/Al2O3 6 1.2, 7.0 6 H2O/Na2O 6 10. The levels for each of the factors were set at three grades (low, intermediate and high). They concluded that Na2O/Al2O3 and H2O/Na2O had signicant impact on the compressive strength. The highest compressive strength (34.9 MPa) was achieved at SiO2/ Al2O3 = 5.5, Na2O/Al2O3 = 1.0 and H2O/Na2O = 7.0. Temuujin et al. [41] studied the effect of Si:Al variation on the adhesion strength of MK geopolymers to stainless and mild steels. Sodium silicate was employed to activate MK. They used different Si:Al ratios of 1. 2 and 2.5 with water/binder (w/b) ratios of 0.61, 0.45 and 0.74, respectively, at xed Na:Al ratio of 1. They reported that adhesion strength > 3.5 MPa to stainless and mild steels for geopolymer with Si:Al = 2.5 and Na:Al = 1 while for compositions with Si:Al = 1 and 2 the adhesion to the metal substrates was very weak. Aquino et al. [42] studied some of mechanical properties of geopolymers synthesized by alkali (NaOH or KOH) activation MK and SiO2 mixture. Samples with K/Al or Na/Al atomic ratios equal to 1, Si/Al atomic ratios in the 1.252.5 range and H2O/Al2O3 molar ratios of 11 or 12 were cured at 80 C for 24 and 48 h. The results indicated that the density of the geopolymers increased with increasing Si/Al ratios for NaOH and KOH activators. Increasing density of the geopolymers with increasing Si/Al ratios had significant effect on increasing Youngs modulus, Vickers hardness, fracture toughness and strengths only at lower Si/Al ratios (below Si/ Al = 1.52). At higher Si/Al ratios, all mechanical properties decreased regardless of increasing density of the geopolymers. The samples cured for 48 h gave higher strength than those cured for 24 h. 3.3. Solid/liquid ratio Strength decreases as the water/solid ratio increases. This trend is analogous to water/cement ratio in the compressive strength in PC system. Mostowicz and Berak [43] mentioned the tendency of zeolitic synthesis mixtures to form larger crystals when the total amount of water in the reaction mixture is increased. Yao et al. [44] reported that high solid/liquid (S/L) ratios resulted in low viscosity of slurry and the lower S/L ratios increased the geopolymerization period. On the other hand, Zuhua et al. [45] stated that low S/L ratios could accelerate the dissolution of source materials. Zhang et al. [46] activated MK/granulated ground blast-furnace slag (donated as slag)-based geopolymers with alkaline activator. Different L/S ratios were employed. Permeability was measured

Fig. 2. Effect of L/S ratio on permeability of air curing geopolymer [46].

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

755

Fig. 3. Compressive strength of MK-based geopolymers with various S/L ratios [49].

by using Darcy method. They reported that the permeability coefcient (k) increased along with increasing L/S ratio (Fig. 2). The more connective pores exist in the geopolymer matrix prepared at higher L/S ratio. When the L/S ratio was 0.6, k was 1.0 106 lm2, almost twice that of 0.55. Kong et al. [47] reported that S/L ratio of 0.8 gave nearly optimum strength and provided good workability. Higher S/L ratio than 0.8 had very low workability and deteriorated the properties of the paste produced. Liew et al. [48] activated MK pastes with alkali activation solution at S/L ratios, by mass, ranging between 0.40 and 1.20. The alkali activation solution was Na2SiO3/NaOH with different ratios. The results of bulk density and compressive strength showed that the S/L of 0.8 gave the highest values at Na2SiO3/NaOH ratio of 0.20. Lin et al. [49] activated MK with alkaline activator. The alkaline activator was a NaOH solution and sodium silicate. The effects of S/L ratios ranging from 0.4 to 1.0 (donated as: SL04, SL06, SL08 and SL10 for S/L ratios of 0.4, 0.6, 0.8 and 1.0, respectively) and SiO2/Na2O ratios ranging from 0.8 to 2.0 on the compressive strengths at ages of 1, 7, 28 and 60 days were studied. The results indicated that after 1 day of curing, a S/L ratio of 0.4 yielded the lowest strength. After 60 days, the strength increased to only 27.7 MPa. The specimen with a S/L ratio of 0.8 was the strongest and its strength increased continuously during 60 days of curing (Fig. 3). The silica and alumina reaction in the MK-based geopolymer system caused the increase in strength. As the S/L ratio approached 1.0, the paste stiffened with low workability. Therefore, the SiO2/Na2O ratio was set to 2, and the S/L ratio ivica et al. [50] studied the effects was set to 0.8. Other authors as Z of the combination of low L/S ratio and pressure compaction of the fresh pastes on the properties of hardened MK geopolymer pastes. The AAMK pastes were prepared with activator solution/MK ratio of 0.08 and compacted by pressure of 300 MPa. The prepared specimens were 20 mm-edge cubes. The reference specimens were prepared with L/S of 0.7 using the compaction by hand. NaOH was used as alkali activator. The results indicated that the use of L/S 0.08 and 300 MPa compaction pressure produced very dense near-nano-pore structure with the degree of the homogeneity and the compressive strength overcoming 500 times reference hardened paste.

optimum curing temperature was 60 C which gave the best geopolymerization process. Perera et al. [52] studied the curing at ambient and controlled RH with mild heating (4060 C) of a MK-based geopolymer of molar ratios Si/Al and Na/Al of 2 and 1, respectively. They monitored the effect of curing condition on the open porosity. They concluded that the curing in the absence of rigorous sealing, in an oven in which the RH was held at 30 70% did not offer any advantage over curing at ambient followed by mild heating (4060 C). Rovnank [53] studied the effect of curing temperatures of 10, 20, 40, 60 and 80 C and time on the compressive strength, exural strength and pore distribution of MK-based geopolymer mortars. Alkaline silicate solution with modulus (SiO2/Na2O) of 1.39 was used as activator. This alkaline was prepared by dissolving of solid hydroxide in sodium silicate. The results showed that the treatment of fresh mixture at higher temperature accelerated the strengths development, but the 28 days mechanical properties were deteriorated in comparison with results obtained for mixtures that were treated at room or slightly decreased temperature. The results also showed that there is an increase in pore size and cumulative pore volume with rising temperature. 5. Effect of ultrasound Feng et al. [54] studied the feasibility of using ultrasound to enhance the geopolymerisation of metakaolinite/sand and y ash (FA)/metakaolinite mixtures. They found that the introduction of ultrasonication into the geopolymerisation systems increased the compressive strength of the formed geopolymers and the strength increased with increasing ultrasonication up to certain time. The dissolution of metakaolinite and FA in alkaline solutions was enhanced by ultrasonication, hence releasing more Al and Si into the gel phase for polycondensation. They also found that the ultrasonication improved the distribution of the gel phase in the geopolymeric matrices and strengthened the binding between the particle surfaces and the gel phases. The ultrasonication enhanced the formation of semi-crystalline to crystalline phases in the formed geopolymers. They also concluded that the improved performance of the ultrasonically formed geopolymers in term of compressive strength could be attributed to the accelerated dissolution of the AlSi source materials, the strengthened bonds at the solid particle/gel phase interfaces, the enhanced polycondensation process and the increased semi-crystalline and crystalline phases. 6. Effect of MK neness Weng et al. [55] studied different specic surface areas of MK activated with sodium silicate and NaOH They concluded that higher specic surface area of MK powders were characterized by quicker setting time, higher compressive strength and more homogeneous microstructure. 7. Resistance of aggressive solutions Davidovits [56] studied the acid corrosion resistance of several different cements in 5% H2SO4 and HCl indicated that AAMK cement had the best acid resistance. PC and Portland slag cement were destroyed easily in acidic environments. Calcium aluminate lost 3060% of the mass, while AAMK lost only 58% mass. Other authors [57] investigated the MK when exposed to aggressive solutions. The alkali activation of MK is a way of producing high strength cementitious materials. Prisms of mortar made of sand and AAMK were immersed in deionized water, ASTM seawater, sodium sulfate solution (4.4 wt.%) and sulfuric acid solution (0.001 M). The prisms were removed from the solutions at 7, 28,

4. Effect of curing condition Muiz-Villarreal et al. [51] studied the effect of curing temperatures on the geopolymerization process, physical, mechanical and optical properties of MK-based geopolymer activated with NaOH and sodium silicate. The inuence of different curing temperatures of 30, 40, 50, 60, 75 and 90 C was studied. They reported that the

756

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

56, 90 180 and 270 days. Their microstructure was characterized and their physical, mechanical and microstructural properties were investigated. It was observed that: (a) the nature of the aggressive solution had little negative effect on the evolution of microstructure and the strength of these materials; (b) 90-day and older specimens experienced a slight increase in their exural strengths with time. This tendency was most pronounced in samples cured in sodium sulfate solutions; and (c) the resultant hydroceramic demonstrated good stability for up to 270 days when submerged in aggressive liquids of various types. 8. Resistance of elevated temperature and re Kuenzel et al. [58] investigated some properties of MK derived geopolymer mortars containing 50%, by weight, silica sand after exposure to elevated temperatures up to 1200 C, for 2 h. The activating solution was mixing sodium silicate, water and NaOH. The specimens were cured at 22 3 C for 77 days. The used sand had three different grades of coarse, medium and ne. The results indicated that the compressive strength, porosity and microstracture of the geopolymer mortar specimens were not signicant affected by temperatures up to 800 C. Nepheline (NaAl/SiO4) and carnegieite (NaAlSiO4) formed at 900 C. After exposure to 1000 C, the mortar specimens were transformed into polycrystalline nepheline/quartz ceramics with relatively high compressive strength. Between 1000 and 1200 C, the specimens softened with gas evolution causing the formation of closed porosity that reduced specimen density and limited mechanical properties. Also, the results indicated that as the grade of sand ner as the higher compressive strength, before or after exposure to elevated temperature (Fig. 4). On the other hand, Kong et al. [59] studied the behaviour of AAMK activated with waterglass and potassium hydroxide after exposure to 800 C at an incremental rate of 4.4 C /min, with re duration 1 h, versus alkali-activated FA. The results indicated that the strength of MK-based geopolymer decreased after exposure to 800 C. On the contrast, the strength of the corresponding FA-based increased after exposed to 800 C. Whilst Davidovits [60], Davidovits et al. [61], Barbosa and Mackenzie [62,63] reported very good heat resistant properties of materials prepared using sodium silicate, potassium silicate and MK, having thermal stability up to 12001400 C. Kong et al. [64] studied the MK geopolymer binder system exposed to elevated temperatures. The combinations of sodium/ potassium silicate and sodium/potassium hydroxide were used as source of activators. They studied different parameters as Si/Al ratio, activator/MK ratio and alkali cation type. The results indicated

that compressive strength of geopolymer that had Si/Al = 1.54, M/ Al = 0.66 higher than those had Si/Al = 1.4, M/Al = 0.42. The strength deterioration due to temperature exposure reduced with increasing Si/Al ratio up to 1.5 then the deterioration began to increase. Potassium-based geopolymer (synthesized with potassium silicate and hydroxide) exhibited less strength regression after elevated temperature exposure compared to equivalent sodiumbased geopolymer systems. Activator/MK around 1.11 seemed to be the optimum that gave higher compressive strength before and after exposure to elevated temperature. Lin et al. [65] studied the effect of addition a-Al2O3 particles to the MK geopolymer activated with potassium silicate solution exposed to 400, 600 C, 800, 1000, 1200 and 1400 C. They reported that no further change until 1400 C. The addition of a-Al2O3 particle ller into MK geopolymers increased the onset crystalline temperature and reduced the crystalline velocity. The thermal shrinkage of MK geopolymers increased with increasing in the heat treatment temperature. The increase in content of a-Al2O3 particles could reduce the thermal shrinkage and maintain a relatively lower density and higher porosity. The increase in content of aAl2O3 particles had no distinct inuence on the exural strength after heat treatment despite an increase in the porosity of the geopolymers. Barbosa and Mackenzie [66] synthesized potassium polysialte (K-PS) and potassium polysialate disiloxo (K-PSDS) geopolymers from metakaolinite and found these kind of geopolymers had excellent thermal stability, especially, the K-PS which showed little sign of melting up to 1400 C. Bernal et al. [67] studied the effect of elevated temperatures of 200, 400, 600, 800 and 1000 C, for 2 h, on geopolymers formulated with an overall SiO2/Al2O3 molar ratio of 3, slag/(slag + MK) ratios of 0.0 and 0.2, constant H2O/ Na2O ratio of 12 and Na2O/SiO2 ratio of 0.25. The results indicated that the geopolymers formulated with MK and slag had higher residual compressive strength than the pure MK-based geopolymer up to 800 C. On the other hand, the pure MK system showed a much higher residual strength upon cooling from 1000 C to room temperature, indicating that the extent of glass formation from the geopolymer gel at 1000 C is reduced by the incorporation of Ca into the gel, as a consequence of formation of CSH type gel that coexisted with the aluminosilicate geopolymer gel. Cheng and Chiu [68] studied re resistance of slag geopolymer blended with metakaolinite. They reported that when a 10 mm thick panel of geopolymer is exposed to 1100 C ame: the measured reverseside temperature reached 240283 C after 35 min. They observed that the re characteristics could be improved by increasing the KOH or the alkali concentration and amount of MK. He et al. [69] studied the effect of elevated temperatures of 1000, 1100, 1200, 1300 and 1400 C, for 90 min in an argon

Fig. 4. Compressive strength of geopolymer/sand mortar specimens as a function of the heat treatment temperature [58].

Fig. 5. Variations of exural strength and work of fracture of the carbon bers reinforced MK geopolymer composite [69].

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

757

atmosphere, on some mechanical properties of MK geopolymer and MK/carbon bers. Composites without and after heat treatment were denoted as C-W, C-1000, C-1100, C-1200, C-1300 and C-1400. Potassium silicate solution was used as activator. The content of carbon bers was 20 vol.% for the composites without heat treatment and 25 vol.% for composites after heat treatment. When the composites were heated in a temperature range from 1100 to 1300 C, it is found that mechanical properties were improved (Fig. 5). The improvement could be attributed to the densied and crystallized matrix, and enhanced bers/matrix bonding based on ne-integrity of carbon bers. On the contrary, for composite heat treated at 1400 C, the strengthening effect of carbon bers was dramatically decompensated. It was resulted strong bers/matrix interface bonding strength. The composite showed substantially decreased in mechanical properties and fractures in a very brittle manner. He et al. [70] prepared unidirectional carbon bers reinforced MK geopolymer composite by ultrasonic-assisted slurry inltration method and heat treated at 1100 C. Then it was impregnated with SolSiO2 to seal the cracks and pores formed during heat treatment. Composites before and after Sol-SiO2 impregnation were denoted as HC and ImHC, respectively. Over an elevated temperatures ranging from 700 to 900 C, the strength of the two composites showed anomalous gained and reached their maximum values at 900 C, 322.1 and 425.1 MPa, respectively, (Fig. 6). These values were 19.8% and 16.8% higher than their ambient ones. At 1100 C, the impregnated composite showed superior high-temperature properties. Bernal et al. [71] studied the mechanical performance of MKbased geopolymer reinforced with refractory aluminosilicate particles and bers after exposed to elevated temperatures of 600, 800 and 1000 C. The aluminosilicate particles were obtained by milling the commercial refractory brick. The aluminasilicazirconia bers were used. The alkali activator was formulated in order to obtain overall matrix SiO2/Al2O3 molar ratios of 3.0 and 3.4 at a constant Na2O/SiO2 ratio of 0.25. The quantity of water was adjusted to achieve an H2O/Na2O ratio of 12. The compressive strength and exural strength results indicated that the inclusion of refractory particles, both with and without additional refractory bers, improved post-exposure compressive and exural strengths compared with sample without reinforcement. The inclusion of higher contents of refractory particles and bers reduced shrinkage of exposed specimens. 9. Fibers effect Lin et al. [72] used short bers (2, 7 and 12 mm) to strengthen MK-based geopolymer. Potassium silicate solution was used as

Fig. 7. Variation of exural strength and work of fracture of geopolymer matrix and bers reinforced MK geopolymer [72].

Fig. 6. Flexural strength versus temperature of HC and InHC [70].

activator. The results of exural strength and work of fracture showed that the composites geopolymer with 7 mm short carbon bers length gave the maximum exural strength and work of fracture values, which were increased by 4.4 times and 118 times, respectively, (Fig. 7). Lin et al. [73] studied the effects of bers content on mechanical properties and fracture behaviour of short carbon bers reinforced MK-based geopolymer matrix. MK powder activated with potassium silicate solution. Different volume fractions of short carbon bers (from 3.5 to 7.5 vol.%) were prepared. Specimens were cured at 80 C for 24 h. At the early stage of curing, a pressure of 0, 0.2, 1.2 or 2.0 MPa was loaded to obtain composites reinforced geopolymers. The results showed that short carbon bers had a great strengthening and toughening effect at low volume contents of bers (3.5 and 4.5 vol.%). With the increase in bers content, the strengthening and toughening effect of short carbon bers reduced. Natali et al. [74] modied some properties of the MK/slag geopolymer with different types of dispersed short bers. Sodium silicate solution, with a SiO2:Na2O ratio of 1.99, and 8 M NaOH solution were used as alkaline activators. The used bers were: HT-carbon bers (average ber diameter 10 lm, tensile strength 5490 MPa), commercial E-glass bers (average ber diameter 10 lm, tensile strength 2500 MPa), PVA bers (average ber diameter 18 lm, tensile strength 1800 MPa) and PVC bers (average ber diameter 400 lm, tensile strength 215 MPa). Regardless the bers type, the same content of bers (1 wt.% fraction on the total mixture) was added to MK/slag mixture. The used bers were cut to obtain a 7 1 mm length. They concluded that all different types of bers had good adhesion properties, micro-cracks propagation along the matrix and created a favourable bridging effect. 1 wt.% of reinforcing bers embedded in the geopolymer matrix was able to increase exural strength from 30% up to 70% depending on bers type. Geopolymers added with PVC and carbon bers exhibited signicantly post-crack improved, resulted more enhancements in ductility after reaching the rst crack load. Li et al. [75] manufactured MK geopolymer composites reinforced with short PVA bers using the extrusion technique. NaOH and sodium silicate solution with a ratio of SiO2 to Na2O of 3.2 were used as reagents. Two types of silica sand (300600 and 90150 lm in diameter) with weight ratio of 3:2 were used as aggregates. The total sand was 32.5%, by weight of the binder. Short PVA bers were used as reinforcements. They concluded that the addition of PVA bers into geopolymer increased extrusion pressures. Extruded geopolymer thin sheets reinforced with microbers showed a good exural strength and reasonable toughness. Yunsheng et al. [76] studied the impact behaviour and

758

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

microstructural characteristics of PVA bers reinforced MK/FA geopolymer boards. They used NaOH and sodium silicate solution with molar ratio of SiO2 to Na2O = 3.2 to activate MK and FA mortars in both cases (separate and blended). Short PVA bers were used as reinforcement materials. The results indicated that the 90% MK blended with 10% FA containing 2% bers, by volume, gave the lowest porosity. The addition of high volume fraction PVA bers changed the impact failure mode of MK/FA geopolymer boards from a brittle pattern to ductile pattern, resulting in a great increase in impact toughness. Zhang et al. [77] studied the effect of polypropylene bers on the mechanical properties and volume stability of geopolymer. Different contents of polypropylene bers were acted in geopolymer matrices. NaOH and water were mixed into sodium silicate to adjust the mole ratio of SiO2 to Na2O of 1.2. The results showed that the 3-day compressive strength, exural strength and impacting energy of geopolymer containing 0.05% bers increased by 67.8, 36.1 and 6.25%, respectively, while shrinkage and modulus of compressibility decreased by 38.6% and 31.3%, respectively. Polypropylene bers offered a bridging effect over the harmful pores, defected and changed the expanding ways of cracks, resulting in a great improvement of strength and toughness. Zhang et al. [46] reported that the large shrinkage problem of the MK with the addition of 10% slag geopolymer, activated with NaOH and sodium silicate, could be solved by appropriate addition of polypropylene bers and MgO as expansion agent as well as careful curing at early age.

10. MK blended with FA Yunsheng et al. [78] studied the durability of alkali-activated blended MK-FA mortars modied with PVA short bers in which the composites manufactured by extrusion technique (SFRGC). The NaOH and sodium silicate were used as source of activation. MK was partially replaced with FA at levels of 0%, 10%, 30% and 50%, by weight. The results indicated that 10% replacement level of MK with FA gave the lower porosity and higher impact strength before and after 20 freezethaw cycles. Zhang et al. [77] modied MK-based geopolymer with FA. The MK was partially replaced with FA at levels of 0%, 33.3%, 50% and 66.7%, by weight. NaOH and water were mixed into sodium silicate to adjust the mole ratio (SiO2 to Na2O) of 1.2. Curing conditions were either in steam at 80 C or in air at 20 C for 1, 3 and 6 days. They concluded that proper addition of FA (33.3%) increased the uidity of fresh paste, prolonged its setting time and improved compressive strength of hardened geopolymer. The compressive strength of the geopolymer containing 33.3%, FA by steam curing for 6 days, was improved by 35.5%. Bankowski et al. [79] activated MK blended with different percentages of FA with solution of sodium silicate and NaOH. The solution had 0.76 M and the sodium silicate concentration was 4.25 M. This geopolymer was used to encapsulate brown coal FA containing high concentrations of heavy metals. The results indicated that leaching of calcium and potassium has been reduced by this geopolymer. Signicant reductions in leaching were found for calcium, arsenic, strontium, selenium and barium. The geopolymer was effective at stabilizing low percentages of FA, but effective as the percentage of FA increased. Phair et al. [80] used sodium silicate and NaOH as activator solutions to activate FA blended with Al sources (such as metakaolinite, kaolinite and K-Feldspar) on the solidication stabilization of heavy metals. They reported that for increasing the efciency of immobilization, it is suggested that the metal waste be pre-treated with the Al source/clay before being added to the geopolymer mixture. This would maximize the sportive capacities of the Al source. They also reported that

all matrices were generally found to be highly efcient in retaining Pb within the matrix with the order of effectiveness: FA > kaolinite > K-feldspar > metakaoline. Xu et al. [81] studied factors affecting the immobilization of heavy metals in FA-based geopolymers. They used a solution of KOH and K2SiO3 to activate the mixture of MK and FA. The results showed that the heavy metals can be effectively immobilized into the geopolymeric matrices. The concentrations of alkali activator and different types of heavy metals had impact on the immobilization behaviour to one metal in the same system. Yunsheng et al. [82] studied the immobilization behaviour of MK/FA (100/0, 90/10, 70/30, 50/50 and 30/70), mortars, activated with NaOH and sodium silicate solution with the molar ratio of, SiO2/Na2O, of 3.2. There were different curing conditions. They concluded that geopolymer containing 70% MK and 30% FA that was synthesized at steam curing (80 C for 8 h), exhibited higher mechanical strength. The compressive and exural strengths were 32.2 and 7.15 MPa, respectively. The synthesized geopolymer can effectively immobilise Cu and Pb heavy metals. Aguilar et al. [83] produced lightweight concretes based on binders composed of MK with 0 and 25% FA activated with 15.2% of Na2O using sodium silicate of modulus SiO2/Na2O = 1.2. Concretes with densities of 1200, 900 and 600 kg/m3 were obtained by aeration by adding aluminum powder, in some formulations lightweight aggregate of blast furnace slag was added at a ratio binder: aggregate 1:1; curing was carried out at 20 and 75 C. The compressive strength development was monitored for 180 days. They concluded that it is possible to produce concrete based on geopolymers of MK of different densities. The substitution of MK with 25% FA was viable to form reactive cementitious pastes. The increment in the curing temperature from 20 to 75 C accelerated the development to the compression strength during the rst day; in the long term curing at 20 C results in similar results.

11. MK blended with slag Buchwald et al. [84] studied the activation of MK blended with slag using NaOH as alkali activator. The activator concentration for the pure MK was higher than that for the pure slag, in order to reach the same Na/Al value of 0.4. The water content was adjusted to give the same workability. They concluded that both type of reaction products CSH from slag and aluminosilicate network from MK were able to coexist. The alkaline activation of MKslag mixtures yields a binder blend containing CSH system and geopolymer system and a mixture of both phases due to interaction at their contact surface. In another investigation, Buchwald et al. [85] studied blended slag with MK activated with two different concentrations of NaOH solution. One with low NaOH concentration (916 wt.%) and the other with high NaOH concentration of 25 wt.%. They measured the reaction progressed of the alkaliactivated pastes by isothermal calorimetry and ultrasonic measurements. They reported that the condensation reaction was accelerated by blending slag and MK but the inuence was much apparent at higher concentrations of activator. This explained by a higher amount of dissolution of both slag and MK, but it is more signicant on MK dissolution. Cheng and Chiu [68] studied slag geopolymer blended with metakaolinite. The results indicated that the more metakaolinite added in the system, the slower setting time. The compressive strength increased with increasing in metakaolinite content. The reason could be that the more metakaolinite added, the more Al gel formed in the system, therefore, giving a higher degree of geopolymer reaction. The density results showed a decrease with the increase in metakaolinite content. Shen et al. [86] partially replaced slag with zeolites or MK in alkali activation. The

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

759

replacement slag with zeolites or MK increased the porosity of the hardened pastes, but the leached fraction of Ca+ and Sr2+ were decreased. The decrease in leached fraction may be attributed to the formation and adsorption properties of (Al + Na) substituted CSH and self-generated zeolite precursor. Yip et al. [87] blended slag with MK using MK/(MK + slag) mass ratios of 1, 0.8 and 0.6 in alkali-activated system activated with commercial sodium silicate solution and NaOH pearls at molar ratios of 2, 1.5 and 1.2. The compressive strength results showed that at MK/(MK + slag) = 0.8 gave the higher strength when molar ratios were 2 and 1.2. At molar ratio of 1.5, MK/(MK + slag) = 1 gave the higher strength. Burciaga-Diaz et al. [88] studied the strength development on alkaline activation of MK/slag pastes. The main parameters were MK/ slag weight ratios of 100/0, 80/20, 50/50, 20/80 and 0/100, modulus of the alkaline solutions of sodium silicate (MSiO2/NaO2 at 0, 1, 1.5 and 2), % Na2O (5%, 10% and 15%) and curing time. Cubes were cured at 20 C for 360 days. The results indicated that the composition of 20% slag plus 80% MK at 10% Na2O gave the highest compressive strength. The modulus of 11.5 was sufcient to promote an adequate activation of binders of slag and MK and their binary mixtures. Yunsheng et al. [89] tested the mechanical strength of MK/slag geopolymer mortars. NaOH and sodium silicate solution with the molar ratio (SiO2/Na2O) of 3.2 were used as alkaline reagents. The ratios of MK/slag were 100/0, 90/10, 70/30, 50/50 and 30/70, by weight. The specimens were cured at 20 C and 100% RH for 28 days. The results showed that geopolymer mortar containing 50% slag gave the highest compressive strength followed by 70% slag and followed by 30% slag (Fig. 8). The results of exural strength showed a similar tendency as compressive strength. They used 50/50 mixture cured at 80 C for 8 h to study the immobilization behaviours of MK/slag based geopolymer in presence of Pb and Cu. They concluded that leaching tests showed that MK/slag based geopolymer could effectively immobilized Cu and Pb heavy metal and the immobilization efciency exceeded 98.5% when the amount of heavy metals contained in slag based geopolymeric matrix was in the range of 0.10.3%, by mass of the binder. The Pb showed better immobilization efciency than Cu in the case of large dosages of heavy metals. Bernal et al. [67] studied the compressive strength of alkali silicate-activated blends of MK and slag pastes. The parameters were slag/(slag + MK) ratios and SiO2/Al2O3 molar ratios. Alkaline activating solutions were prepared by blending sodium silicate solution with solid analytical-grade NaOH to reach the desired modulus values. The results indicated that the addition of slag to MK-based geopolymers led to an increase in compressive strength compared with reference specimens produced using MK as the sole precursor. This effect was more

evident in geopolymers formulated with SiO2/Al2O3 ratios of 3.8 and 4.0. Chen et al. [90] studied the compressive strength of alkali-activated MKslag hydroceramics (AMSHC). The mixtures were made from MK and slag with a pure alkali solution of NaOH and a simulated high-alkaline waste (SAW) liquid via of hydrothermal curing for 24 h. The affecting factors including curing temperature, content of slag and dosage of simulated highly-alkaline waste on the properties of AMSHC. The results indicated that the compressive strength of AMSHC matrixes improved effectively by the addition of slag and the elevated curing temperature. Structural evolution in pastes produced from alkali silicate-activated slag/MK blends was assessed by Bernal et al. [91]. They reported that in the initial period of the reaction, the addition of MK led to an increase in the total setting time, reduced the heat release and affected the reaction mechanism by introduction of a large quantity of additional Al. This effect was more signicant when an activating solution with a higher silicate modulus was used, and led to a slight reduction in the nal mechanical strength. However, the alkaline activating solutions were formulated by blending a commercial sodium silicate solution together with 50% NaOH solution to reach modulus (Ms = molar SiO/Na ratio) of 1.6, 2.0 or 2.4. A constant activator concentration of 5% Na2O by mass of (slag + MK) was used. Mortar samples were formulated with a constant water/ (slag + MK + anhydrous activator) ratio of 0.47 and a binder/sand ratio of 1/2.75. Different amounts of MK were blended with slag, the slag/(slag + MK) were 1.0, 0.9 and 0.8. The compressive strengths were measured up to 180 days. The compressive strength results showed that the increase of MK content in the binders led to a reduction of compressive strength, this effect being more signicant when a higher Ms activator was used. The Ms 2.0 specimens were consistently stronger than the Ms 2.4 specimens at all ages and at each level of MK substitution. Bernal et al. [92] studied the effects of activation conditions on some engineering properties (as compressive, exural strength and accelerated carbonation) of alkali-activated slag/MK blends. The concrete mixtures were formulated with SiO2/Al2O molar ratios of 3.6, 4 and 4.4, slag/(slag + MK) ratios of 0.8, 0.9 and 1.0 (i.e. 20%, 10% and 0% MK, respectively) and a constant Na2O/SiO2 molar ratio of 0.25. They used sodium silicate solution with 50 wt.% NaOH solution as alkali activators. They concluded that the higher alkali activator content gave satisfactory early strength development. The inclusion of MK enhanced exural strength at later age. Accelerated carbonation showed rather rapid carbonation accompanied by a loss in strength, this occurred faster at high MK content. Wang et al. [93] studied the compressive strength

Fig. 8. Effect of slag replacement percentages on compressive strength [89].

Fig. 9. Effects of slag content on permeability of air curing geopolymer prepared at L/S ratios of 0.55 and 0.60 [46].

760

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

and porosity of alkali-activated slagFAMK cementitious materials prepared by hydrothermal method. The activator used was waterglass with the modulus adjusted to 1.0 by dissolving NaOH. The ratio of water to solid was about 0.35. The compressive strength and porosity results indicated that this type of material had higher mechanical strength and compact structure. The average compressive strength was more than 60 MPa after hydrothermal process. The porosity was less than 36%. Zhang et al. [46] studied the permeability, measured by Darcy method, of MK/ slag-based geopolymers. Different L/S ratios of 0.55 and 0.6 were employed. They reported that the inclusion of slag could reduce permeability, particularly at L/S ratio of 0.6 (Fig. 9). The existence of slag had only a slight effect on permeability of geopolymer at L/S ratio of 0.55. When the slag content was P10%, geopolymer had a relatively steady and low permeability, suggesting slag had a packing inuence on geopolymer structure [94].

12. MK blended with calcium hydroxide The addition of a sufcient quantity of Ca to AAMK in form of calcium hydroxide could lead to formation of phase separated Alsubstitute CSH and NASH gels [15,94,95]. This is known to be more prevalent at relatively low alkalinity conditions system, because if the OH concentration is high, the dissolution of calcium hydroxide is hindered and it is also possible that very highly alkaline conditions will lead to dissolution of any CSH type phases which are formed. It has also been suggested that Ca2+ is capable of acting as charge-balancing cation within the geopolymeric binding structure, but it needs to be further studied [22]. However, Alonso and Palomo [27] partially replaced MK with calcium hydroxide at levels of either 30% or 50%. The composites were activated with NaOH at different concentrations of 10, 12, 15 and 18 M. The specimens were cured for 24 h at different temperatures of 35, 45 and 60 C. After curing, the specimens were tested in exural. The results showed that the inclusion of 50% calcium hydroxide gave higher exural strength than 30% calcium hydroxide. The strength increased as curing temperature increased and concentration of NaOH decreased (Fig. 10). Buchwald et al. [96] studied the inuence of calcium content on the compressive strength of AAMK. MK powders have been activated with 8 mol/1 NaOH solution. The calcium content of these model mixtures has been increased by successive exchange of the pure alumosilicate powder against calcium hydroxide up to 40% per mass. Compressive strengths were conducted at 28 days and 111 days. The results indicated that the optimal calcium content seemed to be about 10%. Pacheco-Torgal et al. [97] studied the effect of activator concentration, superplasticizer content and calcium hydroxide content on the workability, compressive strength and exural strength of
Fig. 11. Flow versus NaOH concentration and calcium hydroxide content in the present of 3% superplasticizer [97].

AAMK-based mortars. The concentrations of NaOH were 10, 12, 14 and 16 M, while the contents of superplasticizer dosages were 1%, 2% and 3%. MK was partially replaced with calcium hydroxide at levels of 0%, 5% and 10%. The results indicated that the workability of the fresh mortars decreased with increasing NaOH concentration and increased with the content of calcium hydroxide and superplasticizer. The results also showed that the use of 3% of superplasticizer, combined with 10% calcium hydroxide improved mortar ow from less than 50% to over 90% (Fig. 11), while high compressive and exural strengths were maintained. Yip and Deventer [94], Yip et al. [98] studied the aluminosilicates with similar SiO2/Al2O3 and Al2O3/Na2O molar ratios. They found the existence of an optimum of 20% (about 9% of calcium oxide) slag content in alkali-activation of MK-slag mixtures, led to increase compressive strength, which according to them could be explained by the fact that formed CSH within the geopolymeric binders acts as microaggregates resulting in a dense and homogeneous binder.

13. MK blended with other materials Zhang et al. [99] blended MK and slag in the mass ratio of 1:4. Sodium metasilicate, Na2SiO3.9H2O was used as alkaline activator. The MK/slag based geopolymer reinforced by organic resins (OR) which consist of acrylic resin emulsion and polyvinyl acetate resin. The amount of OR ranging from 0% to 15%. They concluded that the excellent compressive and exural performance was attributed to the OR that prevented growing cracks and increased the fracture toughness of the geopolymer composites. The geopolymer composites modied with 1 wt.% OR displayed the highest compressive and exural strengths. Komnitsas et al. [100] used the frerronickel slag-based inorganic polymers that produced by mixing slag with sodium silicate solution (Na2O:SiO2 = 0.3, Na2O = 7.58.5%, SiO2 = 25.528.5%), sodium or potassium hydroxide and water. The effect of additives such as kaolin or MK and pre-curing period on the nal compressive strength was evaluated. The results indicated that the presence of MK decreased the compressive strength. This is probably due to the increased porosity of the new structure as a result of thermal processing. Bignozzi et al. [101] studied the possibility of using electric arc furnace slag (EAFS) blended with MK as starting materials for geopolymers. The MK/EAFS ratios were 40/60, 30/70 and 20/80. Sodium silicate solution with SiO2/ Na2O = 1.99 and NaOH 8 M solution were used to activate the

Fig. 10. Flexural strength of MK/lime specimens cured at 24 h [27].

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

761

Fig. 12. Compressive strength of alkali-activated MK-ladle slag [102].

starting materials. They concluded that the combination of 40/60 was strongly recommended where it gave better mechanical properties and compact microstructure. In another investigation, Bignozzi et al. [102] investigated alkali-activated materials based on partial substation of MK with ladle slag, deriving from the rening process of steel produced by arc electric furnace technology. The replacement levels were 0%, 25%, 50%, 60%, 70%, 80% and 100%, by weight, donated as G-MK100, C-MK75, G-MK50, G-MK40, GMK30, G-MK20 and G-LS100, respectively. Sodium silicate and 8 M NaOH were used as activating solutions. The compressive strength results showed that G-MK100 and G-MK75 gave the lowest values. Mixture of G-MK20 gave the highest compressive strength followed by G-MK40 and followed by G-MK30 (Fig. 12). They also determined the porosity percentages of the studied mixtures. They found that as compressive strength increased as the porosity percentage decreased. Bernal et al. [103] activated MK/slag blended binders with NaOH which mixed with either SF or rice husk ash (RHA), as alternative silica-based activators. The results indicated that pastes with NaOH/SF or NaOH/RHA showed similar trends in mechanical strength development as a function of activation conditions compared with pastes obtained using commercial silicate solutions as activator. All activating solutions promoted higher compressive strength development with increasing slag contents in the binders, which promoted the coexistence of aluminosilicate reaction products along with calcium silicate hydrate gel. Yunsheng et al. [104] used MK blended with SF activated with KOH pellet and potassium silicate solution with molar ratio of SiO2/K2O of 3.25 and solid content of 40%. The experimental results showed that K2O/Al2O3 had a signicant impact on compressive strength. The highest compressive strength when SiO2/Al2O3 = 6.5, K2O/Al2O = 0.8 and H2O/ K2O = 10. Kouamo et al. [105] modied MK geopolymer with aluminaoxide (Al2O3). MK was partially replaced with Al2O3 at levels of

0%, 10%, 20%, 30% and 40%, by weight. The alkaline solution used was a mixture of aqueous of NaOH (12 M) and sodium silicate. The mass ratio of sodium silicate/NaOH of 2.0 was employed. The results showed that its possibly to improve the compressive strength by replacing MK with Al2O3 up to 30%. Beyond 30% replacement level, the compressive strength decreased, in comparison to pure MK. However, 20% Al2O3 plus 80% MK gave the highest compressive strength in which the original compressive strength increased by 18.1%, followed by 10% Al2O3 and followed by 30% Al2O3. Rajamma et al. [106] partially replaced biomass FA with MK, in mortars, at levels of 0%, 20% and 40%, by weight, activated with different molar volumes of NaOH and sodium silicate solutions. The compressive strength results showed that the inclusion of MK improved the compressive strength values. Also, the results showed that as the MK content increased, as the improvement in compressive strength increased (Fig. 13). Lin et al. [49] partially replaced MK with thin-lm transistor liquidcrystal display (TFT-LCD) waste glass, after crushing and milling to approximately 3000 cm2/g, at replacement levels of 0%, 10%, 20%, 30% and 40%, by weight. The alkaline activator was a NaOH solution and sodium silicate. The results showed that both initial and nal setting times increased as the replacement levels of MK increased. The compressive strength decreased as the replacement levels of MK increased (Fig. 14). However, the geopolymer based on 10% waste glass and 90% MK had a compressive strength of 62 MPa after 60 days curing. Yip et al. [87] replaced MK with calcite or dolomite in geopolymers. The replacement levels of calcite were 0%, 20%, 40%, 60%, 80% and 100%. Sodium silicate solution was used as alkaline activator with different molar (SiO2/ Na2O) ratios of 2, 1.5 and 1.2 by addition of solid NaOH to the sodium silicate solution. The compressive strength and shrinkage were measured at ages of 2, 7, 28, 90, 360 and 560 days. The results indicated that the addition of a moderate amount (20% by mass) of calcite was found to have a positive effect on the compressive strength of MK based geopolymeric binder at all ages. More than 20% calcite had a deleterious effect on strength due to signicant disruption of the geopolymer gel network and the reduced reactive aluminosilicate content. The inclusion of calcite did not induce additional shrinkage during the rst 90 days of aging. However, Fig. 15 shows the compressive strength and relative shrinkage results at age of 560 days. The results of this study also showed that partially replacing MK with 20% dolomite proceeded similar trend of partially replacing MK with 20% calcite. Lampris et al. [107] investigated the production of aggregates from silt produced from aggregate washing plants using the

Fig. 13. Compressive strength values of biomass FA-MK mortars [106].

Fig. 14. Development of compressive strength of waste-glass MK-based geopolymers [49].

762

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

Fig. 15. Compressive strength and relative shrinkage at varying calcite content at Ms = 1.2 [87].

process of geopolymerisation. Silts have been blended with either 20% MK or 20% PFA and activated by highly alkaline sodium silicate activating solution. The activating solution was prepared by dissolving NaOH in distilled water and sodium silicate solution was added. The results indicated that the replacement of silts with either 20% MK or 20% PFA increased the compressive strength, but it was higher for MK than it in PFA. Chen et al. [108] used reservoir sludge as a partial replacement of MK in the production of geopolymers. 30% slag plus 70% MK were used as source materials to produce the reference mixture of the geopolymer. MK was replaced with reservoir sludge at levels of 50%, 70% and 100%, by weight. The amounts of NaOH, sodium silicate and distilled water were used to produce the alkaline activators. The results showed that the compressive strengths decreased with increasing replacement level of MK with reservoir sludge. Wang et al. [109] studied the compressive strength, exural strength and wear behaviour (sliding against AISI-1045 steel) of metakaolinite-based geopolymer composites containing 5%, 10%, 15%, 20%, 25% and 30% (volume fraction) polyteta-uoroethylene (PTFE). The compound activator composed of aqueous NaOH and sodium silicate was used. The results showed that mechanical strength of the composites was lower than corresponding geopolymer, while the wear model became mild. The wear was dramatically reduced by 8699.2%. Wang et al. [110] synthesized metakaolinite-based geopolymer composites containing 5%, 10%, 15%, 20%, 25% and 30% (volume fraction) scalelike graphite, polytetra-uoroethylene (PTFE) and molybdenum disulde (MOS2), respectively, in the presence of a compound activator composed of aqueous NaOH and sodium silicate at room temperature. The tribological behaviours of the resulting composites sliding against AISI-1045 steel were investigated on an MM-200 friction and wear tester, and the bending strength and compressive strength of the composites were determined. The results indicated that all the three kinds of the tested solid lubricants reduced both bending and compressive strengths. On the other hand, the friction, wear of the composites and uctuation of the friction coefcient were reduced. Xiangke et al. [111] employed the activated vanadium tailing (AVT) as the main source material for geopolymerisation in combination with MK. The raw vanadium tailing was blended with solid NaOH at a raw vanadium tailing/NaOH mass ratio of 5/1 and heated at 450 C for 1 h. The heated tailing was cooled naturally to room temperature, then ground in a ball mill for 5 min. The AVT was mixed with MK at various mass ratios of 100/0, 90/10, 80/20, 70/30 and 60/40. The compressive strength results showed that the mixture of 70/30 gave the highest compressive strength. Kouamo et al. [112] improved the compressive strength of the volcanic ash geopolymer mortars. Volcanic ash was replaced with MK at levels of 0%, 30%, 40% and 60%, by weight. An alkali fusion

process was used. The results showed an increase in compressive strength with the inclusion of MK content. Also the results showed that as MK content increased, as the compressive strength of the composition increased. The high amount of MK in fused volcanic ash increased the amount of reactive phase content in the aluminosilicate allowing the dissolution of silica and alumina that improved the polycondensation phenomena and the formation of polymeric binder. Consequently, the compressive strength increased. Kani et al. [113] addressed methods to reduce eforescence in a geopolymer binder based on a pumice-type natural pozzolanic material. This eforescence caused by excess sodium oxide remaining un-reacted in the sodium aluminosilicate geopolymers (as natural pozzolan geopolymers). The deposited alkalis can react with atmospheric CO2 resulting in the formation of white carbonate surface deposits known as eforescence. Carbonation usually results in binder degradation, pH reduction and the deposition of carbonate reaction products in the bulk sample, which may or may not be visible to the naked eye, whereas eforescence causes the formation of visible surface deposits and may or may not be accompanied by further of the binder. However, they partially replaced the pumice with MK at levels of 2%, 4%, 6% and 8%, by weight, aiming to reduce the eforescence. They concluded that the additional of MK slightly reduced the eforescence and improved compressive strength.

14. Special applications of MK based geopolymer He et al. [114] studied the initial efforts of explore a new application of MK-based geopolymers to structural health monitoring. A distributed geopolymer-ber optic sensing (G-FOS) system was prepared, where geopolymers were used as smart adhesives to afx optical bers to existing in-service structures to form as integrated G-FOS sensor. Combination of sodium hydroxide and sodium silicate solution was used as activator. Totally 11 different SiO2/Al2O3 ratios ranging from 2.7 to 6 were selected to synthesize geopolymer precursors, while the NaO2/SiO2 and H2O/Na2O ratios of 0.3 and 17.5, respectively, were maintained constant for all synthesized geopolymers, except that the one with a SiO2/Al2O3 ratio = 3.8 (M) had a modied H2O/Na2O ratio of 11.5. The results indicated that the tensile cracking strain of the geopolymers could be controlled by nely tuning the Si/Al ratios or adding appropriate aggregate llers such as sand, thus rendering the smart nature of geopolymers for deformation-based sensing. Geopolymers with SiO2/Al2O3 ratios P3.8 were viable adhesives that could develop strong bond to concrete, steel and glass bers. Vasconcelos et al. [115] carried out an experimental program about the use of MK-based geopolymers mortars for retrotting purposes. They addressed two main situations, the use of geopolymeric mortars as a repairing layer or as a binding agent between CFRP sheets and the repaired concrete. Several compositions of MK geopolymer mortars were executed by varying the percentage of sand/binder mass ratio (30%, 60% and 90%) and concentration of sodium hydroxide (12 M, 14 M and 16 M). The results showed that MK geopolymer gave a high mechanical resistance and a relevant adhesion to the concrete substrate. Also, low adhesion strength between CFRP and geopolymer mortars was obtained. MK geopolymeric mortars with low sand/binder mass ratio presented low adhesion to concrete substrate due to high shrinkage behaviour deduced by the microcracks in the surface of the specimens. On the same line with this, Hu et al. [116] studied repair materials using MK and MK blended with steel slag activated with NaOH and sodium silicate solution with SiO2/Na2O molar ratio and mass concentration of 1.14% and 38%, respectively. The compressive strengths were measured at ages of 8 h, 1, 3, 7 and 28 days.

A.M. Rashad / Construction and Building Materials 41 (2013) 751765

763

Comparing MK with MKsteel slag geopolymers, it was found that the 8 h, 1, 3, 7 and 28 days compressive strengths increased by 43%, 28%, 17%, 6.9% and 7.6%, respectively. The addition of steel slag accelerated the setting time and signicantly improved compressive strength, which could be due to its latent hydraulic cementitious character. Liew et al. [117] studied the possibility of MK to produce cement powder that could be an alternative to PC by applying geopolymerization process. Cement paste was rstly made by alkaline activation of calcined kaolin with alkaline activator (mixture of 610 M NaOH and Na2SiO3 solution), heated at 80 C forming a solidied product, followed by pulverization to xed particle size powder. Different parameters of SiO2/Al2O3, Na2O/SiO2, Na2O/ Al2O3, H2O/Na2O molar ratios, NaOH/Na2SiO3 ratio, concentration of NaOH, MK/activator and heating conditions were studied. Cement powder was added with water and then cured to produce cubes. They found that the key parameter inuenced the properties of the cement powder was NaOH/Na2SiO3 ratio followed by NaOH concentration, MK/activator ratios and heating condition. The results indicated that the highest compressive strength was obtained when the SiO2/Al2O3, Na2O/SiO2, Na2O/Al2O3 and H2O/Na2O molar ratios were 3.10, 0.37, 1.15 and 14.23, respectively.

15. Conclusions The use of MK as source of alkali activation system has been widely investigated in the recent years. The conclusions of the current literature review can be summarized as following: 1. The Alkali-activated binders seem to be the alternative of PC system. The exact reaction mechanism of AAMK is depending on the prime source materials, alkaline activator type and concentration of activator. The main reaction product of MK activation with sodium silicate + NaOH solutions or highly concentrated alkaline solutions in the presence of calcium hydroxide is an amorphous hydrated sodium aluminosilicate. When NaOH with concentration of 5 M or lower in MK activation in calcium hydroxide presence, the main product is NASH gel. 2. Nature of activator plays important roles in AAMK. Waterglass-activated cements often give much higher strength than alkali hydroxide-activated cements. 3. In most cases, as the activator concentration increases, as the mechanical strengths increase, up to certain concentration. 4. The SiO2/Al2O3 ratio has signicant effect on compressive strength geopolymer. The optimal ratios that give higher strength ranging from 3.5 to 5.5, depending on activator type and curing condition. 5. Although some researchers believed that the optimum curing temperature is around 60 C, but this curing temperature depends on many factors as MK neness, activator type and dosage, etc. 6. Although some researchers believed that S/L ratio of 0.8 gives nearly optimum strength and provides good workability, but this ratio affected with SiO2/Na2O ratio. 7. As L/S ratio increases, as permeability increases. 8. Ultrasonication of geopolymers up to certain time, increases the compressive strength and improves the distribution of gel phase in the geopolymeric matrices. 9. Higher specic surface area of MK powders in alkali activation system gives quicker setting time, higher compressive strength and more homogeneous microstructure. 10. AAMK system has better acid, seawater attack, sodium sulfate resistance than PC system.

11. AAMK system has very good heat resistant, where this system has thermal stability up to 12001400 C. When MK blended with slag, the new composite has higher residual compressive strength than the pure MK up to 800 C. On the other hand, the pure MK system shows a much higher residual strength upon cooling from 1000 C to room temperature. 12. Blended MK with refractory particles modied the re resistance of the composition. The inclusion of higher contents of refractory particles and bers reduced shrinkage of exposed specimens. 13. 0.05% polypropylene bers increase the geopolymer compressive strength, exural strength and impacting energy. 1% short bers (FT-carbon, E-glass, PVA) embedded in the geopolymer matrix can increase exural strength from 30% up to 70%, depending on ber type. PVC and carbon bers signicant improve the post-cracks in the geopolymers and enhancement the ductility. 14. Partially replacing 10% MK with FA in alkali activation system gives lower porosity and higher impact strength. Other researchers believed that the inclusion of 33.3% FA in MK based geopolymer gives the highest compressive strength, but depends on the mole ratio and curing condition. 15. 70% MK plus 30% FA based geopolymer exhibits higher mechanical strength and can effectively immobilise Cu and Pb heavy metals. 16. In general, the more MK adds in alkali-activated slag system, the slower setting time. The addition of slag to MK leads to an increase in compressive strength. However, some researchers believed that 80% MK plus 20% slag gave the highest paste compressive strength which depends on molar ratio and Na2O concentration. Other researchers believed that 50% MK plus 50% slag gave the highest mortar compressive strength followed by 30% MK plus 70% slag and followed by 70% MK plus 30% slag at SiO2/Na2O = 3.2 cured at 20 C and 100% RH for 28 days. 17. Partially replacing 10% of MK with calcium hydroxide in geopolymer improves both workability and compressive strength. 18. The addition of 20% calcite or dolomite has positive effect on the strength of MK based geopolymer binder. 19. Blending 60% of electric arc furnace slag with 40% MK in geopolymer gives better mechanical properties and compacts microstructure. Whilst blending 20% MK with 80% ladle slag gives the highest compressive strength and the lowest porosity. 40% MK plus 60% ladle slag comes in the second place. 20. Blending 20% Al2O3 with 80% MK in geopolymer results 18.1% extra compressive strength related to pure MK. 21. Partially replacing volcanic ash with MK in geopolymer mortars increases compressive strength. 22. The inclusion of MK in geopolymer binder based on a pumice-type natural pozzolanic material can slightly reduce the eforescence and improves the compressive strength. 23. MK-based geopolymer mortars can be used as a repairing layer or as a binding agent between CFRP sheets and the repaired concrete. On the same line with this, MK and blended MK with steel slag (MK/steel slag = 80/20) activated with NaOH and sodium silicate solution at SiO2/Na2O = 1.14 can be used as repair materials.

References
[1] Malhotra VM, Metha PK. High performance, high volume y ash concrete. Supplementary cementing materials for sustainable development. 2nd ed. Ottawa: ONT; 2005.

764

A.M. Rashad / Construction and Building Materials 41 (2013) 751765 [33] Pinto AT. Alkali-activated metakaolin based binders. PhD thesis. University of Minho; 2004 [in Portuguese]. [34] Duxson P, Provis JL, Lukey GC, Mallicoat SW, Kriven WM, van Deventer JSJ. Understanding the relationship between geopolymer composition, microstructure and mechanical properties. Colloids Surf A: Physicochem Eng Aspects 2005;269(1-3):4758. [35] Zhang YS, Sun W, Li JZ. Hydration process of interfacial transition in potassium polysialate 9K-PSDS) geopolymer concrete. Mag Concr Res 2005;57:338. [36] De Silva P, Sagoe-Crenstil K, Sirivivatnanon V. Kinetics of geopolymerization: role of Al2O3 and SiO3. Cem Concr Res 2007;37:5128. [37] Rowles M, OConnor B. Chemical optimization of the compressive strength of aluminosolicate geopolymers synthesized by sodium silicate activation of metakaolinite. J Mater Chem 2003;13:11615. [38] Provis John L, van Deventer SJ. Geopolymerisation kinetics. 1. In situ energy dispersive X-ray diffractometry. Chem Eng Sci 2007;62:230917. [39] Yunsheng Zhang, Wei Sun, Zongjin Li. Synthesis and microstructural characterization of fully reacted potassium-poly (sialate-siloxo) geopolymeric cement matrix. ACI Mater J 2008:15664. 105-M18. [40] Yunsheng Zhang, Wei Sun, Zongjin Li. Composition design and microstructural characterization of calcined kaolin-based geopolymer cement. Appl Clay Sci 2010;47:2715. [41] Temuujin Jadambaa, Minjigmaa Amgalan, Rickard William, Lee Melissa, Williams Iestyn, Arie van Riessen. Preparation of metakaolin based geopolymer coatings on metal substrates as thermal barriers. Appl Clay Sci 2009;46:26570. [42] Aquino W, Lange DA, Olek J. The inuence of metakaolin and silica fume on the chemistry of alkalisilica reaction products. Cem Concr Compos 2001;23(6):48593. [43] Mostowicz R, Berak JM. Factors inuencing the crystal morphology of ZSM-5 type zeolites. In: Drzaj B et al., editors. Zeolites. Amsterdam: Elsevier; 1985. p. 6572. [44] Yao Xiao, Zhang Zuhua, Zhu Huajun, Chen Yue. Geopolymerization process of alkali-metakaolinite characterized by isothermal calorimetry. Thermochim Acta 2009;493:4954. [45] Zuhua Z, Xiao Y, Huajun Z, Yue C. Role of water in the synthesis of calcined kaolin-based geopolymer. Appl Clay Sci 2009;43:21823. [46] Zhang Zuhua, Yao Xiao, Zhu Huajun. Potential application of geopolymers as protection coatings for marine concrete. I. Basic properties. Appl Clay Sci 2010;49:16. [47] Kong Daniel LY, Sanjayan Jay G, Kwesi Sagoe-Crentsil. Comparative performance of geopolymers made with metakaolin and y ash after exposure to elevated temperatures. Cem Concr Res 2007;37:15839. [48] Liew YM, Mustafa Al Bakri AM, Bnhussain M, Luqman M, Khairul Nizar I, Ruzaidi CM, et al. Optimization of solids-to-liquid and alkali activator ratios of calcined kaolin geopolymeric powder. Constr Build Mater 2012;37:44051. [49] Lin Kae-Long, Shiu Hau-Shing, Shie Je-Lueng, Cheng Ta-Wui, Hwang ChaoLung. Effect of composition on characteristics of thin lm transistor liquid crystal display (TFT-LCD) waste glass-metakaolin-based geopolymers. Constr Build Mater 2012;36:5017. ivica Vladimr, Balkovic Svetozar, Drabik Milan. Properties of metakaolin [50] Z geopolymer hardened paste prepared by high-pressure compaction. Constr Build Mater 2011;25:220613. [51] Ramlochan T, Thomas M, Gruber KA. The effect of metakaolin on alkalisilica reaction in concrete. Cem Concr Res 2000;30(3):33944. [52] Perera DS, Uchida O, Vance ER, Finnie KS. Inuence of curing schedule on the integrity of geopolymers. J Mater Sci 2007;42:3099106. [53] Rovnank Pavel. Effect of curing temperature on the development of hard structure of metakaolin-based geopolymer. Constr Build Mater 2010;24:117683. [54] Feng D, Tan H, van Deventer JSJ. Ultrasound enhanced geopolymerisation. J Mater Sci 2004;39:57180. [55] Weng Luqian, Kwesi Sagoe-Crentsil, Brown Trevor, Song Shenhua. Effect of aluminates on the formasion of geopolymers. Mater Sci Eng B 2005;117:1638. [56] Davidovits J. Properties of geopolymer cements. In: Krivenko, editor. 1st International conference on alkaline cements and concretes, Kiev, Ukraine, vol. 1; 1994. p. 13149. [57] Palomo A, Blanco-Varela MT, Granizo ML, Puertas F, Vazquez T, Grutzeck MW. Chemical stability of cementitious materials based on metakaolin. Cem Concr Res 1999;29:9971004. [58] Kuenzel C, Grover LM, Vandeperre L, Boccaccini AR, Cheeseman CR. Production of nepheline/quartz ceramics from geopolymer mortars. J Eur Ceram Soc 2013;33:2518. [59] Kong Daniel LY, Danjayan Jay G, Kwesi Sagoe-Crentsil. Comparative performance of geopolymers made with metakaolin and y ash after exposure to elevated temperatures. Cem Concr Res 2007;37:15839. [60] Davidovits J. Synthetic mineral polymer compound of the silicoaluminates family and preparation process. US Patent 4472199; 1984. [61] Davidovits J, Davidovits M, Davidovits N. Process for obtaining a geopolymeric alumino-silicate and prodicts thus obtained. US Patent 5342595; 1994. [62] Barbosa Valeria FF, MacKenzie KJD. Thermal behavior of inorganic reopolymers and composites derived from sodium polysialate. Mater Res Bull 2003;38:31931.

[2] Gartner E. Industrially interesting approaches to low-CO2 cements. Cem Concr Res 2004;34:148998. [3] Scrivener KL, Kirkpatrick RJ. Innovation in use and research on cementitious material. In: 12th International congress of chemistry of cement, Montreal, Canada; 2007. [4] Rashad Alaa M, Zeedan Sayieda R. The effect of activator concentration on the residual strength of alkali-activated y ash pastes subjected to thermal load. Constr Build Mater 2011;25:3098107. [5] Park Sang-Sook, Kang Hwa-Young. Characterization of y ash-pastes synthesized at different activator conditions. Korean J Chem Eng 2008;25(1):7883. [6] Davidovits J., Synthesis of new high temperature geo-polymers for reinforced plastics/composites. In: SPE PACTEC 79 Society of Plastic Engineers, Brookeld Center; 1979. p. 1514. [7] Davidovits J. Minerals polymers and methods of making them. US Patent 4472199; 1982. [8] Davidovits. Geopolymers: inorganic polymeric new materials. J Therm Anal 1991;37:163356. [9] Granizo ML, Blanco-Varela MT, Puertas F, Paloma A, Alkali, activation of metakaolin: inuence of synthesis parameters. In: Proceedings of the tenth international congress on the chemistry of cement, Gthenburg, Sweden, vol. 3; 1997. [10] Krivenko P. Alkaline cements: terminology, classication, aspects of durability. In: Proceedings of the tenth international congress on the chemistry of cement, Gthenburg, Sweden, vol. 4; 1997. [11] Barrer RM, Mainwaring DE. Chemistry of soil minerals. Part XIII. Reactions of metakaolinite with single and mixed bases. J Chem Soc Dalton Trans 1972:253446. [12] Madani A, Aznar A, Sanz J, Serratosa JM. 29Si and 27Al NMR study of zeolite formation from alkali-leached kaolinites-inuence of thermal reactivation. J Phys Chem 1990;94:7605. [13] Granizo ML, Blanco MT. Alkaline activation of metakoalin an isothermal conduction calorimetry study. J Therm Anal 1998;52:95765. [14] Granizo ML, Blanco-Varela MT, Martnez-Ramrez S. Alkali activation of metakaolins: parameters affecting mechanical, structural and microstructural properties. J Mater Sci 2007;42:293443. [15] Granizo M,L, Alonso S, Blanco-Varela MT, Palomo A. Alkaline activation of metakaolin: effect of calcium hydroxide in the products of reaction. J Am Ceram Soc 2002;85(1):22531. [16] Granizo ML, Blanco-Varela MT, Palomo A. Inuence of the starting kaolin on alkali-activated materials based on metakaolin. Study of the reaction parameters by isothermal conduction calorimetry. J Mater Sci 2000;35:630915. [17] Davidovits J. Early high strength mineral polymer. US Patent 45009985; 1985. [18] Duxson P, Fernndez-Jimnez A, Provis JL, Lukey GC, Palomo A, van Deventer JSJ. Geopolymer technology: the current state of the art. J Mater Sci 2007;42:291733. [19] Rahier H, Simons W, van Mele B, Biesemans M. Recent literature in geopolymer science and technology. J Mater Sci 1997;32(9):223747. [20] Barbosa VFF, Mackensie KJD, Thaumaturgo C. Synthesis and characterization of materials based on inorganic polymers of alumina and silica: sodium polysialate polymers. Int J Inorg Mater 2000;2:30917. [21] Zhang Bo, MacKenzie Kenneth JD, Brown Ian WM. Crystalline phase formation in metakaolinite geopolymers activated with NaOH and sodium silicate. J Mater Sci 2009;44:466876. [22] Li Chao, Sun Hengh, Li Longtu. A review: the comparison between alkaliactivated slag (Si + Ca) and metakaolin (Si + Al) cements. Cem Concr Res 2010;40:13419. [23] Provis J, van Deventer J. Geopolymers: structure, processing, properties and industrial applications. UK: Woodhead Publishing Limited; 2009. [24] Yunsheng Zhang, We Sun, Zuquan Jin, Hongfa Yu, Jia Yantao. In situ observing the hydration process of K-PSS geopolymeric cement with environment scanning electron microscopy. Mater Lett 2007;61:15527. [25] Sum Wei, Zhang Yun-sheng, Lin Wei, Liu Zhi-yong. In situ monitoring of the hydration process of K-PS geopolymer cement with ESEM. Cem Concr Res 2004;34:93540. [26] Zhang YS, Sun W, Li ZJ. Hydration process of potassium polysialate (K-PSDS) geopolymer cement. Adv Cem Res 2005;17(1):238. [27] Alonso S, Palomo A. Alkaline activation of metakaolin and calcium hydroxide mixtures: inuence of temperature, activator concentration and solids ratio. Mater Lett 2001;47:5562. [28] Alonso S, Palomo A. Calorimetric study of alkaline activation of calcium hydroxide + metakaolin solid mixtures. Cem Concr Res 2001;31:2530. [29] Krivenko PV. Alkalen cements. In: 9th International congress on the chemistry of cement, New Delhi, India, vol. IV; 1992. p. 4828. [30] Palomo A, Macias A, Blanco MT, Puertas F. Physical chemical and mechanical characterization of geopolymers. In: 9th International congress on the chemistry of cement, New Delhi, India, vol. V; 1992. p. 50511. [31] Popel GN. Synthesis of a mineral-like stone on alkaline aluminosilicate binders to produce the materials with the increased corrosion resistance. In: Krivenko, editor. 2nd International conference on alkaline cements and concretes, Kiev, Ukraine; 1999. p. 20819. [32] Wang Hongling, Li Haihong, Yan Fengyuan. Synthesis and mechanical properties of metakaolinite-based geopolymer. Colloids Surf A: Physicochem Eng Aspects 2005;268:16.

A.M. Rashad / Construction and Building Materials 41 (2013) 751765 [63] Barbosa VFF, MacKenzie KJD. Synthesis and thermal behavior of potassium sialate geopolymers. Mater Lett 2003;57:147782. [64] Kong Daniel LY, Sanjayan Jay G, Kwesi Sagoe-Crentsil. Factors affecting the performance of metakaolin geopolymers exposed to elevated temperatures. J Mater Sci 2008;45:82431. [65] Lin TS, Jia DC, He PG, Wang MR. Thermo-mechanical and microstructural characterization of geopolymers with a-Al2O3 particle ller. Int J Thermophys 2009;30:156877. [66] Barbosa VFF, Mackenzie KJD. Synthesis and thermal behavior of potassium sialate geopolymers. Mater Lett 2003;57:147782. 9 guez Erich D, de Gutierrez Ruby Meji 9 a, Gordillo [67] Bernal Susan A, Rodri Marisol, Provis John L. Mechanical and thermal characterisation of geopolymers based on silicate-activated metakaolin/slag blends. J Mater Sci 2011. doi:10.1007/s10853-011-5490-z. [68] Cheng TW, Chiu JP. Fire-resistant geopolymer produced by granulated blast furnace slag. Minerals Eng 2003;16:20510. [69] He Peigang, Jia Dechang, Lin Tiesong, Wang Meirong, Yu Zhou. Effects of high-temperature heat treatment on the mechanical properties of unidirectional carbon reinforced geopolymer composites. Ceram Int 2010;36:144753. [70] He Peigang, Jia Dechang, Wang Meirong, Yu Zhou. Improvement of hightemperature mechanical properties of heat treated Cf/geopolymer composites by SolSiO2 impregnation. J Eur Ceram Soc 2010;30:305361. [71] Bernal Susan A, Julian Bejarano, Cristhian Garzn, de Gutirrez Ruby Meja, Silvio Delvasto, Rodrguez Erich D. Performance of refractory aluminosilicate particle/ber-reinforced geopolymer composites. Composites Part B 2012;43:191928. [72] Lin Tiesong, Jia Dechang, He Peigang, Wang Meirong, Liang Defu. Effects of ber length on mechanical properties and fracture behavior of short carbon ber reinforced geopolymer matrix composites. Mater Sci Eng A 2008;497:1815. [73] Lin Tiesong, Jia Dechang, Wang Meirong, He Peigang, Liang Defu. Effects of bre content on mechanical properties and fracture behavior of short carbon bre reinforced geopolymer matrix. Build Mater Sci 2009;32(1):7781. [74] Natali A, Manzi S, Bignozzi MC. Novel ber-reinforced composite materials based on sustainable geopolymer matrix. Procedia Eng 2011;21:112431. [75] Li Zongjin, Yunsheng Zhang, Zhou Xiangming. Short ber reinforced geopolymer composites manufactured by extrusion. J Mater Civil Eng, ASCE 2005:62431. [76] Yunsheng Zhang, Wei Sun, Zongjin Li. Impact behavior and microstructural characteristics of PVA bers reinforced y ash-geopolymer boards prepared by extrusion technique. J Mater Sci 2006;41:278794. [77] Zhang Zu-hua, Yao Xiao, Zhu Hua-jun, Hua Su-dong, Chen Yue. Preparation and mechanical properties of polypropylene ber reinforced calcined kaoliny ash based geopolymer. J Cent South Univ Technol 2009;16:004952. [78] Yunsheng Zhang, Wei Sun, Li Zongjin, Zhou Xiangming. Impact properties of geopolymer based extrudates incorporated with y ash and PVA short ber. Constr Build Mater 2008;22:37083. [79] Bankowski P, Zou L, Hodges R. Reduction of metal leaching in brown coal y ash using geopolymers. J Hazard Mater 2004;B114:5967. [80] Phair JW, van Deventer JSJ, Smith JD. Effect of Al source and alkali activation on Pb and Cu immobilization in y-ash based geopolymers. Appl Geochem 2004;19:42334. [81] Xu JZ, Zhou YL, Chang Q, Qu HQ. Study on the factors of affecting the immobilization of heavy metals in y ash-based geopolymers. Mater Lett 2006;60:8202. [82] Yunsheng Zhang, Wei Sun, Wei She, Guowei Sun. Synthesis and heavy metal immobilization behaviors of y ash based gepolymer. J Wuhan Univ Technol Mater Sci Ed 2009;24(5):81925. 9 az O, Escalante Garci 9 a JI. Lightweight [83] Arellano Aguilar R, Burciaga Di concretes of activated metakaolin-y ash binders, with blast furnace slag aggregates. Constr Build Mater 2010;24:116675. [84] Buchwald A, Hilbig H, Kaps Ch. Alkali-activated metakaolin-slag blendsperformance and structure in dependence of their composition. J Mater Sci 2007;42:302432. [85] Buchwald Anja, Tatarin R, Stephan D. Reaction progress of alkaline-activated metakaolin-ground granulated blast furnace slag blends. J Mater Sci 2009;44:560917. [86] Shen X, Yan S, Wu X, Tang M, Yang L. Immobilization of simulated high level wastes into AASC waste form. Cem Concr Res 1994;24(1):1338. [87] Yip Christina K, Provis John L, Lukey Grant C, van Deventer Jannie SJ. Carbonate mineral addition to metakaolin-based geopolymers. Cem Concr Compos 2008;30:97985. 9 az Oswaldo, Escalante-Garca Jose Ivan, Arellano-Aguilar Rat 9 l, [88] Burciaga-Di Gorokhovsky Alexander. Statistical analysis of strength development as a function of various parameters on activated metakaolin/slag cements. J Am Ceram Soc 2010;93(2):5417. [89] Yunsheng Z, Wei S, Qianli C, Lin C. Synthesis and heavy metal immobilization behaviors of slag based geopolymers. J Hazard Mater 2007;143(12):20613. [90] Chen Song, Wu Mengqiang, Zhang Shuren. Mineral pgases and properties of alkali-activated metakaolin-slag hydroceramics for a disposal of simulated highly-alkaline wastes. J Nucl Mater 2010;402:1738.

765

9 a. Evolution [91] Bernal Susan A, Provis John L, Rose Volker, de Gutierrez Ruby Meji of binder structure in sodium silicate-activated slag-metakaolin blends. Cem Concr Compos 2011;33:4654. [92] Bernal Susan A, de Gutirrez Ruby Meja, Provis John L. Engineering and durability properties of concretes based on alkali-activated granulated blast furnace slag/metakaolin blends. Constr Build Mater 2012;33:99108. [93] Wang Jin, Wu Xiu-ling, Wang Jun-xia, Liu Chang-zhen, Lai Yuan-ming, Hong zun-ke. Hydrothermal synthesis and characterization of alkali-activated slagy ash-metakaolin cementitious materials. Micropor Mesopor Mater 2012;155:18691. [94] Yip CK, Lukey GC, Deventer SJs. The coexistence of geopolymeric gel and calcium silicate hydrate gel at the early stage of alkaline activation. Cem Concr Res 2005;35(9):168897. [95] Davidovits J. Chemistry of geopolymric systems, terminology. In: Gopolymre 99. 2nd International conference, Saint-Quentin, France; 1999. p. 939. [96] Buchwald A, Dombrowski K, Weil M. The inuence of calcium content on the performance of geopolymeric binder especially the resistance against acids. In: Davidovits J, editor. Proceedings of the world congress geopolymer, Saint Quentin, France, 28 June1 July, 2005. p. 359. [97] Pacheco-Torgal F, Moura D, Yining Ding, Said Jalali. Composition, strength and workability of alkali-activated metakoalin based mortars. Constr Build Mater 2011;25:373245. [98] Yip CK, Deventer SJS. Microanalysis of calcium silicate hydrate gel formed within a geopolymeric binder. J Mater Sci 2003;38:385160. [99] Zhang Yao Jun, Wang Ya Chao, Xu De Long, Li Sheng. Mechanical performance and hydration mechanism of geopolymer composite reinforced by resin. Mater Sci Eng A 2010;527:657480. [100] Komnitsas Kostas, Zaharaki Dimitra, Perdikatsis Vasillios. Effect of synthesis parameters on the compressive strength of low-calcium ferronickel slag inorganic polymers. J Hazard Mater 2009;161:7608. [101] Bignozzi MC, Barbieri L, Lancellotti I. New geopolymers based on electric arc furnace slag. Adv Sci Technol 2010;69:11722. [102] Bignozzi Maria Chiara, Manzi Stefania, Lancellotti Isabella, Kamseu Elie, Barbieri Luisa, Leonelli Cristina. Mix-design and characterization of alkali activated materials based on metakaolin and ladle slag. Appl Clay Sci, in press, corrected proof. [103] Bernal Susan A, Rodrguez Erich D, de Gutirrez Ruby Mejia, Provis John L, Delvasto Silvio. Activation of metakaolin/slag blended alkaline solutions based on chemically modied silica fume and rice husk ash. Waste Biomass Valor 2011. doi:10.1007/s12649-011-9093-3. [104] Yunsheng Zhang, Wei Sun, Li Zongjin. Preparation and microstructure of KPSDS geopolymeric binder. Colloids Surf A: Physicochem Eng Aspects 2007;302:47382. [105] Kouamo Tchakoute H, Elimbi A, Ngally Sabouang, Njopwouo D. The effect of adding alumina-oxide to metakaolin and volcanic ash on geopolymer products: a comparative study. Constr Build Mater 2012;35:9609. [106] Rajamma Rejini, Labrincha Joo A, Ferreira Victor M. Alkali activation of biomass y ash-metakoalin blends. Fuel 2012;98:26571. [107] Lampris C, Lupo R, Cheeseman CR. Geopolymerisation of silt generated from construction and demolition waste washing plants. Waste Manage 2009;29:36873. [108] Chen Ji-Hsien, Huang Jong-Shin, Chang Yi-Wen. Use of reservoir sludge as a partial replacement of metakaolin in the production of geopolymers. Cem Concr Compos 2011;33:60210. [109] Wang Hongling, Li Haihong, Yan Fengyuan. Reduction in wear of metakaolinite-based geopolymer composite through lling of PTFE. Wear 2005;258:15626. [110] Wang Hongling, Li Haihong, Yan Fengyuan. Synthesis and tribological behavior of metakaolinite-based geopolymer composites. Mater Lett 2005;59:397681. [111] Xiangke Jiao, Yimin Zhang, Tiejun Chen, Shenxu Bao, Tao Liu, Jing Huang. Geopolymerisation of a silica-rich tailing. Minerals Eng 2011;24:17102. [112] Kouamo Techakoute H, Mbey JA, Elimbi A, Diffo Kenne, Njopwouo D. Synthesis of volcanic ash-based geopolymer mortars by fusion method: effect of adding metakaolin to fused volcanic ash. Ceram Int 2013;39:161321. [113] Kani Ebrahim Naja, Ali Allahverdi, Provis John L. Eforescence control in geopolymer binder based on natural pozzolan. Cem Concr Compos 2012;34:2533. [114] He Jian, ASCE Guoping Zhanh PEM, Hou Shuang, ASCE CS Cai M. Geopolymerbased smart adhesives for infrastructure health monitoring: concept and feasibility. J Mater Civil Eng, ASCE 2011;February:1009. [115] Vasconcelos E, Fernandes S, Barroso De Aguiar JL, Pacheco-Torgal F. Concrete retrotting using metakaolin geopolymer mortars and CFRP. Constr Build Mater 2011;25:321321. [116] Hu Shuguang, Wang Hongxi, Zhang Gaozhan, Ding Qingjun. Bonding and abrasion resistance of geopolymeric repair material made with steel slag. Cem Concr Compos 2008;30:23944. [117] Liew YM, Kamarudin H, Mustafa AM, Luqman M, Khairul Nizar I, Ruzaidi CM, et al. Processing and characterization of calcined kaolin cement powder. Constr Build Mater 2012;30:794802.

Das könnte Ihnen auch gefallen