You are on page 1of 27

Construction and Building Materials 47 (2013) 2955

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


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

A comprehensive overview about the inuence of different additives on the properties of alkali-activated slag A guide for Civil Engineer
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
 Water absorption and permeability porous were markedly improved with the addition of steel bres.  Superplasticizer (naphthalene-based) improved workability and compressive.  SF in AAS matrix decreased workability and increased compressive strength.  10% PC in AAS gave higher compressive strength and the lowest drying shrinkage.  Gypsum in AAS reduced drying shrinkage, setting time and increased compressive strength.

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 slag (AAS) binder is obtained by a manufacturing process less energy-intensive than PC and involves lower greenhouse gasses emission. AAS belongs to prospective materials in the eld of Civil Engineering. Researchers have employed bres, chemical admixtures, mineral admixtures and other materials as additives in AAS system aiming to modify some properties of this system. This paper presents a comprehensive overview of the previous works carried out on using different additives in AAS system. 2013 Elsevier Ltd. All rights reserved.

Article history: Received 29 January 2013 Received in revised form 1 April 2013 Accepted 3 April 2013 Available online 28 May 2013 Keywords: Alkali-activated slag Fibres Chemical admixtures Mineral admixtures Other admixtures

1. Introduction Practices in the delivery of build infrastructure, there are specic concerns over atmospheric CO2 concentrations which at 390 ppm reached record breaking levels (U. N. Intergovernmental Panel on Climate Change) [1]. Major CO2 producing sectors, such as power generation, transportation, oil rening and manufacturing of steel and concrete are under pressure to adopt measures that would drastically reduce the global CO2 emission rate by 2030. Within the concrete industry, cement manufacturing is the main culprit [2]. Each year, the concrete industry produces approximately 12 billion tonnes of concrete and uses about 1.6 billion tonnes of PC worldwide [3]. The production of cement is increasing about 3% annually [4]. Indeed, with the manufacture of 1 tonne of cement approximately 0.94 tonnes of CO2 are launched into the atmosphere [5]. The IEA (International Energy Authority) holds the cement industry responsible for emitting between 6% and 7% of
Tel.: +20 1228527302; fax: +20 233351564.
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.2013.04.011

all the CO2 emission into the atmosphere [6]. The projections for the global demand of PC show that in the next 40 years it will have a twofold increase reaching 6 Gt/year [7]. Among the green house gases, CO2 contributes about 65% of global warming. Additionally, cement production and resulting emissions are expected to increase by 100% from the current level by 2020 [8]. On the same line with this, global demand will have increased almost 200% by 2050 from 2010 [7]. Beside the emission of CO2, cement industry launches SO3 and NOx which can cause the greenhouse effect and acid rain [9,10]. This is particularly serious in the current context of climate change caused by CO2 emissions worldwide, causing a rise in sea level and the occurrence of natural disasters and being responsible for future meltdown in the world economy [11]. In many countries around the world cement production consumes huge amounts of energy, in particular, arising from the calcination of raw materials at around 1500 C and the grinding of raw materials, cement clinker and gypsum [12]. The energy demand associated with PC production is about 17001800 MJ/tonne clinker [1315], which is the third largest use of energy, after aluminium and steel manufacturing industries [6,13]. The cement

30

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

production account about 5% of worldwide industrial energy consumption [16,17]. In addition, cement production consumes considerable amounts of virgin materials (limestone and sand), producing each tonne of PC of which about 1.5 tonnes of raw material is needed [14]. Further, concrete made of PC is subject to certain durability problems that are difcult to solve. In the light of these problems, the scientic community has undertaken to seek new processes, technologies and materials to provide the construction industry with alternative binders. One avenue that is expected to signicant reduce cement is use of blended cement [1825]. The second alternative is to recycling in cement industry [2634] which has become indispensable processes in some countries. The third alternative is the use alkali-activation of slag, y ash (FA), burned clay and other aluminosilicate materials. The new binder materials that can replace PC, by alkali activation, can generate about 8090% less carbon dioxide than PC [35]. Comparison to PC concrete, the global warming potential of alkali-activated concrete is 70% lower [36]. In addition, there are numerous advantages of this system as lower heat of hydration [37], the development of earlier and higher mechanical properties [38,39], low heat release [40], better resistance of chemical attack [41,42], freezethaw resistance [43], re resistance [44,45], higher reduction in chloride diffusion [46] and stronger aggregate-matrix interface formation [47,48]. On the contrary, the AAS system presents some problems such as rapid setting periods [49], higher shrinkage values [50], higher formation of salt eforescences [39], higher carbonation [51] and tendency to crack during curing [52,53]. However, these advantages and disadvantages are depended on the types of PC and blended cement, and AAS and not always in the case. For specic problems, adequate solutions have been established. Researchers have been tried to solve some of these problems by using different additives in AAS system. The most commonly used activators are sodium silicate, sodium hydroxide, sodium carbonate or a mixture of sodium - potassium hydroxide (NaOH, KOH) with sodium silicate - potassium silicate, the mixture of sodium hydroxide with sodium silicate has been generally agreed to be the most effective activator and provides the best formulation for high strength and other advantageous properties. However, both the sodium hydroxide and sodium silicate do not exist naturally and they are obtained from energy intensive manufacturing processes. This is particularly the case for sodium silicate, which is produced by melting sand and sodium carbonate at 13501450 C, followed by dissolving in an autoclave at 140160 C under suitable steam pressure [54]. As a result, even though AAS could potentially be considered as a low energy and low carbon cement system. Other uncommon activator as Na2SO4 can be obtained either from natural occurring sodium-sulfatebearing brines, crystalline evaporate deposits or as a by-product during the manufacture of various products such as viscose rayon, hydrochloric acid and silica pigments [13]. However, in general the energy consumption of alkali activation system is calculated to be approximately 60% less than that of PC [55]. Ground granulated blast furnace slag (denoted as slag) is a byproduct obtained in the manufacturing of iron in a blast furnace where iron ore, limestone and coke are heated between 1400 C and 1600 C. Quenching process of molten slag by water is converting it into a ne (no larger than 5 mm diameter), granulated slag of whitish colour. The rapid cooling prevents the formation of larger crystals. The resulting granular material comprises some 8595% non-crystalline calcium-aluminosilicates (glassy materials) that higher in energy than the crystalline material [56]. Slag is commonly utilized in the cement industry for preparation of blended cement CEM II, CEM IV and CEM V where cement clinker substitution by slag range from 6 to 95 wt% [57]. This may be a partially solution to the disposal problem. Despite this use, the vast majority of this slag is still disposed in landlls [58]. One option to eliminate

this disposal of the slag in an ecologically sensitive manner is to reuse it as cementing material by alkali activation. The alkali activation system is a chemical process that transforms partially or wholly vitreous structures into cementitious skeletons. In this context, the idea of applying alkali activation was put forward in the researchers priorities. Thus, AAS cement and concrete were invented in the USSR in 1957 by Glukhowsk of the Kiev Institute of Civil Engineering, Ukraine [59], although the possibility of using alkaline activation of slag can be traced back to the 1940s [60]. In recent years AAS cement and concrete have received great attention worldwide, with applications in the Far East [61,62] and Europe [6365]. A variation of AAS cement has been used in the USA since 1987 [66,67]. Today the focus is no longer on obtaining new binders, but on developing materials with sustainably high mechanical strength and other characteristics. One option to improve some special properties of AAS system is to add one or more additives into this system. Already the literature has abundant of review papers in alkali activation system [68 78]. In addition, the previous authors reviewed the opportunities and challenges of AAS system [58]; historical background, terminology, reaction mechanisms, hydration products and materials and binders manufacture [79,80]; reinforced geopolymer composites [81]; durability [82]; geopolymers with recycled aggregate [83]; effect of various chemical activators on pozzolanic reactivity [84] and the pursuit of an alternative of PC [85]. Indeed, there is no published literature review paper which reviewed the previous works curried out on AAS modied with different types of additives. However, in this investigation, the author conducted a comprehensive literature review focused on the effect of different additives on some properties of AAS system. A review on bres, chemical admixtures, mineral admixtures and other additive materials that were added into slag in alkali activation system is presented.

2. Fibres In the literature, there are few references concerning the inuence of bres on the mechanical properties of alkali-activated cements. However, the available references can be summarized as following: Bernal et al. [86] activated slag concretes with waterglass. They added steel bres in amounts of 40 kg/m3 and 120 kg/m3 into the AAS concrete mixtures. They reported that the water absorption and permeable porous quantity were markedly improved with the addition of the bres. Both splitting tensile and exural strengths were largely improved with increasing bre volume. On the other hand, a reduction in compressive strength with bre incorporations was observed. Aydn and Baradan [87] studied the effects of length and volume fraction of steel bres on the mechanical and drying shrinkage behaviour of steel bre reinforced alkali-activated slag/silica fume (SF) mortars. Steel bres with two different lengths of 6 mm and 13 mm, and four different volume fractions of 0.5%, 1.0%, 1.5% and 2.0% were used. The composite ratio of slag/SF was 80/20. This composite was activated with waterglass and NaOH solution. The results showed a reduction in the workability and drying shrinkage with the inclusion of bres. As bres content increased as the workability and drying shrinkage decreased. Compressive strength (Fig. 1) and exural strength as well as toughness increased with the increase in bres contents and bres lengths (Fig. 2). Bernal et al. [88] modied AAS concretes by steel bres. The steel bres contents were 0%, 0.1% and 0.3%, by weight. Waterglass was used as alkaline activator. They reported that the inclusion of bre increased the splitting tensile strength, exural strength and toughness, at ages of 7, 14 and 28 days. Water absorption was reduced with the inclusion of bre, at ages of 7, 14 and 28 days. 0.3% bre content showed better performance than 0.1%.

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

31

Fig. 1. The inuence of bre length and content on the compressive strength of activated mortars [87].

Fig. 2. The effect of bre length and content on toughness of activated mortars [87].

Puertas et al. [43] modied alkali-activated mortars by polypropylene bres. The bres contents were 0%, 0.5% and 1%, by mortar volume. The source binder materials consisted of 100% slag or 50% slag coupled with 50% FA. Neat slag was activated with a mixture of Na2SiO3 and NaOH with 4% concentration of Na2O, by slag mass. The composite of slag/FA was activated with 8 M NaOH. The results showed that the incorporation of the bre did not affect positively the mechanical behaviour and freezing/thawing resistance of the mortars. The modulus of elasticity decreased with the inclusion of 1% bres. The impact resistance of the mortars after wet/dry cycles increased with the inclusion of the bre. 1% bre increased the drying shrinkage of the specimens that cured at 95% RH, whilst 1% bre reduced the drying shrinkage of the specimens cured at 21 C and 50% RH. Lee et al. [89] presented a feasibility study of strainhardening bre reinforced cementless composite using AAS based mortar and polyvinyl alcohol (PVA) bre. Test results showed the feasibility of attaining tensile strain up to 4.7% in bre reinforced AAS composite, compared to 0.02% for the mortar matrix alone. Alcaide et al. [90] reported that the inclusion of carbon bres in AAS

mortars failed to improve the compressive strength. On the other hand, using bres caused a reduction in the drying shrinkage. Puertas et al. [91] reported that the inclusion of glass bres in AAS mortar caused a reduction in the drying shrinkage with no adverse effect on its mechanical properties. Li and Xu [92,93] prepared sodium silicate activated slag/FA blends with less than 0.3% basalt bres. Tests using a split Hopkinson pressure bar (SHPB) system revealed no change in the dynamic compressive strength, but improvements in energy absorption were observed. 0.3% was estimated to be the optimal bre loading based on energy absorption. Strain hardening was not observed. The composite properties (energy absorbed, etc.) were determined to be strain rate dependent under impact loading. Silva and Thaumaturgo [94] presented the fracture toughness in mortars made of PSS (slag + MK) cement matrix reinforced by wollastonite microbres (Ca(SiO3)). The bres volumes were 1%, 2%, 3% and 5%. They reported that wollastonite microbres improved the fracture toughness of the geopolymer. The best volume of the bre was 2%. The better reinforcing efciency obtained by the geopolymer was related to

32

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

the nature of the bond between bre and matrix, and consequently the toughening mechanism in operation. PSS cement composites showed debonding and bre pullout as the main toughening mechanisms. The bre contributed to stiffness and increased the peak load. Natali et al. [95] modied some properties of the slag/MK geopolymer by adding different types of dispersed short bres. Sodium silicate solution, with a SiO2:Na2O ratio of 1.99, and 8 M NaOH solution were used as alkaline activators. The used bres were: HT-carbon bres (average bre diameter 10 lm, tensile strength 5490 MPa), commercial E-glass bres (average bre diameter 10 lm, tensile strength 2500 MPa), PVA bres (average bre diameter 18 lm, tensile strength 1800 MPa) and PVC bres (average bre diameter 400 lm, tensile strength 215 MPa). Regardless the bres type, the same content of bres (1 wt% fraction on the total mixture) was added to MK/slag mixture. The used bres were cut to obtain a 7 1 mm length. They concluded that all different types of bres had good adhesion properties, micro-cracks propagation along the matrix and created a favourable bridging effect. 1 wt% of reinforcing bre embedded in the geopolymer matrix was able to increase the exural strength from 30% up to 70% depending on bre type. Geopolymers added with PVC and carbon bres exhibited signicantly post-crack improved, resulted more enhancements in ductility after reaching the rst crack load. From the above review of literature of this part, it can be noted that the main advantage of employing steel or polypropylene or carbon or glass bres in AAS system is reducing the drying shrinkage that AAS system suffers. On the same line with this, the tendency of the post cracks was reduced by employing HT-carbon or commercial E-glass or PVA or PVC bres. On the contrary, the inclusion of steel bres in AAS matrix decreased the workability and compressive strength. However, Table 1 summarizes the effect of different type of bres on some properties of AAS system.
Table 1 Effect of different types of bres on some properties of AAS system. Author Bernal et al. [86] Fibre type Steel

3. Chemical admixtures Douglas and Brandstetr [96] reported that the addition of sodium lignosulfonate or sulfonated naphthalene-based superplasticizer did not improve the workability of AAS mortars. Wang et al. [97] studied the inclusion of water-reducing admixtures, such as sodium lignosulphonate and naphthalene-based superplasticizer in AAS mortars. They concluded that such admixtures caused a decrease in the compressive strength without improving the workability. Puertas et al. [98] studied the effect of two superplasticizer admixtures, based on vinyl copolymers and polycarboxylates, on waterglass-activated slag mortars and pastes. The dosage of the admixtures was constant and equivalent 2%, in solid mass of slag. They concluded that the vinyl copolymer-based admixture decreased mortar mechanical strength after 2 and 28 days without increasing paste workability, whilst the polycarboxylate admixture had no effect on the mechanical performance of the mortar but did not improve the workability. The calorimetric results revealed that with the admixture vinyl copolymer, the slag activation was delayed. This delay in the reaction process accounted for the lower mechanical strength of this mortar after 2 days. Polyacrylate copolymer had not a signicant effect on the activation process of the slag, which explained why its mechanical strength level was about the same as that for admixture-free mortar. Palacios and Puertas [99] investigated how several superplasticizers (polycarboxylates, vinyl copolymers, melamine and naphthalene-based) and shrinkage-reducing (polypopylenglycol derivatives) admixtures affect the mechanical and rheological properties and setting times of AAS pastes and mortars. Two alkaline activator solutions, waterglass and NaOH, were used; along with two concentrations of 4% and 5% of Na2O, by slag mass. The results indicated that all admixtures reduced liquid/solid ratio. The greatest reduction induced when the activator was NaOH and the admixture was naphthalene derivative. The admixtures

Positive effects Improved the absorption, and permeability porous quantity Improved the splitting and exural strengths Improved the compressive, exural strengths as well as toughness Reduced the drying shrinkage Increased the splitting tensile strength, exural strength and toughness Reduced the waster absorption Improved the impact resistance after wet/dry cycles Reduced the drying shrinkage of the specimens that cured at 21 C and 50% RH Improved the tensile strain Reduced the drying shrinkage Reduced the shrinkage Improvements in the energy absorption Increased the toughness, contributed to stiffness and increased peak load Improved the adhesion properties Micro-cracks propagation along the matrix and created a favourable bridging effect Created a favourable bridging effect Increased the exural strength Improved the post-crack, resulted more enhancements in ductility after reaching the rst crack load (for PVC and carbon bres)

No or negative effects Decreased the compressive strength

Aydin and Baradan [87]

Steel

Decreased the workability

Bernal et al. [88]

Steel

Decreased the compressive strength

Puertas et al. [40]

Polypropylene

No positively effects on the mechanical behaviour and freeze/thawing resistance Decreased the modulus of elasticity Increased the drying shrinkage of the specimens that cured at 95% RH No effect on the compressive strength No effects on the mechanical properties No effect on the dynamic compressive strength

Lee et al. [89] Alcaide et al. [90] Puertas et al. [91] Li and Xu [92,93] Silva and Thaumaturgo [94] Natali et al. [95]

Polyvinyl alcohol Carbon Glass Basalt Wollastonite microbers (Ca(SiO3)) HT-carbon, commercial E-glass, PVA and PVC

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

33

did not increase the slump of the pastes when waterglass was used as activator. When the activator was NaOH, owability increased slightly in polycarboxylate, melamine-based and vinyl copolymer admixtures, whilst the naphthalene-based admixture increased ow rate signicantly. The shrinkage-reducing admixture had no impact on paste slump. In the case of waterglass-AAS pastes with 4% of Na2O, the admixtures had no signicant effect on setting, with the exception of the vinyl copolymer. In the case of waterglass-AAS pastes with 5% of Na2O, the vinyl copolymer accelerated the initial set slightly, but lengthened the nal setting time. The admixtures based on melamine and polycarboxylate retarded the initial and nal sets. The naphthalene-based admixture shortened both initial and nal setting times. However, they concluded that naphthalene-based admixture combined with NaOH-AAS mortars raised mechanical strength values compared to slag mortar with no admixtures. Naphthalene-based admixture combined with NaOH-AAS pastes improved workability and retarded the initial and nal setting times compared to slag paste with no admixtures. Palacios et al. [100] studied the effect of different types of superplasticizers (naphthalene-based, melamine-based and vinyl copolymer) on the yield stress and plastic viscosity of AAS pastes. Slag was activated with NaOH solutions at pH of 11.7 and 13.6. PC paste mixtures were employed for comparison. They reported that the dosages of the superplasticizers required to attain similar reduction in the yield stress were 10-fold higher for PC than for 11.7-pH NaOH-activated slag pastes. Vinyl copolymer admixture induced the highest reduction of the yield stress in 11.7-pH NaOH-activated slag pastes. The only admixture observed to affect the rheological parameters in 13.6-pH NaOH-activated slag was naphthalene-based admixture due to its structural stability in such extremely alkaline media. Dosages as low as 1.26 mg naphthalene/ g slag were observed to induce the maximum reduction in yield stress (98%). The presence of the three types of superplasticizers lowered the plastic viscosity in the 11.7-pH NaOH-AAS pastes and only naphthalene-based admixture in 13.6-pH NaOH-AAS pastes. Palacios et al. [101] established the effect of HRWR on the rheology and setting of AAS pastes and mortars. Two different alkaline activators were used. The activators were waterglass solution with SiO2/Na2O = 1 and NaOH solution with 4% Na2O, by mass. Five different chemical admixtures were employed four HRWR: polycarboxylate-based, melamine formaldehyde derivative, naphthalene formaldehyde derivative and vinyl copolymer; and shrinkage reducing agent. The dosages of admixtures were 0%, 0.3%, 0.5%, 1.0%, 1.5% and 2.0%, by mass of the binder. They reported that the greatest drop in the yield stress in AAS cements was found when naphthalene-derivative HRWR was added to NaOH-activated slag pastes and mortars because of its inherent stability in alkaline media. The other admixtures incorporated in AAS pastes and mortars did not signicantly modify their rheological parameters. Increasing the mixing time could dramatically retard the setting of waterglass-activated slag pastes, with initial and nal setting times of approximately 3 and 10 h, respectively. Bakharev et al. [102] studied the effect of superplasticizer, airentraining agent (AEA) and water-reducing on some properties of AAS concretes. The activators were liquid sodium silicate (47% Na, mass of slag) and a multi-compound activator of NaOH + Na2CO3 (8% Na, mass of slag).The superplasticizer based on modied naphthalene formaldehyde polymers, water-reducing based on lignosulphonates and AEA with a soluble salt of an alkyl aryl sulphonate. The slump results showed that sodium silicate activated slag concrete with 4% Na, had a slump of 55 mm immediately after mixing. The same concrete with water-reducing of 6 ml/kg had a slump of 80 mm and had better workability than concrete without water-reducing. Water-reducing admixture of 10 ml/kg in the same concrete mixture produced a slump of 200 mm. AEA at a dosage of 6 ml/kg produced a considerable improvement in workability.

Superplasticizer admixture showed very uid concrete with slump more than 200 mm, but the concrete lost its uidity after 10 min. The AAS concrete activated with sodium carbonate and NaOH and containing 10 ml/kg of water-reducing exhibited a slump exceeded 200 mm. However, they recorded that using water-reducing and AEA admixtures were the most effective for improving workability. The compressive strength results showed that using superplasticizer admixture caused 25% loss of 28 days strength, the use of waterreducing (based on lignosulphonates) reduced the early strength up to 14 days. AEA had some effect on early strength up to 7 days, after that strength development was similar to AAS concrete without admixtures (Fig. 3). The drying shrinkage results showed that the specimens with superplasticizer exhibited the highest drying shrinkage, AAS specimens without admixtures came in the second place and the specimens admixed with water-reducing came in the third place, whilst the specimens admixed with AEA came in the fourth place. They concluded that AEA admixture was the most suitable for use in AAS concretes and it was not desirable to use superplasticizer admixture in AAS concretes. Bilim et al. [103] studied the effect of shrinkage-reducing (SHR) and superplasticizing and set-retarding admixtures (SSR) on some properties of slag pastes and mortars activated with alkaline activator. Liquid sodium silicate was used to activate the slag at two sodium concentrations of 4% and 6%, by mass of slag. Liquid sodium silicate and NaOH were blended to obtain 0.75 and 1 modulus of SiO2/Na2O. SHR based on polypropylenglycol and SSR based on modied polymer liquid were used as chemical additives. 1% of each admixture, by binder mass, was added to the activator solution. Fixed water/binder (w/b) ratio of 0.5 was used to prepare paste and mortar specimens. The results indicated that both SSR and SHR admixtures generally did not have an impact on setting times of the pastes, with the exception for the sodium silicate activated slag paste with Ms = 0.75 and 4% Na, where SSR caused longer initial and nal setting times in comparison with those mixture without admixture. Both SSR and SHR chemical admixtures reduced compressive and exural strengths of the AAS mortars at age of 2 days. After 2 days, SSR had no impact on the compressive and exural strengths, whereas SHR increased the exural strength of AAS without changing in the compressive strength. SSR admixture had no impact on AAS mortars after carbonation, whilst SHR admixture somewhat decreased the carbonation depths of AAS mortars. Both SSR and SHR reduced the drying shrinkage of AAS mortars, but these shrinkages were still higher than that of PC. Collins and Sanjayan [53] studied the drying shrinkage of AAS concretes without or with 1.5% shrinkage-reducing chemical admixture, compared to PC concrete. The slag was supplied with gypsum (2% SO3), which was blended with slag. The activators were powdered sodium metasilicate and hydrated lime. The results showed that AAS concrete had 1.6 times greater drying shrinkage than PC concrete at 56 days. The AAS concrete containing 1.5%

Fig. 3. Compressive strength of AAS concrete prepared using different admixtures [102].

34

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

shrinkage reducing showed lower drying shrinkage compared to pure AAS concrete (Fig. 4). Palacios and Puertas [104] studied the effect of shrinkage-reducing admixture, based on polypropylenglycol, on the compressive strength, exural strength and dimensional stability of waterglass-activated slag mortars. The concentrations of shrinkage-reducing were 0%, 1% and 2%, by slag mass. Specimens were cured at 20 2 C and 99% RH for 48 h. Then, the specimens subsequently removed from the moulds and cured until testing date at either 99% or 50% RH. The results showed that the addition of 1% shrinkage-reducing did not signicantly modify the compressive strength or the exural strength for the specimens cured at either 99% or 50% RH, although after 28 days, the exural strength of the specimens cured at 99% RH was 25% higher than recorded for the mortars without admixture. The addition of 2% shrinkagereducing increased the compressive strength and exural strength at both curing conditions, although more signicantly in the mortars cured at 99% RH: exural strength was increased by approximately 75% at 7 days and by more than 100% at 28 days related to the mortars without shrinkage-reducing. The shrinkage results indicated that the shrinkage-reducing reduced the shrinkage by up to 85% and 50% when the AAS mortar specimens were cured at 99% and 50% RH, respectively. They reported that the mechanism primarily involved in shrinkage reduction was the decrease in the surface tension of pore water prompted by the admixture. From the above review of literature of this part, it can be noted that the main advantage of employing some chemical admixtures as shrinkage-reducing (based on polypropylenglycol) or airentraining (with a soluble salt of an alkyl aryl sulphonate) or water-reducing (based on lignosulphonates) in AAS matrix reduced the drying shrinkage that AAS system suffers. Other chemical admixtures as superplasticizer (based on polycarboxylate) or superplasticizer (based on vinyl copolymer) or superplasticizer (based on melamine) or superplasticizer (based on naphthalene) retarded the setting time that AAS suffers, but depending on activator type and Na2O concentration. Shrinkage-reducing (based on polypropylenglycol) somewhat decreased the carbonation depth that AAS suffers. On the contrary, some chemical admixtures as superplasticizer (based on lignosulfonate) or superplasticizer (based on sulfonated naphthalene) or superplasticizer (based on vinylcopolymer) decreased the compressive strength and had no effect on the workability of AAS system. However, Table 2 summarizes the previous researches that studied the effect of chemical admixtures on some properties of AAS system. 4. Mineral admixtures 4.1. Silica fume and quartz powder 4.1.1. Workability Collins and Sanjayan [105] partially replaced slag with 10% condensed silica fume (CSF) in alkali-activated concrete activated with

powdered sodium metasilicate and hydrated lime. They showed that alkali-activated slag/CSF concrete had signicantly less workability than neat AAS concrete. 4.1.2. Strength Rashad et al. [106] partially replaced slag with quartz powder (QP) at replacement levels of 0%, 5%, 10%, 15%, 20%, 25% and 30%, by weight, in alkali-activated pastes. Sodium silicate was used as activator. The results showed that the compressive strength value, at age of 28 days, increased with increasing QP content (Fig. 5). The composite of 30% QP coupled with 70% slag gave the highest compressive strength. Rashad and Khalil [44] activated slag with sodium silicate. The slag was partially replaced with SF at levels of 0%, 5%, 10% and 15%, by weight. The paste specimens were cured at 20 1 C and 90 5% RH. They concluded that the inclusion of SF in slag enhanced and increased the compressive strength. The optimal content of SF that gave the highest compressive strength at ages of 7 and 28 days was 5% (Fig. 6). Escalante-Garca et al. [107] activated slag mortars with 6 wt% Na2O equivalent of NaOH. The slag was partially replaced with SF at levels of 0%, 5%, 10%, 15% and 20%, by weight. The compressive strength increased with the presence of SF up to 15%, then decreased with the replacement level of 20%. The replacement levels of SF at 510% gave the highest compressive strength. In another investigation Escalante-Garca et al. [108] reported that partially replacing 10% of slag with geothermal silica waste in mortar specimens, activated with 6 wt% Na2O equivalent of NaOH and waterglass, enhanced the formation of hydration products and produced denser microstructure in comparison with the mortar containing neat slag. Aydn [109] activated slag mortars with NaOH and sodium silicate. Slag was partially replaced with SF at levels of 0%, 10% and 20%, by weight. The specimens were cured at 20 C and 90% RH for 5 h, then in steam at 70 C for 6 h. The results showed that 10% SF enhanced the compressive strength by 4.24%, whilst 20% SF reduced the compressive strength by 15.42%. On the contrary, the exural strength decreased by 29.59% and 32.65% with the inclusion of 10% and 20% SF, respectively. Collins and Sanjayan [105] reported that partially replacing 10% of slag with condensed SF in concrete specimens, activated with sodium metasilicate powder and hydrated lime, improved the one-day compressive strength by 10% higher than that of AAS concrete without SF. Rousekov et al. [110] reported that the NaOH activated pastes prepared with the proportion of slag: SF of 1:2 and 1:1 gave higher strength at age of 56 days than the proportions of 2:1 when NaOH/SF was 20%. Roy and Silsbee [66] reported higher early and later strength of alkali-activated pastes containing <4% CSF addition. Douglas and Brandstetr [96] reported that the compressive strength of alkali-activated mortars activated with sodium silicate and modied with 8% SF and 2% lime reduced the compressive strength at age of 1 day, but increased the compressive strength at ages of 7 and 28 days, compared to that activated slag mortar without modication. They also reported that mortar activated with sodium silicate solution gave higher compressive strength when the slag binder containing 2% lime plus 1% Na2SO4 plus 8% SF. 4.1.3. Durability, re resistance and drying shrinkage Rashad et al. [106] partially replaced slag with QP at replacement levels of 0%, 5%, 10%, 15%, 20%, 25% and 30%, by weight, in alkali-activated pastes. Sodium silicate was used as activator. The pastes were autoclaved at a pressure of 8 bars and a temperature of 170 C at different autoclaving times of 0.5, 1, 4, 6 and 10 h. They reported that the replacement of slag with QP increased the strength values of autoclaved specimens. The composite of 30% QP coupled with 70% slag gave the highest compressive strength after all autoclaving conditions.

Fig. 4. Drying shrinkage of concrete prisms subjected to sealed curing at 23 C for 24 h followed by exposure to 23 C and 50% RH [53].

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 Table 2 Effect of chemical admixtures of some properties of AAS system. Author Douglas and Brandster [96] Admixture Superplasticizer (sodium lignosulfonate or sulfonated naphthalene-based Superplasticizer (sodium lignosulfonate or sulfonated naphthalene-based) Superplasticizer (vinylcopolymer-based) Superplasticizer (polycarboxylate-based) Positive effects No or negative effects No effect on the workability

35

Wang et al. [97]

No effect on the workability. Decreased the compressive strength No effect on the workability Decreased the compressive strength Delayed slag activation process No effect on the workability No effect on the mechanical performance No signicant effect on the slag activation process No effect on the workability when waterglass was used as activator No signicant effect on the setting time when waterglass with 4% Na2O was used as activator

Puertas et al. [98]

Palacios and Purtas [99]

Superplasticizer (polycarboxylate-based)

Reduced the liquid/solid ratio Improved the workability when NaOH was used as activator Retarded the initial and nal setting times when 5% Na2O of waterglass was used as activator Reduced the liquid/solid ratio Improved the workability when NaOH was used as activator Lengthened the nal setting time when waterglass with 5% Na2O was used as activator Reduced the liquid/solid ratio Improved the workability when NaOH was used as activator Retarded the initial and nal setting times when 5% Na2O of waterglass was used as activator Reduced the liquid/solid ratio Improved the workability when NaOH was used as activator Retarded the initial and nal setting times when NaOH was used as activator Improved the compressive strength when NaOH was used as activator Reduced the liquid/solid ratio

Superplasticizer (vinyl copolymer-based)

No effect on the workability when waterglass was used as activator Slightly accelerated the initial setting time when waterglass with 5% Na2O was used as activator

Superplasticizer (melaminebased)

No effect on the workability when waterglass was used as activator No signicant effect on the setting time when waterglass with 4% Na2O was used as activator

Superplasticizer (naphthalenebased)

No effect on the workability when waterglass was used as activator No signicant effect on the setting time when waterglass with 4% Na2O was used as activator Shortened the initial and nal setting times when 5% Na2O waterglass was used as activator No effect on the workability when waterglass was used as activator No signicant effect on the setting time when waterglass with 4% Na2O was used as activator

Shrinkage-reducing (polypopylenglycol derivatives)

Palacios et al. [100]

Superplasticizer (naphthalene-based)

Superplasticizer (melamine-based)

Superplasticizer (vinyl copolymer-based)

Lowered the yield stress when either of 11.7-pH or 13.6-pH NaOH was used as activator Improved the uidization when 13.6-pH NaOH was use as activator Lowered the plastic viscosity when either of 11.7-pH or 13.6-pH NaOH was used as activator Lowered yield stress when 11.7-pH NaOH was used as activator Lowered plastic viscosity when 11.7-pH NaOH was used as activator Lowered (the highest reduction) yield stress when 11.7-pH NaOH was used as activator Lowered plastic viscosity when 11.7-pH NaOH was used as activator

No effect on the owability when 13.6pH NaOH was used as activator

No effect on the owability when 13.6pH NaOH was used as activator

Palacios et al. [101]

HRWR(polycarboxylate-based) HRWR (melamine formaldehyde derivative) HRWR (naphthalene formaldehyde derivative) HRWR (vinyl copolymer-based) Shrinkage reducing

Did not signicantly modify its rheological parameter Did not signicantly modify its rheological parameter Lowered (the highest reduction) the yield stress when NaOH was used Did not signicantly modify its rheological parameter Did not signicantly modify its (continued on next page)

36 Table 2 (continued) Author Bakharev et al. [102]

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

Admixture Superplasticizer (naphthalene formaldehyde polymers) Water-reducing (based on lignosulphonates) Air-entraining (with a soluble salt of an alkyl aryl sulphonate)

Positive effects Improved the workability, but mixture lost its uidity after 10 min Improved the workability (the most effective) Reduced the drying shrinkage Improved the workability (the most effective admixture) Reduced the drying shrinkage (the lowest reduction in drying shrinkage) Increased the exural strength after age of 2 days Somewhat decreased the carbonation depth Reduced the drying shrinkage Longer the initial and nal setting times when sodium silicate with 4% Na2O and Ms = 0.75 was used Reduced the drying shrinkage

No or negative effects rheological parameter Reduced the compressive strength Increased the drying shrinkage (the highest increase in the drying shrinkage) Reduced the early compressive strength up to 14 days No effect on the compressive strength after 7 days

Bilim et al. [103]

Shrinkage-reducing (based on polypropylenglycol)

No effect on the setting times Reduced the compressive and exural strengths at age of 2 days No impact on the compressive strength after age of 2 days No effect on the setting times

Superplasticizer set retarder (based on modied polymer)

Reduced the compressive and exural strengths at age of 2 days No impact on the compressive and exural strengths after age of 2 days No impact on the carbonation depth 1% did not signicantly modify the compressive and exural strengths

Collins and Sanjayan [53] Palacios and Puertas [104]

Shrinkage-reducing Shrinkage-reducing (based on polypropylenglycol)

Reduced the drying shrinkage 2% improved the compressive and exural strengths Reduced the drying shrinkage

Fig. 5. Compressive strength development versus QP content [106].

Fig. 6. Compressive strength development versus SF content [44].

Rashad and Khalil [44] partially replaced slag with SF at levels of 0%, 5%, 10% and 15%, by weight, in alkali-activated pastes that designated as SF0, SF5, SF10 and SF15, respectively. Sodium silicate was used as activator. The specimens were cured at 20 1 C and 90 5% RH. After 28 days, some specimens were crushed directly in compression without exposure to elevated temperatures. Other specimens were exposed to 400, 600, 800 and 1000 C, for 2 h. The results indicated that the compressive strengths after ring increased with the presence of SF up to 800 C, then decreased at 1000 C in comparison with neat AAS specimen. Both residual strength and relative strength of hardened neat AAS paste at 1000 C were signicantly higher than that of hardened alkali-activated slag/SF. They concluded that neat AAS was more efcient to resist elevated temperature, especially at 1000 C, than alkali-activated slag/SF. They found that the volume of the specimens made of neat slag did not exhibit any change in their dimensions and showed volume stability. On the contrary, the specimens made of slag/SF mixtures exhibited slightly volume instability after exposure to elevated temperatures. This volume instability marginally

increased as the content of SF increased. Rashad and Khalil [44] also studied the repeated thermal shock resistance of the previous mixtures. After drying, the specimens were exposed to 800 C for 40 min, then quenched into water at room temperature for 5 min. This called one cycle. Such cycles were repeated until the specimens were broken, damaged or deteriorated. The results showed that the hardened neat AAS paste had thermal shock 1.75 times greater than other hardened alkali-activated slag/SF pastes (Fig. 7). They concluded that neat AAS exhibited more resistance of sudden thermal shock than alkali-activated slag/SF. Aydn [109] studied the percentage of permeable volume and drying shrinkage of activated slag mortars with NaOH and sodium silicate. Slag was partially replaced with SF at levels of 0%, 10% and 20%, by weight. The specimens were cured at 20 C and 90% RH for 5 h, then in steam at 70 C for 6 h. After that, the specimens were kept at 20 C and 55% RH. The results showed a reduction in the percentage of permeable volume by 11.4% and 17.93% with the inclusion of 10% and 20% SF, respectively. The measurement of

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

37

Fig. 7. Thermal shock resistance of hardened neat AAS and alkali-activated slag/SF pastes [44].

drying shrinkage of the specimens started from the end of heat treatment up to 4 months. The results showed a reduction in the drying shrinkage with the inclusion of SF. The inclusion of 10% SF showed the lowest drying shrinkage. From the above review of literature of this section, it can be noted that the main advantage of employing SF and QP in AAS system is enhancing the compressive strength and producing denser microstructure. On the same line with this, the inclusion of SF in AAS matrix can reduce the drying shrinkage that AAS system suffers. On the contrary, the inclusion of SF in AAS matrix decreased the residual compressive strength after exposure to 1000 C. On the same line with this, the inclusion of SF in AAS matrix reduced the thermal chock resistance by about 42.85%, related to the neat AAS paste. The inclusion of SF and QP in AAS matrix is still needs more investigations that studying the effect of these materials on setting time, formation of salt eforescences, carbonation resistance and the tendency to cracks during curing. However, Table 3 summarizes the effect of SF and QP on some properties of AAS system. 4.1.4. As activator Rousekov et al. [110] explored new type of alkali activator manufactured from SF. The data obtained showed that the optimal proportion of slag and SF alkali activator allowed an increase in the

compressive strength and signicant increase in the resistance to water (improve the hydraulicity of the binder). Further, the improved properties of the composites containing the alkali activator over those prepared with NaOH conrmed the higher effectiveness of SF activator over the NaOH. Brew and Mackenzie [111] exploited solgel type condensation reactions between sodium silicate, formed in situ by alkaline dissolution of SF, and a solution of sodium aluminate. They concluded that the specimens prepared from this activator showed good compressive strength, indicated that the products displayed all the characteristics of typical aluminosilicate geopolymers. Bernal et al. [112] activated slag/metakaolin (MK) 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 the increase contents of slag in the binders, which promoted the coexistence of aluminosilicate reaction products along with calcium silicate hydrate gel. ivica [113] studied the acidic resistance of slag/PC with chemZ ically modied silica fume (MSF). The results of this study indicated that MSF enabled the possibility of a signicant solution of general insufcient acidic resistance of slag/PC materials, where pore structure, mechanical properties and high acidic resistance ivica [114] emwere found on the slag/PC when MSF was applied. Z ployed SF as activator to activate slag/PC, other traditional activators as NaOH and waterglass were used for comparison. The results showed that the composite of 100/0 activated with SF activator gave higher compressive strength at 28 days than that activated with NaOH. The composite of 30/70 activated with SF activator gave higher compressive strength at age of 28 days than that acti ivica [115] conducted alkali activator manvated with waterglass. Z ufactured from SF. The experimental results showed that the SF activator as a very alkali activator for the binding systems based on the combination of PC, SF and slag and slag alone. The positive effect of SF activator seemed to be based on the intensication of the production of CSH and the densifying of the forming pore structure. This activator was more effective at slag binder. Table 4 summarizes the previous researches about the effect of using SF as activator on AAS system.

Table 3 Effect of SF and QP on some properties of AAS system. Author Collins and Sanjayan [105] Rashad et al. [106] Rashad and khalil [44] Escalante-Garca et al. [107] Escalante-Garca et al. [108] Aydn [109] Collins and Sanjayan [105] Rousekov et al. [110] Roy and Silsbee [66] Douglas and Brandstetr [96] Content 10% CSF Up to 30% QP Up to 15% SF Up to 20% SF 10% geothermal silica waste 10% CSF Slag: SF of 1: 2 and 1: 1 <4% CSF 8% SF + 2% lime 8% SF + 2% line + 1% Na2SO4 Up to 30% QP Up to 15% SF Positive effects Increased the compressive strength Increased the compressive strength Increased the compressive strength (up to 15%) Gave denser microstructure Increased the compressive strength Increased 1 day compressive strength Increased the compressive strength compared to slag: SF of 1: 1 Increased the early and late ages compressive strength Increased 7 and 28 days compressive strength Increased the compressive strength Increased the compressive strength of autoclaved specimens Increased the residual compressive strength up to 800 C Reduced the percentage of permeable volume Reduced the drying shrinkage No or negative effects Decreased the workability

Decreased the compressive strength (20% SF)

Decreased 1 day compressive strength

Rashad et al. [106] Rashad and khalil [44]

Reduced the residual compressive strength at 1000 C Reduced the thermal shock resistance

Aydn [109]

10 and 20

38 Table 4 Effect of SF as activator on AAS system. Author Rousekov et al. [110] Brew and Mackenzie [111] Bernal et al. [112] ivica [113] Z ivica [114] Z Activator SF SF SF or RHA Modied SF SF

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

Effects Increased the compressive strength Increased the resistance to water (improved the hydraulicity of the binder) Increased compressive strength Similar trends of mechanical strength development as a function of activation compared with specimens activated with commercial silicate solution Higher acidic resistance of slag/PC At slag/PC ratio of 100/0, specimens activated with SF gave higher compressive strength at 7 days age than specimens activated with NaOH At slag/PC ratio of 30/70, specimens activated with SF gave higher compressive strength at 28 days age than specimens activated with waterglass Intensication of the production of CSH

ivica [115] Z

SF

4.2. Fly ash 4.2.1. Workability Yang and Song [116] studied the workability of AAS and alkaliactivated y ash (AAFA) activated with a combination of sodium silicate and NaOH powders. They noted that higher initial ow in AAFA than that in the AAS. Yang et al. [117] reported that AAS mortars exhibited slightly less ow than AAFA mortars for the same mixing condition. Yang et al. [118] used sodium silicate powder to activate either slag or FA mortar. They used constant w/b ratio of 0.5 and sand to binder of 3. They reported that the AAS mortar had lower workability than FA-based alkali-activated mortar. Collins and Sanjayan [105] explained that partial replacement slag with 10% ultrane FA in AAS concrete activated with powdered sodium metasilicate and hydrated lime, showed higher workability than neat AAS concrete. Talling and Brandstetr [119] reported that FA might improve the workability of the fresh mixture of AAS cement. Rashad [120] studied the workability of slag/FA-based geopolymer concrete mixtures. The ratios of slag/FA were 15/85, 10/ 90, 5/95 and 0/100, by weight. The results indicated that as FA content increased as the workability increased. Wang et al. [97] reported that an addition of FA (below 10%) had little effect on improving the workability of the mixture. On the contrary, Parameswaran and Chatterjee [121] reported that an Indian FA did not improve the workability even at 40% by weight of total binder, implying that the characteristics of FA are important. Table 5 summarizes the previous researches about the effect of FA on the workability of AAS system.

cate with a 2.00 SiO2/Na2O ratio led to rapid setting. An increase in the proportion of FA shortened the setting time. Kumar et al. [123] replaced slag with FA at levels of 50%, 65%, 75%, 85%, 95% and 100% in slag/FA-based geopolymers. They reported that the setting time increased as the content of FA increased.

4.2.2. Setting time Sugama et al. [122] studied the setting time of alkali-activated slag/Class F FA. Slag was partially replaced with FA at levels of 0%, 10%, 20%, 30%, 40% and 50%, by weight. Sodium silicate was used as alkaline activator with SiO2/Na2O molar ratios of 3.22, 2.50 and 2.00. The results indicated that the setting time increased with the increase in FA content when using sodium silicate with molar ratios of 3.22 and 2.50. On the contrary, using sodium sili-

4.2.3. Strength Puertas et al. [124] investigated both compressive and exural strengths of mortar consisted of 50% slag coupled with 50% FA activated with 8 M NaOH versus neat slag mortar activated with a mixture of Na2siO3 and NaOH with 5% concentration of Na2O, by slag mass. The results showed lower compressive and exural strengths at ages of 2 and 28 days with the inclusion of 50%FA. Shi and Day [125] partially replaced slag with FA at levels of 0% and 50%, by weight. They activated the mixtures with either sodium silicate or NaOH. They reported that when NaOH was used as an activator, the slag replacement with FA did not show a significant effect on compressive strength. The compressive strength decreased with the inclusion of FA in the slag matrix, when sodium silicate was used as an activator. Puertas and Fernndez-Jimnez [126] activated the composite of 50% slag/50% FA with 10 M NaOH. The liquid/solid ratio was 0.35. Two curing temperatures of 22 and 65 C were used. The pastes were maintained at 65 C during the rst 5 h. For the rest of the curing time, the specimens were maintained at ambient temperature and 98% RH; the same as the specimens cured at 22 C. The results showed that higher compressive and exural strengths of the specimens cured at 22 C at ages of 7 and 28 days compared to those specimens cured at 65 C. Lu [127] produced waterglass-activated slag/FA concrete with strength of 36 MPa at 7 days and 56 MPa at 28 days when cured in a fog room in the temperature range of 1020 C. Smith and Osborne [128] and Bijen and Waltje [129] found that waterglass-activated slag/FA blends showed very low strength, but NaOH activation gave higher strength. On the contrary, Dai and Cheng [130] found that waterglass was much more effective than NaOH. Smith and Osbrone [63] investigated comprised of nely slag, FA and NaOH. They found good early strength values, but there was

Table 5 Effect of FA on the workability of AAS system. Author Yang and Song [116] Yang et al. [117] Yang et al. [118] Collins and Sanjayan [105] Talling and Brandstetr [119] Rashad [120] Wang et al. [97] Parameswaran and Chatterjee [121] Increased workability p p p p p p p No effect Notes AAS had less workability than AAFA AAS had less workability than AAFA AAS had less workability than AAFA

Little improvement

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

39

little gain in the later ages. On the same line with this, Douglas and Brandstetr [96] activated slag mortars with sodium silicate. They modied the mortars by 2% lime, 2% lime + 5% FA and 2% lime + 10% FA. The w/b ratio was xed at 0.41. The results indicated that the inclusion of 5% FA gave lower strength at age of 1 day in comparison with the mixture did not include FA, whist it gave higher strength at ages of 7 and 28 days. The inclusion of 10% FA gave lower strength at ages of 1 day and 7 days, whist it gave higher strength at age of 28 days in comparison with those containing 5% FA. On the contrary, Collins and Sanjayan [105] explained that partial replacement slag with 10% ultrane FA in AAS concrete activated with powdered sodium metasilicate and hydrated lime, showed 24% better 1-day strength, but lower the strength at ages beyond 28 days, compared to neat AAS concrete. Other authors as Talling and Brandstetr [119] reported that partial substitution slag with 10% FA, by weight, might result in an increase in strength. On the contrary, Wang et al. [97] reported that the addition of FA (below 10%) in AAS mortar often cause slight reduction in compressive strength under normal curing. However, substitution slag with FA at higher proportions reduced the strength noticeably [61,121]. On the same line with this, Rashad [120] reported that substitutions slag with higher proportions of FA (i.e. 85%, 90%, 95% and 100%) in slag/FA-based geopolymer concretes activated with mixture of NaOH and sodium silicate, reduced the compressive strength, splitting tensile strength and exural strength. The compressive strength, splitting tensile strength and exural strength decreased as the content of FA increased. 9 a et al. [131] prepared pastes of slag and FA in proportions Garci of 100/0, 75/25, 50/50, 25/75 and 0/100, by weight, activated with sodium silicate. The used moduli (SiO2/Na2O) were 0, 0.75, 1, 1.5 and 2. The% Na2O was added at 4%, 6% and 8% related to the binder weight. The pastes were cured at 75 C for 24 h and then at 20 C up to 28 days. The results indicated an increase in the compressive strength with increasing slag content. The highest compressive strength was noted for the neat slag (8085 MPa) when 4% of Na2O and 1.5 SiO2/Na2O were used. For 0/100 paste, the higher% Na2O, the better the strength, whereas the highest strength of 25 MPa was reached using modulus 1. For the composite of 75/ 25, the strengths were (5660 MPa) at 4% Na2O and the modulus 1 and 1.5. For the composite of 50/50 paste, the strengths were (4548 MPa) at 4% Na2O and the best modulus was 11.5. For the composite of 25/75 paste, the strength reached 3035 MPa at 4% Na2O and modulus 1.5. Weiguo et al. [132] replaced slag with FA at levels of 0%, 30%, 40%, 50%, 60%, 70% and 100%, by weight. Waterglass with SiO2/Na2O ratio of 2.4 and NaOH which used to adjust the SiO2/Na2O ratio to 1.0 was used as alkaline activator. The results showed that the compressive and exural strengths decreased as the FA content increased. Kumar et al. [123] replaced slag with FA at levels of 50%, 65%, 75%, 85%, 95% and 100%, by weight, in slag/FA-based geopolymers. The specimens were cured at either 27 C or 60 C. The reduction in compressive strength was observed with the inclusion of FA. As FA content increased as the compressive strength decreased. Kim and Kim [133] replaced slag with FA at levels of 0%, 50% and 100%, by weight, in slag/FA-based geopolymers. 2.78 M NaOH was used as activator. The compressive strength results of the mortars, at ages of 1, 7 and 28 days, decreased as the content of FA increased. Puertas et al. [134] activated slag/FA pastes with NaOH solution. The parameters of this study were: activator concentration (NaOH 2 and 10 M), curing temperature (25 and 65 C) and slag/FA ratio (100/0, 70/30, 50/50, 30/70 and 0/100). In the curing temperature process, the pastes were maintained at 65 C during the rst 5 h. for the rest of the curing time, the specimens were maintained at ambient temperature and 98% RH; the same as the specimens cured at 25 C. The compressive strength results at 1 day showed

an increase in strength with the increase in slag content. 65 C coupled with 10 M of NaOH gave the highest strength. At 7 days, 70/30 gave the highest strength at 25 C coupled with 10 M of NaOH, whilst the strength increased with the increase in slag content, in the remaining conditions. At age of 28 and 90 days, as the slag content increased as the compressive strength increased. 25 C coupled with 10 M of NaOH seemed to be the optimum condition, followed by 65 C coupled with 10 M of NaOH, followed by 25 C coupled with 2 M of NaOH. 56 C coupled with 2 M of NaOH came in the last place. Bakharev et al. [38] studied the alkali-activated Australian slag mortar using different activators. In addition, they prepared mixture of alkali-activated slag/FA. The ratio of slag/FA was 30/70. They concluded that FA introduced in AAS at 30% reduced the compressive strength of the mortar. Smith and Osborne [63] investigated cements made of the combination of 60% nely slag and 40% FA activated with 7% NaOH. They found that early strength properties were good but there was little gain in strength beyond 28 days. Aydn [109] activated slag mortars with NaOH and sodium silicate. Slag was partially replaced with FA at levels of 0%, 20% and 40%, by weight. The specimens were cured at 20 C and 90% RH for 5 h, then in steam at 70 C for 6 h. The results showed a reduction in the compressive strength with the inclusion of FA. The reduction in the compressive strength was 3.11% and 13% with the inclusion of 20% and 40% FA, respectively. Zhang et al. [135] solidied municipal solid waste incinerator (MSWI) FA with the Na2SiO3-activated slag. The Na2SiO3 activated slag was added to MSWI FA at 25%, 30%, 35% and 45%, by weight. The compressive strength results showed that 45% gave the highest compressive strength at 7 days, whilst 35% gave the highest compressive strength at 28 and 60 days. Guerrieri and Sanjayan [136] presented the compressive strength of geopolymer pastes made of different combinations of slag/FA ratios. The ratios of slag/FA were 100/0, 65/35, 50/50, 35/ 65 and 0/100, by weight. The alkaline activators were mixtures of sodium silicate liquid and 8 M NaOH. The activators were mixed in the proportion so that Ms were 0, 0.5, 1.0, 1.5 and 2.0. Industrial grade-powdered sodium metasilicate with hydrated lime was also used. The concentrations of the activators were 4% and 8% Na. After casting, the specimens were kept at 23 1 C and 50 5% RH for 2 h and then were cured at 80 1 C and 95 3% RH for 22 h, after that the specimens were allowed to cool down to room temperature. The compressive strength was measured at 24 h after curing period. The results showed that the composite of 35/65 achieved the highest compressive strength at 8% Na and Ms 1.01.5, followed closely by the 65/35, followed by 100/0 and followed by 50/50. Chi and Huang [137] presented the compressive strength and exural strength, at ages of 7, 14 and 28 days, of geopolymer mortars made of different combinations of slag/FA ratios of 100/0, 70/30, 50/50, 30/70, 0/100, by weight. Sodium silicate with modulus ratio of 1 was used as alkaline activator. Two concentrations of Na2O of 4% and 6%, by cementitious weight, were employed. The results showed that the composition of 50/50 achieved the highest compressive strength and exural strength followed by 70/30, 100/0, 30/70 and 0/100, respectively, at both 4% and 6% Na2O. Goretta et al. [138] measured the compressive strength of Class C FA, slag and sodium silicate alkali-activated concrete. The aggregate constituted 52% and an alkali activator 11.2% of the total mass; the mass ratio of silicate to FA + slag was 0.29. 50 mm diameter cylinders having a 1:2 diameter-tolength ratio were used for testing compressive strength. They reported that compressive strength of 35 MPa could be obtained at 14 days. Table 6 summarizes the pervious researches that studied the effect of FA on the strength of AAS system. 4.2.4. Durability and drying shrinkage Sugama et al. [122] exposed alkali-activated slag/Class F FA to CO2-laden H2SO4 with pH value of 1.1 for 15 days at 90 C, after

40 Table 6 Effect of FA on the strength of AAS system. Author Puertas and Chatterjee [121] Shi and Day [125] Smith and Osborne [128] Bijen and Waltje [129] Douglas and Brandstetr [96] 5 10 Collins and Sanjayan [105] Talling and Brandstetr [119] Wang et al. [97] Wang [61] Parameswaran and Chatterjee [121] Rashad [120] Garca et al. [131] 10

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

% incorporation 50 50

Positive effects

Negative effects p p p p

Notes

p p p p p p p p p p p p p p 0% optimal (8085 MPa)

Activated with waterglass Activated with NaOH Activated with waterglass Activated with NaOH At 1 day At 7 and 28 days At 1 and 7 days At 28 days At 1 day Beyond 28 days

10 Below 10 Higher proportions Higher proportions 85, 90, 95 and 100 0, 25, 50, 75 and 100

4% Na2O, 1.5 SiO2/Na2O 25% optimal (5660 MPa) p 50% optimal (4548 MPa) p 75% optimal (3035 MPa) p 4% Na2O, 1 and1.5 SiO2/Na2O 4% Na2O, 11.5 SiO2/Na2O 4% Na2O, 1.5 SiO2/Na2O As FA content increased as strength decreased As FA content increased as strength decreased p Curing 65 C for 5 h, then in room. 10 M NaOH at 1 day Curing 25 C for 5 h, then in room. 10 M NaOH. at 7 days At 28 and 90 days Good at early ages and little gain beyond 28 days At 7 days At 28 and 60 days 8% Na2O, 11.5 SiO2/Na2O

Weiguo et al. [132] Kumar et al. [123] Kim and Kim [133] Puertas et al. [134]

0, 30, 40, 50, 60, 70 and 100 50, 65, 75, 85, 95 and 100 0, 50 and 100 0, 30, 50, 70 and 100

p 0% optimal p 0% optimal p 0% optimal p 30% optimal p 0% optimal p

Bakharev et al. [38] Smith and Osborne [63] Aydn [109] Zhang et al. [135] Guerrieri and Sanjayan [136]

30 40 20 and 40 55, 65, 70 and 75 0, 35, 50, 65 and 100

p p

Chi and Huang [137]

0, 30, 50, 70 and 100

p 55% optimal p 65% optimal p 65% optimal p 35% second place p 0% third place p 50% fourth place p 50% optimal p 30% second place p 0% third place p 100% last place

4% or 6% Na2O, 1 SiO2/Na2O

autoclaving at 100, 200 and 300 C. The paste specimens were prepared by varying two parameters, slag/FA weight ratios of 100/0, 90/10, 70/30 and 50/50 and SiO2/Na2O molar ratios, in the sodium silicate, of 3.22, 2.50 and 2.00. For all autoclaved temperatures, the weight loss results showed that the proportion of 50/50 mixture had the lowest weight loss among all studied molar ratios, followed by 70/30 and followed by 90/10. The mixture of 100/0 came in the last place. Ismail et al. [139] studied the performance of alkali-activated slag/FA geopolymer binders to different forms of sulfate exposure immersed in 5 wt% MgSO4 solution or 5 wt% Na2SO4 solution for 3 months. Sodium metasilicate was used as alkaline activator at concentration of 8 wt%. They reported that MgSO4 was more aggressive to geopolymer paste than Na2SO4. The presence of magnesium led to decalcication of the Ca-rich gel phases presented in the blended slag/FA geopolymer system, causing degradation of the binder system and the precipitation of gypsum. The products of magnesium sulfate attack were poorly cohesive and expansive, led to dimensional instability and loss mechanical performance. On the contrary, immersion of geopolymer pastes in Na2SO4 did not lead to any apparent degradation of the binder and no conversion of the binder phase components into sulfatecontaining precipitates. Chi and Huang [137] studied the percent-

age of water absorption of geopolymer mortars made of different combinations of slag/FA ratios of 100/0, 70/30, 50/50, 30/70, 0/ 100, by weight. Sodium silicate with modulus ratio of 1 was used as alkaline activator. Two concentrations of Na2O of 4% and 6%, by cementitious weight, were employed. The results showed a reduction in the percentage of water absorption with increasing slag and decreasing FA content at both concentrations. Aydn [109] studied the percentage of permeable volume of activated slag mortars with NaOH and sodium silicate. Slag was partially replaced with SF at levels of 0%, 20% and 40%, by weight. The results showed an increase in the percentage of permeable volume by 3.8% and 3.26% with the inclusion of 20% and 40% FA, respectively. As known, shrinkage is the reduction in volume at constant temperature without external loading. It is as important material property that signicant effects on long-term performance of designed structures. It also inuences structural properties and durability of the material. However, Rashad [120] reported that substitutions slag with higher proportions of FA (i.e. 85%, 90%, 95% and 100%) in slag/FA-based geopolymer mortars activated with mixture of NaOH and sodium silicate reduced the drying shrinkage. The drying shrinkage decreased as the content of FA increased. On the same line with this, Chi and Huang [137] prepared

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

41

geopolymer mortars made of different combinations of slag/FA ratios of 100/0, 70/30, 50/50, 30/70, 0/100, by weight. They measure the drying shrinkage at ages of 7, 14 and 28 days. They reported that the drying shrinkage decreased as the content of FA increased. Aydn [109] activated slag mortars with NaOH and sodium silicate. Slag was partially replaced with FA at levels of 0%, 20% and 40%, by weight. The specimens were cured at 20 C and 90% RH for 5 h, then in steam at 70 C for 6 h. After that, the specimens were kept at 20 C and 55% RH. The measurement of drying shrinkage of the specimens started from the end of heat treatment up to 4 months. The results showed a reduction in the drying shrinkage with the inclusion of FA. The drying shrinkage decreased as the content of FA increased. 4.2.5. Fire resistance and metal leaching Guerrieri and Sanjayan [136] studied the high temperature performance up to 800 C, for 1 h, of alkali-activated slag/FA pastes activated with a combination of sodium silicate and 8 M NaOH. Slag was replaced with FA at replacement levels of 0%, 35%, 50% and 100%, by weight. They reported that the specimens with very low initial strength (<7.6 MPa) experienced an increase in residual strength up to 90% gain. Specimens with initial strength in the order of 28 MPa had residual strength losses of approximately 70%, whilst specimens with higher initial strength approaching 83 MPa had residual strength lose in order of 90% after exposure to 800 C. Izquierdo et al. [140] studied leaching of FA/slag/K-silicate/H2O yielded geopolymer bodies. The activator consists of potassium silicate and potassium hydroxide solution with SiO2/K2O = 1.25. The proportions of FA differ between specimens, ranging from 57% to 70%. They reported that geopolymer matrices gave rise to a leaching, which inhibited the leaching of heavy metals with the result that a number of traced pollutants such as Ba, Be, Bi, Cd, Co, Cr, Nb, Ni, Pb, REE, Sr, Th, U, Y and Zr were retained. They also reported that the main mechanism of immobilization in the FA/slag/K-silicate/H2O geopolymers could be essentially physical encapsulation rather than chemical stabilization. Bai et al. [141] used Na2SO4 to activate slag. The slag was partially replaced with pulverized FA at levels of 0%, 10%, 20% and 30%, by weight. The water to solid ratio was xed at 0.35. The pH and corrosion of Al were determined and used as primary criteria for judging the feasibility for development of a Na2SO4 activated slag/PFA matrix for immobilizing nuclear wastes. Leaching studies were carried out to examine the possibility of immobilizing Cs+ within activated slag/PFA matrices. They concluded that replacing slag with PFA slightly reduced the Cs+ immobilization efciency. They also reported that the corrosion of aluminium could be reduced with up to 20% slag replaced with PFA. From the above review of literature of this part, it can be noted that the main advantage of employing FA in AAS matrix is decreasing the setting time and drying shrinkage that AAS system suffers. Workability also increased with the inclusion of FA. The inclusion of FA in AAS matrix reduced the corrosion of aluminium. On the contrary, the inclusion of FA reduced the compressive strength, as reported by some authors. However, the inclusion of FA in AAS matrix is still needs a lot of researches studying the effect of this material on formation of salt eforescences, carbonation resistance and the tendency to cracks during curing. 4.3. Metakaolin 4.3.1. Setting time and heat of hydration Bernal et al. [142] partially replaced slag with MK at levels of 0%, 10% and 20%, by weight, in alkali-activated pastes and mortars activated with a combination of sodium silicate and NaOH. They reported that the inclusion of MK increased total setting time

and reduced the heat release. Cheng and Chiu [143] studied some properties of slag geopolymer blended with metakaolinite. The results indicated that the more metakaolinite added in the system, the slower setting time. Other author [144] reported that MK or MK doped with sodium could be successfully used as heat evolution/hydration accelerating addition in AAS mixtures. Buchwald et al. [145] studied some properties of slag blended with MK-based geopolymers. The composition was 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 alkali-activated 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 concentration of activator. This explained by a higher amount of dissolution of both slag and MK, but it was more signicant on MK dissolution. Table 7 summarizes the pervious researches that studied the effect of MK on the setting time and heat of hydration of AAS system. 4.3.2. Strength Yip et al. [146] replaced slag with MK at levels of 0%, 20%, 40%, 60%, 80% and 100%, by weight, in mortars. The compositions were activated with sodium silicate and NaOH solutions. The activator solutions had two different molar ratios of 2 and 1.2. The binder: sand ratio was 1: 3. The specimens were cured at 40 C for 24 h, then at 25 C up to testing date. The compressive strength results showed that the slag/MK mixtures gave higher strength than neat slag mixture in both molar ratios. When Ms = 2 was used, the presence of 8060 wt% of MK gave an optimum compressive strength. When Ms = 1.2 was used, the presence of 10080 wt% of MK gave an optimum strength. In both cases, the mixtures containing slag/MK ratio of 20/80 gave the highest strength. Bernal et al. [147] studied the compressive strength of alkali-activated slag/ MK pastes. Slag was replaced with MK at levels of 20%, 40%, 60%, 80% and 100%, by weight. The alkaline activating solutions were formulated by blending sodium silicate with NaOH. The alkaline activators were formulated in order to obtain overall (activator + solid precursor) SiO2/Al2O3 molar ratios of 3.0, 3.4, 3.8 and 4.0 and a constant Na2O/SiO2 ratio of 0.25. The results indicated that at 3.0 SiO2/Al2O3 molar ratio, the 60/40 slag/MK gave the highest compressive strength. At the remaining SiO2/Al2O3 molar ratios the 80/20 slag/MK gave the highest compressive strength. Bernal et al. [148] activated slag/MK blended binders with NaOH which mixed with either SF or RHA, as alternative silica-based activator. The activator solutions had Na2O/Al2O3 ratios of 3.0 and 3.8. The ratios of slag/MK were 0/100, 20/80, 40/60, 60/40 and 80/20, by weight. The results showed that the hardened mixture of 60% slag coupled with 40% MK gave the highest compressive strength at Na2O/Al2O3 ratio of 3, at age of 7 days, whilst the hardened mixture of 80% slag coupled with 20% MK gave the highest compressive strength at Na2O/Al2O3 ratio of 3.8, at the same age; which promoted the coexistence of aluminosilicate reaction products along with calcium silicate hydrate gel. Bernal et al. [142] partially replaced slag with MK at levels of 0%, 10% and 20% in alkali-activated pastes and mortars activated

Table 7 Effect of MK on the setting time and heat of hydration of AAS system. Author Bernal et al. [142] Cheng and Chiu [143] -Wczelik [144] Nocun Buchwald et al. [145] Effects Increased total setting time Reduced the heat release The more MK, the slower setting time Accelerated heat of hydration Accelerated the condensation reaction

42

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

with a combination of sodium silicate and NaOH. They reported that the inclusion of MK in AAS matrix decreased the compressive strength. Bernal et al. [149] activated slag/MK mortars with blending sodium silicate and NaOH to reach the overall molar ratios of 1.6, 2.0 or 2.4. The ratios of slag/MK were 100/0, 90/10 and 80/ 20, by weight. The binder/sand ratio was 1:2.74 and water/ (slag + MK + anhydrous activator) of 0.47 was employed. The results showed that the neat slag mixture gave the highest compressive strength, for all Ms ratios. The compressive strength decreased with increasing MK content. Chen et al. [150] manufactured different compositions of slag/MK pastes. The ratios of slag/MK were 50/ 50, 40/60, 30/70, 20/80 and 0/100, by weight. The compositions were activated with NaOH solution. The ratio of liquid to solid was 0.45. The pastes were cured hydrothermally at 90, 120 and 180 C. The results showed a reduction in compressive strength with increasing MK content, at all curing conditions. Buchwald et al. [151] studied the compressive and bending strengths of neat AAS paste and alkali-activated slag/MK pastes. Slag was replace with MK at levels of 0%, 25%, 50% and 100%, by weight. The results showed that as MK content increased as the compressive strength and bending tensile strength decreased. The neat slag gave the highest compressive and bending tensile strengths, whilst neat MK gave the lowest. Wang et al. [152] studied the compressive strength and porosity of alkali-activated slagFA-MK cementitious materials prepared by hydrothermal method. Waterglass was used as alkaline activator. The modulus of waterglass was adjusted to 1.0 by dissolving NaOH. The ratio of water to solid was about 0.35. Different mixtures with different contents of slag, MK and FA were employed. The slag contents ranging from 16.2% to 31.33%, whilst the FA contents ranging from 20.46% to 73.52% and MK contents ranging from 7.22% to 49.39%. The compressive strength results indicated that this type of material had higher mechanical strength. The highest compressive strength value reached about 80 MPa. They suggested that the higher compressive strength was attributed to the addition of slag. The more contents of slag in the system gave more hydration products (CSH and hydrated aluminates calcium). 9 az et al. [153] studied the compressive strength of Burciaga-Di AAS pastes activated with a combination of sodium silicate and NaOH. They used different concentrations of Na2O (i.e. 5%, 10% and 15%). Slag was replaced with MK at levels of 0%, 20%, 50%, 80% and 100%, by weight. The neat slag pastes gave the highest compressive strength at 5% of Na2O, whilst neat MK and slag/MK of 20/80 pastes required 15% Na2O to reach the highest compressive strength. The replacement levels of 20% and 50% with MK required 10% Na2O to reach the highest compressive strength. Bernal et al. [154] partially replaced slag with MK at levels of 0%, 10% and 20%, by weight, in alkali-activated concretes. The alkaline activating solutions were formulated by blending sodium silicate and NaOH to reach the overall molar ratios (SiO2/Al2O3) of 3.6, 4.0 and 4.4. They used different activator concentrations. They reported that at high activator concentration, the inclusion of MK enhanced the compressive and exural strengths at early age. Cheng and Chiu [143] studied slag geopolymers blended with metakaolinite. The results indicated that the compressive strength increased with increasing metakaolinite content, whilst the density decreased with increasing metakaolinite content. Yip et al. [155] blended slag/MK at ratios of 0/100, 20/80 and 40/60, by weight. Commercial sodium silicate solution and NaOH pearl at molar ratios (SiO2/Na2O) of 2.0, 1.5 and 1.2 were used to activate the slag/MK. The compressive strength results showed that slag/MK at ratio of 20/80 gave the highest compressive strength at 2 and 1.2 M ratios, whilst slag/MK = 0/100 gave the highest compressive strength at 1.5 M ratio. Yunsheng et al. [156] tested the mechanical strength of slag/MK-based geopolymer mortars. NaOH and sodium silicate solution with the molar

ratio (SiO2/Na2O) of 3.2 were used as alkaline reagents. The ratios of slag/MK were 0/100, 10/90, 30/70, 50/50 and 70/30, by weight. The specimens were cured at 20 C and 100% RH for 28 days. The results showed that geopolymer mortar containing 50/50 of slag/ MK gave the highest compressive strength, followed by 70/30 and followed by 30/70. The geopolymer mortar containing 100% MK gave the lowest compressive strength. The results of exural strength showed a similar tendency as compressive strength. Table 8 summarizes the pervious researches that studied the effect of MK on the strength of AAS system. 4.3.3. Durability Shen et al. [157] partially replaced slag with zeolites or MK in alkali activation pastes. They reported that replacement slag with zeolites or MK increased the porosity of the hardened pastes, but the leaching fraction of Ca + and Sr2+ were decreased. The decrease in leaching fraction might be attributed to the formation and adsorption properties of (Al + Na) substituted CSH and self generated zeolite precursor. Zhang et al. [158] studied the permeability, measured by Darcy method, of slag/MK-based geopolymers. Different liquid/solid ratios of 0.55 and 0.6 were employed. They reported that the inclusion of slag could reduce permeability, particularly at liquid/solid ratio of 0.6. The existence of slag had only a slight effect on permeability of geopolymer at liquid/solid 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 [146]. Bernal et al. [154] studied water absorption, capillary sorptivity and rapid chloride permeability test (RCPT) of alkali-activated slag/ MK concrete. The ratios of slag/(slag + MK) were 0.8, 0.9 and 1.0. The alkaline activating solutions were formulated by blending sodium silicate and NaOH to reach the overall molar ratios (SiO2/ Al2O3) of 3.6, 4.0 and 4.4. They concluded that the increase in MK contents and higher activator concentrations led, in most cases, to reduce water absorption (Fig. 8) and water sorptivity and gave lower chloride permeability (Fig. 9). Wang et al. [152] studied the porosity of alkali-activated slag-FA-MK cementitious materials prepared by hydrothermal method. Waterglass was used as alkaline activator with the modulus adjusted to 1.0 by dissolving NaOH. The ratio of water to solid was about 0.35. The porosity results indicated that this type of material had compact structure. The porosity was less than 36%, after hydrothermal process. 4.3.4. Carbonation resistance The accelerated carbonation test was used to induce the carbonation of alkali-activated slag/MK mortars activated with blending sodium silicate and NaOH to reach the overall molar ratios of 1.6, 2.0 or 2.4 [149]. The binder/sand ratio was 1:2.74. The water/(slag + MK + anhydrous activator) of 0.47 was employed. The specimens were exposed to an accelerated carbonation with a concentration of 3.0 0.2% at temperature of 20 2 C and RH of 65 5%. The specimens were exposed to these conditions for 340 or 540 h. After 340 h of carbonation at Ms = 2.4, the mortar based on neat slag showed the lowest carbonation depth. The carbonation depth increased with the increase in MK content. After 540 h of carbonation at any Ms, all specimens were fully carbonated. The results also showed that the compressive strength decreased after carbonation. In another investigation [154] the accelerated carbonation test was used to induce the carbonation of alkali-activated slag/MK concretes activated with blending sodium silicate and NaOH to reach the overall molar ratios of 3.6, 4.0 and 4.4 [154]. The ratios of slag/slag + MK were 1, 0.9 and 0.8, by weight, were employed. Carbonation concentration was 3.0 0.2% at temperature of 20 2 C and RH of 65 5%. The specimens were exposed to carbonation for 250, 500, 750 and 1000 h. The results, after carbonation, showed increases in the total pore

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 Table 8 Effect of MK on the strength of AAS system. Author Yip et al. [146] Bernal et al. [147] Bernal et al. [148] Bernal et al. [142] Bernal et al. [149] Chen et al. [150] Buchwald et al. [151] Wang et al. [152] 9 az et al. [153] Burciaga-Di Replacement levels (%) 0, 20, 40, 60, 80 and 100 20, 40, 60, 80 and 100 20, 40, 60 and 100 0, 10 and 20 0, 10 and 20 50, 60, 70, 80 and 100 0, 25, 50 and 100 7.2249.39 0, 20, 50, 80 and 100 Positive effects p 20% optimal p 40% optimal p 20% optimal p 40% optimal p 20% optimal Negative effects Notes SiO2/Al2O3 = 3 SiO2/Al2O3 = 3.4, 3.8 and 4 Na2O/Al2O3 = 3 at age 7 days Na2O/Al2O3 = 3.8 at age 7 days

43

p p p p p p

Bernal et al. [154] Cheng and Chiu [143] Yip et al. [155] Yunsheng et al. [156]

0, 10 and 20 Different contents 60, 80 and 100 30, 50, 70, 90 and 100

p p p p p p p p p p

100% optimal 80% optimal 20% and 50% optimal At early age 80% optimal 100% optimal 50% optimal 30% second place 70% third place

Na2O = 5% Na2O = 15% Na2O = 15% Na2O = 10% At higher activator concentration Increased with increasing MK content Molar = 2 and 1.2 Molar = 1.5

Fig. 8. Water absorption in the rst 48 h of alkali-activated slag/MK concretes with 28 days of curing [154].

4.3.5. Fire resistance Bernal et al. [147] studied the effect of elevated temperatures of 200, 400, 600, 800 and 1000 C, for 2 h, on the geopolymers formulated with an overall SiO2/Al2O3 molar ratio of 3, slag/(slag + MK) ratios of 0.0 and 0.2 (i.e. slag/MK ratios of 0/100 and 20/80). Constant H2O/Na2O ratio of 12 and Na2O/SiO2 ratio of 0.25 were employer. The results indicated that the geopolymers formulated with MK and slag had higher residual compressive strength than the neat MK-based geopolymer up to 800 C. On the other hand, the neat 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 [143] studied re resistance of slag geopolymer blended with metakaolinite, when a 10 mm thick panel of geopolymer 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. Table 9 summarizes the previous researches that studied the effect of MK on durability, carbonation resistance and re resistance of AAS system. 4.3.6. Immobilization Guangren et al. [159] evaluated the effects of MK on simulated radioactive Sr or Cs immobilizing behaviour of AAS matrix by cation exchange capacity, distribution ratio of selective adsorption, leaching test and porosity analyses. PC matrix was used as a reference. The alkaline activators were waterglass and sodium carbonate. The addition of alkaline activator was adjusted to 5%, by Na2O, of slag in weight. Slag was partially replaced with MK at levels of 0%, 5%, 10% and 15%, by weight. The pastes containing 0.5 wt% Cs+ (CsCl) or Sr2+ (SrCl2.6H2O) were cast in a cylindrical mould, with a diameter of 2.5 cm and a height of 5 cm which was cured in fog room at 25 C for 28 days. Ratio of water to solid in the pastes was 0.3. The effects of MK on Sr or Cs ion selective adsorption of AAS matrix were evaluated by the distribution ratio (Kd) which dened as the ratio of the amounts of ions adsorbed by a unit mass of solid adsorbent to the equilibrium concentration of ion in the aqueous phase. The results of the Kd showed that PC had the lowest Kd, whilst Kd of AAS matrix was highly increased by 38% and 151% for Sr ion and Cs ion, respectively. This indicated that the improvement of AAS matrix on selective adsorption

Fig. 9. Rapid chloride permeability test results for 28 days cured activated slag/MK [154].

volumes of the specimens and decreases in the compressive strengths with the increase in MK content.

44

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

Table 9 Effect of MK on the durability, carbonation resistance and re resistance of AAS system. Author Shen et al. [157] Zhang et al. [158] Bernal et al. [154] Reduced water absorption and water sorptivity Lowered chloride permeability Reduced the porosity Increased carbonation depth After carbonation Increase in the total pore volume Decrease in the compressive strength Neat MK-based geopolymer gave higher residual compressive strength at 1000 C than slag/MK-based geopolymer 20/80 slag/MK-based geopolymer gave higher residual compressive strength than the neat MK-based geopolymer after exposure to temperatures up to 800 C Fire resistance improved with increasing MK content Positive effects Decreased leaching fraction of Ca+ and Sr2+ No or negative effects Increased the porosity Increased the permeability With higher activator concentration Notes

Wang et al. [152] Bernal [149] Bernal et al. [154]

The degradation increased as the content of MK increased

Bernal et al. [147]

Cheng and Chiu [143]

behaviour for Cs ion was much more effective than for Sr ion. Additions of MK into AAS matrix had positive effect to further enhance their selective adsorption on Sr and Cs ions. As MK content increased in AAS matrix, as the Kd increased. For slag/MK ratio of 85/15, its Kd was larger than that of neat AAS matrix by about 40% and 56% for Sr ion and Cs ion, respectively. The results of leachability showed that the leaching rate for Sr or Cs ion decreased in the order PC > neat AAS > alkali-activated slag/MK. The porosity results showed that PC matrix had the largest porosity. Neat AAS matrix and alkali-activated slag/MK matrix possessed similar pore structure. Their pore size distributions were concentrated on small pores (<10 mm), which represented over 80% of the porosity. Yunsheng et al. [156] used slag/MK ratio of 50/50 mortar activated with NaOH and sodium silicate solution, with SiO2/Na2O ratio of 3.2. The specimens were cured at 80 C for 8 h to study the immobilization behaviour of slag/MK-based geopolymer in the presence of Pb and Cu. The leaching tests showed that slag/MKbased geopolymer could effectively immobilized Cu and Pb heavy metal and the immobilization efciency exceeded 98.5% when the amount of heavy metals containing 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. From the above review of literature of this part, it can be noted that the main advantage of employing MK in AAS matrix is decreasing the setting time that the AAS system suffers as well as decreasing the leaching formation of Ca+ and Sr2+. The inclusion of MK in AAS matrix, in most cases, reduced the porosity, reduced chloride permeability, reduced water absorption and water sorptivity, increased the compressive strength and improved the resistance of re. On the contrary, the inclusion of MK in AAS matrix increased the carbonation depth. However, the inclusion of MK in AAS matrix is still needs more investigations that studying the effect of this material on drying shrinkage, formation of salt eforescences, and the tendency to cracks during curing. 4.4. Portland cement 4.4.1. Setting time Wu et al. [160] reported that the setting times (initial and nal) of activated 30%slag/70%PC with Na2SO4 was almost same as that

of PC. Fu-Sheng et al. [161] partially replaced slag with PC at levels of 0%, 10%, 20% and 30%, by weight. The compositions were activated with the same concentration of Na2SiO3. They reported that the mixture which containing 30% PC exhibited rapid setting. 4.4.2. Strength Roy et al. [162] studied the compressive strength of normal Portland slag cement (NPSC), which containing slag levels ranging from 30% to 80%, in comparison with alkali-activated Portland slag cement (PSC). The PSC was activated with 1% waterglass. The alkali-activated PSC pastes showed higher early compressive strength than that of NPSC. At age of 28 days, the NPSC pastes showed higher compressive strength than alkali-activated PSC for up to 60% slag substitution. Beyond 60% substitution, alkali-activated PSC showed higher compressive strength compared to NPSC. Wu et al. [160] activated 30%slag/70%PC with Na2SO4. The early ages (3 and 7 days) compressive strengths of the activated slag/PC were lower than that of traditional PC pastes. On the contrary, at ages of 28 and 90 days, the compressive strengths of activated slag/PC surpassed that of traditional PC pastes. Roy et al. [46] studied the compressive strength of 60/40 slag/PC without or with alkali activation. The alkaline activator was 2 M NaOH. They reported that the activated slag/PC gave higher strength than that of slag/PC without activation. Using 2 M NaOH was sufcient to achieve strength equal to that of neat PC, at age of 2 days, whilst little difference was found beyond 28 days (Fig. 10). Talling and Brandstetr [119] reported that in AAS activated with sodium silicate, the addition of PC might cause harmful to strength where PC containing gypsum, which neutralizes the alkaline component and produced expansive non-binding mirabilite (Na2SO4.10H2O). On the same line with this, Bilim and Atis [163] studied the compressive strength and exural strength of AAS mortars activated with sodium silicate. Three different Na dosages of activator were used. The slag was partially replaced with PC at levels of 0%, 20%, 40% and 80%, by weight. 100% PC mortar without alkali activator was also employed. They reported that the compressive and exural strengths decreased as PC content in activated slag increased. The maximum compressive strength and exural strength was obtained from AAS mortar without any replacement of PC. Bakharev et al. [38] activated slag with sodium silicate to manufacture AAS pastes. The concentration of sodium silicate was 4% Na with Ms 0.75. The slag was partially replaced

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

45

ites of slag/PC, in mortars, with waterglass at 010 wt% of Na2O. The ratios of slag/PC were 100/0, 80/20, 50/50, 30/70 and 0/100, by weight. The compressive strength results showed that the composite of 0/100 gave the highest strength at 0% Na2O, followed by 30/70 and followed by 50/50. At each of 4%, 6% or 10% Na2O, the composite of 80/20 exhibited the highest compressive strength, followed by 50/50 and followed by 30/70. Singh et al. [166] reported that the use of 4% Na2SO4 signicantly increased the strength of slag with 50% PC. Table 10 summarizes the pervious researches that studied the effect of PC on the strength of AAS system. 4.4.3. Durability Roy et al. [46] studied porosity, median pore size and density of 65/35 slag/PC without or with alkali activation. The alkaline activator was 4 M NaOH. They reported that the alkaline activator reduced the porosity and median pore size, whilst density was increased. The porosity, median pore size and the density of activated slag/PC were higher than that of neat AAS, whilst the porosity and median pore size of activated slag/PC were lower than that of neat PC without alkaline activator. On the same line with this, ivica [114] activated the composites of slag/PC mortars with SF Z activator. The ratios of slag/PC were 100/0, 90/10 and 70/30, by weight. The results of total porosity and pore median decreased as the content of PC increased. Veiga and Gastaldini [167] partially replaced slag, in mortars, with either 50% white PC or grey PC, by weight. The composites were non-activated or activated with 4% Na2SO4, by total binder weight. Also, neat white PC and neat grey PC without activator were employed as references. The hardened specimens were exposed to 5% Na2SO4 for 2 years according to ASTM C1012/04. The expansion results, after 2 years, indicated that the composition of 50% slag/50% white PC activated with Na2SO4 or without activation presented expansion reductions of 95.2% and 94.9%, respectively, in comparison with neat white PC. The composition of 50% slag/50% grey PC activated with Na2SO4 or without activation presented expansion reductions of 96.9% and 95.7%, respectively, in compression with neat grey PC. These mean that the activated 50% slag blend had higher capability to resist sodium ivica [113] sulfate attack and classied as high sulfate resistance. Z studied the acidic resistance of slag/PC activated with chemically MSF. The ratios of slag/PC were 100/0, 90/10 and 70/30, by weight. The results showed that as the cement content increased as the acidic resistance decreased. Roy et al. [46] investigated the steady state chloride diffusion through the hardened pastes. The pastes were manufactured from slag/PC at different levels ranging from 0/100 to 100/0. The different mixtures of slag/PC were non-activated or activated with alkaline activator. The results showed a clear trend of decreasing diffusion rate with decreasing PC content (Fig. 11). In addition, the activated hardened pastes showed lower diffusion rate than that of un-activated hardened pastes. Table 11 summarizes the previous researches that studied the effect of PC on the durability of AAS system. 4.4.4. Drying shrinkage Fu-Sheng et al. [161] studied the drying shrinkage of blended alkali-activated slag/PC mortars activated with Na2SiO3. Slag was replaced with PC at levels of 0%, 10%, 20% and 30%, by weight. PC mortar mixture was employed as a reference. The results showed that the neat AAS specimen exhibited higher drying shrinkage at age of 35 days, whilst the blend of 10% PC specimen gave lower drying shrinkage which similar to the drying shrinkage of PC specimen. From the above review of literature of this part, it can be noted that the main advantage of employing PC in AAS matrix is decreasing the drying shrinkage that the AAS system suffers as well as

Fig. 10. Strength development of different types of mortars [46].

with PC at level of 30%. The compressive strength results at 1, 3 and 28 days were conducted. The results revealed that the neat AAS paste gave higher strength than 70/30 slag/PC paste at all ages. Indeed, the results of Douglas and Brandstetr [96] contrast with the above results. However, they activated slag mortars with sodium silicate. The mortars were modied with some additives as PC. In some mortar mixtures, 2% PC and 8% PC were added in the AAS mortars. The w/b ratios were 0.38 and 0.48. The activator concentration was 4.6 g Na2O/100 g binder. For specimens prepared with w/b ratio of 0.38, the results showed that the mortars containing 2% PC gave higher compressive strengths than that containing 8% PC at ages of 3, 7 and 28 days. At 0.48 w/b ratio, the compressive strength value of 8% PC was higher than that of 2% PC, at ages of 3 and 7 days. On the other hand, the compressive strength of 2% PC was higher than that of 8% PC at age of 28 days. ivica [164] activated different compoOn the same line with this, Z sitions of slag/PC with alkali-silicate admixture. The ratios of slag/ PC were 100/0, 90/10 and 70/30, by weight. The results showed that replacing slag with 10% PC gave the highest compressive strength at ages of 28 and 90 days. Replacing slag with 30% PC came in the second place. Fu-Sheng et al. [161] studied the compressive strength of blended alkali-activated slag/PC activated with Na2SiO3. Slag was partially replaced with PC at levels of 0%, 10%, 20% and 30%, by weight. Neat PC mixture was employed for comparison. The results showed that partially replacing slag with 10% PC gave strength higher than neat PC or those of 20% or 30% replacements. Also, partially replacement slag with 10% PC gave higher strength than that of neat AAS at ages of 1 and 28 days, whilst it gave lower strength at age of 3 days, compared to neat AAS. ivica [115] partially replaced slag, in mortars, with PC at levels Z of 10% and 30%, by weight. The composites were non-activated or activated with SF activator. The compressive strength results showed that the activated slag/PC at ratio of 90/10 gave higher strength than that of activated 70/30 at age of 28 days, whilst the activated 70/30 gave higher strength than that of activated 90/10 at ages of 1 and 90 days. Both the activated composites gave higher strength than either un-activated composites or traditional neat PC ivica [114] activated different mortars. In another investigation Z composites of slag/PC mortars with SF activator. The ratios of slag/PC were 100/0, 90/10 and 70/30, by weight. The results showed that 70/30 composite gave the highest compressive strength at age of 28 days, followed by 90/10 and followed by 100/0. Acevedo-Martinez et al. [165] activated different compos-

46 Table 10 Effect of PC on the strength of AAS system. Author Wu et al. [160] Roy et al. [46] Talling and Brandstetr [119] Bilim and Atis [163] Bakharev et al. [38] Douglas and Brandstetr [96]

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

Replacement levels (%) 70 40 Undetermined 0, 20, 40 and 80 0 and 30 2 and 8

Positive effects

Negative effects p

Notes At 3 and 7 days, compared to PC At 28 and 90 days, compared to PC Compared to slag/PC of 60/40 without alkali activation Compressive and exural strengths decreased as PC content increased. 0% gave the highest strength 0% gave the highest strength w/b = 0.38 at ages of 3, 7 and 28 days w/b = 0.48 At ages of 3 and 7 days w/b = 0.48 At age of 28 days

p p p p p 2% higher than 8% p 8% higher than 2% p 2% higher than 8% p 10% optimal p 30% second place p 10% optimal p 10% optimal p 30% optimal p 30% optimal p 10% second place p 0% third place p 100% optimal p 70% second place p 50% third place p 20% optimal p 50 % second place p 70% third place p p

ivica [164] Z

0, 10 and 30

Fu-Sheng et al. [161] ivica [115] Z ivica [114] Z

0, 10, 20 and 30 10 and 30 0, 10 and 30

At At At At

ages of 1 and 28 days age of 3 days, 0% gave the highest strength age of 28 days ages 1 and 90 days

Acevedo-Martinez et al. [165]

0, 20, 50, 70 and 100

At Na2O = 0% At Na2O = 0% At Na2O = 0% At Na2O = 4%, 6% and 10% At Na2O = 4%, 6% and 10% At Na2O = 4%, 6% and 10% With 4% Na2SO4

Singh et al. [166]

50

mation of salt eforescences and the tendency to cracks during curing. 4.5. Lime In AAS system, there is a growing interest in the development of calcium hydroxide-activated slag binder due to its low cost and good durability [37]. However, lime or lime slurry has been used to improve the early strength [119,168] and control the setting time of AAS cement using sodium silicate with Ms = 2.85 [169]. Wang et al. [97] reported that lime and lime slurry could be used in AAS to improve the early strength (less than 7 days), whilst the later strength might be slightly reduced, in many cases. They also reported that the addition of 510% lime resulted in a dramatic increase in strength, when Na2SO4 was used as main activator. Shi and Day [125] studied the effect of two types of FA and the addition of lime on the strength development and hydration of slag/FA mixtures activated with NaOH and sodium silicate. They concluded that the addition of a small amount of hydrated lime signicantly increased the early age strength, but slightly decreased the later age strength of the cements. Cheng et al. [170] activated slag with waterglass solution with Ms = 1.5. They reported that both early and 28 days strengths increased dramatically by adding 1.93.4% lime. Collins and Sanjayan [50] investigated some properties of AAS concrete mixtures, activated with different activators, in comparison with PC concrete. Two types of activators were used: NaOH in combination with sodium carbonate and sodium silicate in combination with hydrated

Fig. 11. The effective diffusion coefcient of activated and un-activated slag/PC hardened pastes [46].

decreasing the total porosity and pore median. The inclusion of PC in AAS matrix, in most cases, increased the compressive strength. On the same line with this, the inclusion of PC in AAS matrix had higher capacity to resistance sodium sulfate attack and increased the acidic resistance. On the contrary, the inclusion of PC in AAS matrix increased the chloride diffusion rate as reported by Roy et al. [46]. However, the inclusion of PC in AAS matrix is still needs more investigations that studying the effect of this material on for-

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 Table 11 Effect of PC on the durability of AAS system. Author Roy et al. [46] ivica [114] Z Veiga and Gastaldini [167] ivica [113] Z Roy et al. [46] % Incorporation 35 0, 10 and 30 50 0, 10 and 30 0100 Positive effects Lowed porosity and median pore size Decreased the total porosity and pore median as PC content increased Higher capability to resist sodium sulfate attack and classied as high sulfate resistance Increased in the acidic resistance as the PC content increased Increased chloride diffusion rate with increasing PC content No or negative effects

47

lime. The sodium silicate was either liquid or powder. The paste that activated with powdered sodium silicate and lime slurry demonstrated considerably better workability and higher compressive and exural strengths than the other concrete mixtures. On the other hand, concrete activated with sodium silicate and lime slurry showed greater drying shrinkage than other concrete mixtures. Yang et al. [171] studied the compressive strength development of AAS pasts and mortars activated with different activators. 7.5% calcium hydroxide was used for the main activator and either 1% Na2SiO3 or 2% Na2CO3 was added for an auxiliary activator. The compressive strength of the pastes activated with calcium hydroxide combined with Na2SiO3 gave the highest compressive strength, whilst pastes activated with a combination of calcium hydroxide and Na2CO3 came in the second place. Finally, pastes activated with individual calcium hydroxide came in the last place. The activated mortars with the combination of calcium hydroxide and Na2SiO3 gave compressive strength higher than that activated with the combination of calcium hydroxide and Na2CO3. Yang and Song [172] studied some mechanical properties of calcium hydroxide based AAS concrete. As an alkali activator, S-type for a combination of 7.5% calcium hydroxide and 1% Na2SiO3 and C-type for a combination of 7.5% calcium hydroxide and 2% Na2CO3. They concluded that the 28-day compressive strength of calcium hydroxide based AAS concrete showed higher compressive strength developed in calcium hydroxide and Na2SiO3-activated slag concrete than that in calcium hydroxide and Na2CO3-activated concrete. Yang et al. [173] presented some properties of calcium hydroxide-slag concrete activated with different auxiliary activators with different w/b ratios. They reported that calcium hydroxide-slag concrete displayed an enhanced workability and delayed setting time. Compressive strength increased with the decrease in w/b ratio. Table 12 summarizes the previous researches that studied the effect of lime on some properties of AAS system.

5. Other materials as additives Li and Liu [174] utilized 4% slag, by weight, as additive to 90% FA + 10% MK based geopolymer. Na2SiO3 and NaOH solution was used as alkaline activator. The paste specimens were cured at 30 or 70 C for 14 days. After 14 days, the specimens were tested in compression. The results showed 51.96% and 25.72% increases in compressive strengths, due to the inclusion of 4% slag, cured at 30 and 70 C, respectively. 4% slag addition also inuenced the pore structure of the geopolymer, signicantly. A rened pore size and reduced porosity volume were exhibited after 4% slag addition. Zhang et al. [175] blended slag and MK in the mass ratio of 4:1. Sodium metasilicate, Na2SiO3.9H2O was used as alkaline activator. The slag/MK based geopolymer reinforced with organic resins (OR) that 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. Zhang et al. [176] modied the composite of 80% slag/20% MK, which activated with sodium metasilicate, by different contents of resin powder of ethylenevinyl acetate copolymer (1%, 5%, 10% and 15 wt%). The results showed that 1% resin improved the compressive strength and exural strength by 8.64% and 23.91%, respectively, at age of 3 days; 30.51% and 21.82%, at age of 7 days; and 11.15% and 40.98%, respectively, at age of 28 days. These improvements in strengths were ascribed to: (1) the microcracks on the surface of geopolymer were signicantly ameliorated; (2) incorporation of resin was able to prevent the cracking growth through lling effect; (3) the doping resin had an effect of immobilizing water molecules and effectively postpone the water evaporation. Zhang et al. [177] modied the alkali-activated slag/MK based geopolymer with either 1% or 5% of the polyvinyl acetate (PVA) resin. The mixture of slag and MK/activator/organic resin/water was 1:0.15:0.010.05:0.36. The ratio of slag/MK was 80/20. The specimens were cured at 20 C and 99% RH for one day, then at room temperature for additional 27 days. After curing, the specimens were exposed to elevated temperatures of 150, 300, 450, 600 and 850 C for 1 h. The compressive and exural strengths results at age of 28 days showed that the composite with 1% resin gave the highest strength, followed by the geopolymer without any resin, whilst the composite with 5% resin gave the lowest strength. After exposure to elevated temperatures, the compressive strength progressively increased with the increase in the temperatures from 150 to 300 C due to promotion of polycondensation reaction. The compressive and exural strengths signicantly decreased in the temperature range of 450850 C resulting from the dehydration of geopolymer matrix, thermal decomposition of resin and phase transformation. The heat resistance of the geopolymer could be improved by incorporating content of 1 wt% resin. The geopolymer and the composites were composed of amorphous phase and some crystalline phases such as akermanite, gehlenite and nonstoichiometric sodium aluminum silicate after the specimen was bared to 850 C for 1 h. However, the composite with 5% resin showed the lowest compressive strength after exposure to elevated temperatures (Fig. 12). Sanyin et al. [178] studied the effect of replacing different levels of carbonatite with slag, in alkali-activated mortars. Carbonatite was partially replaced with slag at levels of 25%, 30%, 35% and 40% in pastes and 30%, 40% and 50% in mortars, by weight. The inclusion of barium chloride in the system was investigated. Sodium silicate solution was prepared by dissolving the sodium silicate and industrially pure NaOH in tap water. Barium chloride was weighted by the mass percentage of slag and dissolved in tap water. The ratio of mixing solution (sodium silicate solution + barium chloride solution or water) to slag, by weight, was 0.5. The setting time results, in the case without barium chloride addition,

48 Table 12 Effect of lime on some properties of AAS system. Author Pu et al. [119] Talling and Brandstetr [168] Douglas et al. [169] Wang et al. [97] Positive effects

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

No or negative effects

Notes

Increased the early age compressive strength Increased the early age compressive strength Controlled the setting time Increased the early age strength

In many cases, slightly decreased the later age strength

Shi and Day [125] Cheng et al. [170] Collins and Sanjayan [50] Yang et al. [171]

The addition 510% lime, increased the compressive strength when Na2SO4 was used as main activator Increased the early age strength Increased both early and 28 days strengths Improved the workability Increased the strength Gave the highest compressive strength Compressive strength came in the second place Compressive strength came in the third place

Slightly decreased the later age strength Activator: waterglass with Ms = 1.5 Activator: sodium silicate and lime slurry Activator: 7.5% lime + 1% Na2SiO3 Activator: 7.5% lime + 2% Na2CO3 Activator: 7.5% lime Activator: 7.5% lime + 1% Na2SiO3 Activator: 7.5% lime + 2% Na2CO3

Increased the drying shrinkage

Yang and Song [172]

Gave the highest compressive strength Compressive strength came in the second place

Yang et al. [173]

Improved the workability Delayed the setting time

Fig. 12. Effects of elevated temperature on compressive strength of geopolymer without or with either 1% or 5% PVA resin [177].

Fig. 13. Effect of Gypsum on the drying shrinkage [185].

showed that the more the slag replacement, the shorter the setting time. Barium chloride had a good reduction effect on the setting of alkali-activated carbonatite with slag replacement; setting time increased with the amount of barium chloride added. The compressive strength increased with increasing slag content. The inclusion of barium chloride reduced the compressive strength. This might be attributed to the formation of a covering layer on the surface of slag grain. Fang et al. [179] studied the effect of the inclusion of burnt magnesia (0%, 2%, 4%, 6% and 8 wt%) on some properties of AAS mortar. Setting time, compressive strength, exural strength and drying shrinkage were studied. Magnesia was burnt at different temperatures of 800, 850, 900 and 950 C for 2 h. Waterglass with SiO2/Na2O modulus of 1.42 and concentration of 5 wt% was used as alkaline activator. The specimens were cured at 20 1 C and 95 2% RH. The results showed that as magnesia content

increased, as the initial setting time decreased. Increasing burnet temperature of magnesia led to increasing setting time. Both compressive and exural strengths decreased with the increase of magnesia content. Increasing burnt temperature led to increasing compressive and exural strengths. Specimens with magnesia exhibited lower drying shrinkage compared to AAS mortar specimen without magnesia. The drying shrinkage decreased as the content of magnesia increased. Drying shrinkage increased as the burnt temperature of magnesia increased. They concluded that magnesia burnt at 850950 C could be used to reduce the shrinkage of AAS mortar when its additive amount did not exceed 8%; otherwise, the setting time might be too short and compressive and exural strengths might severely decrease. Brough et al. [180] studied some properties of AAS mortars admixed with sodium chloride (NaCl) and malic acid. 1.5 M of Na2Si2O3 was used as alkaline activator. The slag/sand ratio was 1/2.33. The additions of malic acid or NaCl were dissolved in the activator

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

49

solution prior mixing and the concentrations were given as percentage by mass with respect to slag mass. Compressive strength measurement indicated that the addition of 8% NaCl to the mixture signicantly retarded the strength development. The reductions in compressive strengths were 100%, 94.74%, 93.75%, 13.79% and 0% at ages of 1, 3, 7, 28 days and 3 months, respectively. The setting time results showed that 8% NaCl acted as retarder. On the other hand, small additions of NaCl, 1% and 4%, acted as accelerator. The addition of 0.5% malic acid signicantly retarded the initial and nal setting times. The retarding effect of malic acid was much higher than that of 8% NaCl. The combination of 0.5% malic acid and 8% NaCl gave the highest retarder during this study.

ek et al. [181] activated three different types of steel slag Vlc with a solution of NaOH and waterglass. Steel slag was partially replaced with either 50% slag or 25% slag and 25% FA, by weight. They concluded that the studied materials, both dense and lightweight, displayed good product manufacturing qualities while their preparation was economically acceptable as well as being environmentally friendly. El-Didamony et al. [182] partially replaced slag with air cooled slag (ACS) in alkali activation system. Slag was partially replaced with ACS at levels of 0%, 20%, 40% and 60%, by weight. The slag/ACS specimens activated with 3.3 NaOH: Na2SiO3 wt% at optimum value of 6 wt% mixed with sea water were cured at room temperature and 100% RH. The results showed that the

Table 13 Effect of different additives on some properties of AAS system. Author Li and Liu [174] Zhang et al. [175] Additives 4% slag added to 90% FA + 10% MK Organic resins (acrylic resin) with slag/MK Positive effects Reduced the porosity and rened pore size improved the compressive and exural strengths improved the fracture toughness Improved the compressive and exural strengths Improved the compressive and exural strengths Improved the heat resistance Decreased the compressive and exural strengths Decreased the heat resistance Reduced the setting time Reduced the compressive strength Reduced the compressive strength Decreased the initial setting time Decreased the compressive and exural strengths decreased the compressive strength development Accelerated the setting time (1%, 4%) No or negative effects

Zhang et al. [176] Zhang et al. [177]

Resin powder of ethylenevinyl acetate copolymer with slag/MK 1% polyvinyl acetate (PVA) resin with slag/MK

5% polyvinyl acetate (PVA) resin with slag/MK

Sanyin et al. [178]

Carbonatite blended with slag Carbonatite with slag and barium chloride Burnt magnesia Increased the setting time Decreased the drying shrinkage

Fang et al. [179]

Brough et al. [180]

NaCl

Retarded the setting time (8%)

Malic acid ek et al. [181] Vlc El-Didamony et al. [182] Allahverdi et al. [183] Kani et al. [184] Steel slag/slag Steel slag/slag/FA ACS 5% slag with 95% pumice-type natural pozzolan Slag with pumice-type natural pozzolan

Retarded the initial and nal setting times Dense and lightweight Shortened the setting time Decreased the compressive strength Decreased the compressive strength

Bakharev et al. [102] Chang et al. [185]

Gypsum Gypsum

Phosphoric acid

No eforescences Increased the compressive strength Slightly decreased the eforescences Reduced the drying shrinkage Increased the compressive strength Reduced the drying shrinkage Increased setting time

Reduced the setting time

Khater [187] Yongde and Yao [188] Brylicki et al. [189]

25% CKD FA, ination agent, zeolite, kieselgur Natural clinoptilolite zeolite (5% soda as activator) Natural clinoptilolite zeolite (5% soda coupled with 5% PC as activator) Aluminum powder/lauryl sulfate = 1 Ground synthetic quartz sand with surface area 2000 cm2/g Ground synthetic quartz sand with surface area 5000 cm2/g

Improved the compressive strength development Increased the compressive strength

Reduced the compressive strength at early age Increased the drying shrinkage Decreased the compressive strength Crack tendency Decreased the exural strength Decreased the exural strength Decreased the compressive and exural strengths

Esmaily and Nuranian [190] Rakhimova and Rakhimov [191]

Ground synthetic quartz sand with surface area 8000 cm2/g

The best compressive strength 3060% increased the setting time 3060% increased the setting time 1030% increased the compressive strength 3060% increased the setting time 1040% increased the compressive strength

1030%, no effect on the setting time Decreased compressive strength 1030%, no effect on the setting time Beyond 3060% decreased the compressive strength 1030%, no effect on the setting time Beyond 40% to 60% decreased the compressive strength

50 Table 14 Problems of AAS system and their limitations. Defect Rapid setting

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

Additive required to limit the defect Superplasticizer (polycarboxylate-based) retarded the setting time (at 5% Na2O of waterglass) Superplasticizer (vinyl copolymer-based) lengthened the nal setting time (at 5% Na2O of waterglass) Superplasticizer (melamine-based) retarded the setting time (at 5% Na2O of waterglass) Superplasticizer (naphthalene-based) retarded the setting time (NaOH activator) Superplasticizer set retarder (based on modied polymer) retarded the setting time (at 4% Na2O, Ms = 0.75, sodium silicate) FA and MK retarded the setting time Calcium hydroxide delayed the setting time Phosphoric acid increased the setting time Barium chloride in slag/carbonatite increased the setting time 8% NaCl retarded the setting time 0.5% malic acid retarded the setting time Ground synthetic quartz sand activated with soda increased the setting time Steel bers reduced the drying shrinkage Polypropylene bers reduced the drying shrinkage Carbon bers reduced the drying shrinkage Glass bers reduced the drying shrinkage Water-reducing (based on lignosulphonates) reduced the drying shrinkage Air-entraining (with a soluble salt of an alkyl aryl sulphonate) reduced the drying shrinkage Shrinkage-reducing (based on polypropylenglycol) reduced the drying shrinkage SF, FA and PC reduced the drying shrinkage Up to 8% burnet magnesia reduced the drying shrinkage 6% gypsum signicantly reduced the drying shrinkage

Higher shrinkage values

Higher formation of salt eforescences Higher carbonation Tendency to cracks during curing

(This needs studies) (This needs more studies) Shrinkage-reducing (based on polypropylenglycol) somewhat decreased the carbonation depth HT-carbon, commercial E-glass, PVA and PVC bers reduced cracks 1% resin powder of ethylenevinyl acetate copolymer in slag/MK-based geopolymer prevented the growing of cracks 1% organic resins, which consist of resin emulsion and polyvinyl acetate resin in slag/MK-based geopolymer prevented the growing of cracks

setting times of 100/0 were shortened with the inclusion of ACS. The blend of 80/20 gave higher compressive strength and bulk density than those of 60/40 and 40/60, but lower than 100/0 up to 90 days. Allahverdi et al. [183] studied the effect of slag on some properties of alkali-activated pumice-type natural pozzolan. Pumice natural pozzolan was partially replaced with slag at levels of 0%, 5%, 15% and 25%, by weight. Waterglass and NaOH solution with silica modulus of 0.6 was used as alkaline activator. The Na2O of the activator had three different concentrations of 4%, 6% and 8%. The w/b ratio was adjusted at four different values of 0.26, 0.28, 0.4 and 0.32. The specimens were cured at 25 C and 40% RH for 24 h, then in ambient conditions until testing date. They concluded that incorporation of the slag did not provide any improvement in setting time and compressive strength behaviours of the studied alkali-activated natural pozzolan. The pastes comprising of 5 wt% slag activated with 8% Na2O and produced at w/b ratio of 0.3 were sound with no eforescences and exhibited the highest 28-day compressive strength along with almost acceptable initial and nal setting times. On the other hand, Kani et al. [184] addressed methods to reduce eforescence in a geopolymer binder based on a pumice-type natural pozzolanic material. The pumice was partially replaced with slag at levels of 2%, 4%, 6% and 8%, by weight, aiming to reduce the eforescence. They concluded that the additional of slag had very minor decrease in the tendency towards eforescence and improved the compressive strength. Compressive strength increased with the increase in slag content. Bakharev et al. [102] studied the drying shrinkage of AAS concrete admixed with 6% gypsum, by weight. The results showed that using gypsum as additive signicantly reduced the drying shrinkage of AAS concrete. Chang et al. [185] examined the effects of gypsum and/or phosphoric acid on the compressive strength development, setting time and drying shrinkage of AAS pastes. A

sodium silicate-based alkaline activator was used. Slag was partially replaced with gypsum at levels of 0%, 2%, 3%, 4% and 6%, by weight. A dosage of 0.85 M phosphoric acid was used alone or combined to gypsum in some mixtures. For the drying shrinkage and compressive strength, the amounts of gypsum used were 0, 2% and 4%. They concluded that using phosphoric acid alone increased the setting time, reduced the compressive strength at early age and increased the drying shrinkage. Adding gypsum alone reduced the setting time, increased the compressive strength and decreased the drying shrinkage (Fig. 13). The combined use of phosphoric acid and gypsum blocked the retarding effect of phosphoric acid, the compressive strength development was similar to that using phosphoric acid alone and the drying shrinkage reducing effect of gypsum disappeared. Dondxu et al. [186] studied two mixtures containing clinker, anhydride and slag activated with Na2SO4. The rst mixture containing 22% clinker, 70% slag and 4% anhydride, activated with 4% Na2SO4. Whilst the second mixture containing 19% clinker, 70% slag and 5% anhydride, activated with 6% Na2SO4. The second mixture showed higher compressive strength than the rst one. Khater [187] partially replaced slag with 25% cement kiln dust (CKD), by weight, in activated pastes. 2% of NaOH was used as alkaline activator. The results showed a signicant decrease in compressive strength with replacing slag with 25% CKD. Yongde and Yao [188] studied compressive strength, exural strength and crack tendency of AAS pastes admixed with various kinds of mineral powders. The pastes were activated with Na2CO3 and NaOH and its combination. The additives were FA, ination agent, composite ination agent, zeolite, uorgypsum and kieselgur. The additives were either alone or combined. They reported that some compressive strengths at 28 days were even higher than 60 MPa. FA, ination agent, zeolite, kieselgur benet the strength development, but a severe problem in crack tendency that caused

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

51

decrease in exural strength was obtained. Brylicki et al. [189] studied compressive and exural strengths of activated slag modied with natural clinoptilolite zeolite (Ca,K2,Na2,Mg)4Al8Si40O9624H2O). They added clinoptilolite to the slag at levels of 0%, 1%, 3% and 5%, by weight. 5% soda (Na2CO310H2O) or 5% soda coupled with 5% PC was used as activator. The results indicated that the compressive strength values increased with increasing clinoptilolite content, when soda was used as activator. On the contrary, the compressive strength values decreased with increasing clinoptilolite content, when blended activator was used. The exural strength values decreased with increasing clinoptilolite contend, when soda or soda coupled with PC was used as activator. Esmaily and Nuranian [190] used AAS in the production of autoclave aerated concrete (AAC). Different amounts of aluminum powder (Al) and sodium lauryl sulfate (SLS) were employed in AAS. The specimens were cured at different temperatures of 70, 78 and 87 C. The best compressive strength was obtained when SLS/Al = 1. Also, the temperature of 87 C led to the best compressive strength. Rakhimova and Rakhimov [191] studied the setting time and the compressive strength of the composite of slag and ground synthetic quartz sand (US). Soda was used as activator. The inclusions of US were 0%, 10%, 20%, 30%, 40%, 50% and 60%, by weight. The surface area of the slag was 3000 cm2/g, whilst three different surface areas of US (2000, 5000 and 8000 cm2/g) were employed. They reported that the addition of US with concentrations of 1020 wt% had no effect on the setting time. Increasing additive content of US up to 60% led to increasing setting time by a factor of 23. The addition of US with 2000 cm2/g surface area led to decreasing the compressive strength at 1 year age, whilst the incaution of US at range1030% with 5000 cm2/g surface area, increased the compressive strength. Higher levels of US, beyond 30%, decreased the compressive strength. 1040% US with 8000 cm2/g surface area led to increasing the compressive strength. Beyond 40% of US, the reduction in compressive strength was noted. From the above review of literature of this part, it can be noted that the main advantage of employing barium chloride with carbonatite/slag is increasing the setting time. On the same line with this, the inclusion of 8% NaCl or malic acid or ground synthetic quartz sand in AAS matrix retarded the setting time that AAS system suffers. The inclusion of phosphoric acid in AAS matrix retarded the setting time, but increased the drying shrinkage. The inclusion of gypsum or burnt magnesia in AAS matrix reduced the drying shrinkage that AAS suffers. On the contrary, the inclusion of gypsum in AAS matrix reduced the setting time. However, Table 13 summarizes the pervious researches that studied the effect of different additives on some properties of AAS system. From this literature review, it can be noted that the main disadvantages of AAS system that previously mentioned in the introduction part (i.e. rapid setting, higher shrinkage values, higher formation of salt eforescences, higher carbonation and tendency to cracks during curing) can be mitigated or limited by using the suitable additives. However, Table 14 summarizes the main disadvantages or the main defects of AAS system and the suitable additives used to limit or mitigate these defects or disadvantages.

6. Conclusions The admixed slag as a source of alkali activation has been widely investigated in the recent years. The disadvantages of AAS system that previously mentioned in the introduction part (i.e. rapid setting, higher shrinkage values, higher formation of salt eforescences, higher carbonation and tendency to cracks during curing) can be mitigated and limited by using the suitable additives as shown in Table 14. However, the general conclusions of this literature review can be summarized as following:

1. Water absorption and permeability porous quality were markedly improved with the addition of steel bres in AAS matrix. Both splitting and exural strengths were largely improved with increasing bres volumes, but compressive strength decreased. In the alkali-activated slag/SF with the ratio of 80/20, as the steel bres increased as the workability and drying shrinkage decreased. Compressive and exural strengths as well as toughness increased with the increase in bres contents and bres lengths. 2. Incorporation of polypropylene bres did not affect positively the mechanical behaviour, modulus of elasticity and freezing/ thawing resistance of the AAS mortars, whilst PVA bres signicantly improved the tensile strain of AAS mortars. 3. Carbon bres in AAS mortars failed to improve the compressive strength, but reduced the drying shrinkage, whilst glass bres lowered the shrinkage without adverse effect on the mechanical properties. Basalt bres did not affect the dynamic compressive strength, but increased the energy absorption. 4. Superplasticizer with naphthalene-based improved the workability, the compressive strength and retarded the initial and nal setting times, but this depending on the activator type and activator concentration. On the other hand, some researchers recorded that it was not desirable to use Superplasticizer, based on modied naphthalene formaldehyde polymers, in AAS concrete, where it decreased the compressive strength and increased the drying shrinkage. 5. AEA was the most suitable for using in AAS concrete. It reduced the drying shrinkage and improved the workability and there was no reduction in the compressive strength after 7 days. 6. Both SSR and SHR reduced the compressive strength and the exural strength of AAS mortars, at age of 2 days. Beyond 2 days, SSR had no impact on the compressive and exural strengths, whilst SHR increased the exural strength of the mortars without changing in the compressive strength. Both SSR and SHR reduced the drying shrinkage of the AAS mortars. 7. The addition of 2% shrinkage reducing admixture increased the exural strength and signicantly reduced the drying shrinkage. 8. The inclusion of SF in AAS matrix decreased the workability and increased the compressive strength. The optimal replacement was 5% or 510%. 9. Both residual strength and relative strength of hardened neat AAS paste at 1000 C were signicantly higher than that of hardened alkali-activated slag/SF pastes. On the same line with this, the hardened neat AAS paste had thermal shock 1.75 times greater than that of the other hardened alkaliactivated slag/SF pastes. 10. SF as activator had positive effect in AAS matrix. The positive effect of SF activator seemed to be based on the intensication of the production of CSH and the densifying of the forming pore structure. 11. The inclusion of FA in AAS matrix increased the workability and setting time. Most of researchers believed that as the FA content increased in slag matrix, as the compressive strength decreased. Other researchers believed that the optimal content of FA in slag matrix that gave highest compressive strength depends on many factors as activator type, activator concentration, modulus ratio and curing condition. 12. Alkali-activated slag/FA (50/50) gave the lowest weight loss after exposure to CO2-laden H2SO4. On the other hand, slag/ FA geopolymer exhibited dimensional instability after magnesium sulfate attack, whilst sodium sulfate attack did not cause any degradation.

52

A.M. Rashad / Construction and Building Materials 47 (2013) 2955

13. FA/slag/K-silicate/H2O geopolymers inhibited the leaching of heavy metals, with the result that a number of traced pollutants, such as Ba, Be, Bi, Cd, Co, Cr, Nb, Ni, Pb, REE, Sr, Th, U, Y and Zr were retained. 14. The inclusion of MK in AAS matrix increased total setting time and reduced the heat release. 15. Although some researchers believed that replacement slag with 20% MK gave the highest compressive strength, but it depends on activator type, activator concentration and modulus Na2O/Al2O3 ratio. Other researchers believed that the inclusion of MK in AAS matrix reduced the compressive strength; as the MK content increased, as the compressive strength decreased. 16. Some researchers believed that the inclusion of MK in AAS matrix increased the permeability. Others believed that the inclusion of MK associated with high concentration of activator, in most cases, gave lower chloride permeability. On the contrary, the inclusion of MK, in slag matrix, increased the carbonation depth, increased the total pore volumes of the specimens and decreased the compressive strength, after exposure to carbonation. 17. The inclusion of 20% slag in MK matrix gave higher residual compressive strength than neat MK-based geopolymer after exposure to elevated temperatures up to 800 C. 18. Replacing slag with 5%, 10% and 15% MK in alkali activation slag/MK had positive effect on increasing Kd ratio for Sr ion and Cs ion better than neat AAS system, as the MK increased as the Kd ratio increased. In addition, slag/MK-based geopolymer could effectively immobilized Cu and Pb heavy metal. 19. Some researchers believed that replacing 10% slag with PC in AAS matrix gave higher compressive strength and exhibited the lowest drying shrinkage which similar to the drying shrinkage of the traditional PC. On the contrary, other researchers believed that the inclusion of PC in AAS matrix reduced compressive strength. 20. The inclusion of alkali activator in slag/PC reduced the porosity and median pore size and increased density. The chloride diffusion showed decreasing diffusion rate with decreasing PC content. The acidic resistance decreased with increasing PC content in AAS matrix. 21. The inclusion of lime in AAS matrix delayed setting time, enhanced workability and early age compressive strength. On the other hand, the inclusion of lime in AAS matrix reduced later age compressive strength. 22. 1% of resin powder of ethylenevinyl acetate copolymer or 1% of polyvinyl acetate resin increased the compressive and exural strengths of slag/MK composite. 23. The inclusion of barium chloride in alkali-activated slag/carbonatite, increased the setting time and reduced the compressive strength. The inclusion of air cooled slag in AAS matrix, shortened the setting times and reduced the compressive strength. 24. Burnt magnesia at 850950 C could be used to reduce the shrinkage of AAS mortar when its amount did not exceed 8%; otherwise, the setting time might be too short and the compressive and exural strengths might severely decreased. 25. 8% NaCl, 0.5% malic and its combination could be used as a retarder in AAS system. 26. Replacing pumice with slag, in alkali-activated pumice-type natural pozzolan, had very minor decrease in the tendency towards eforescence and improved the compressive strength. Compressive strength increased with the increase slag content.

27. The incorporation of gypsum in AAS matrix signicantly reduced the drying shrinkage, reduced the setting time and increased the compressive strength. On the other hand, the inclusion of phosphoric acid increased the setting time, reduced the compressive strength at early age and increased the drying shrinkage. 7. Comments for further studied It is recommended to study the effects of more than one menial admixture (ternary or quaternary) blends with slag in AAS system. It is recommended to study the effects of chemical admixtures and/or mineral admixtures on the carbonation depth of AAS system. It is recommended to study the effects of chemical admixtures and/or mineral admixtures on the formation of salt eforescences and the tendency to cracks during curing. It is recommended to study the effect of SF, QP, MK and their combinations of the drying shrinkage and setting time of AAS system. References
[1] Metz B, Davidson OR, Bosch PR, Dave R, Meyer LA. Climate change, mitigation, construction of working group III to the fourth assessment report of intergovernmental panel on climate change IPCC. Cambridge (UK): Cambridge University Press; 2007. [2] Mehta PK, Walters M. Roadmap to a sustainable concrete construction industry. Constr Specier 2008;61:4857. [3] Malhotra VM, Mehta PK. High-performance, high-volume y ash concrete: materials, mixture, proportioning, properties, construction practice, and case histories. In: 2nd edn. supplementary cementing materials for sustainable development incorporated, Ottawa, 2005. p. 1124. [4] Rangan BV, Hardjto D. Development and properties of low calcium y ash based geopolymer concrete. Research report GC-1, Faculty of Engineering, Curtins University of Technology, Perth, Australia, 2005. [5] Gartner E. Industrially interesting approaches to low-CO2 cements. Cem Concr Res 2004;34:148998. 9 pez-Hombrados Cecilio, Lieyda Jos [6] Palomo ngel, Fernndez-Jimnez Ana, Lo Luis. Railway sleepers made of alkali activated y ash concrete. Revista 9 a de Construccio 9 n 2007;22(2):7580. Ingenieri [7] Taylor M, Gielen D. Energy efciency and CO2 emissions from the global cement industry. Int. Energy Agency; 2006. [8] Naik T. Sustainability of concrete construction. ASCE Pract Period Struct Des Constr 2008;13(2):98103. [9] Anand S, Vrat P, Dahiya RP. Application of a system dynamics approach for assessment and mitigation of CO2 emission from the cement industry. J Environ Manage 2006;79(4):38398. [10] Hendriks CA, Worrell E, Price L, Mertin N, Ozawa Media L. Greenhouse gases from cement production. IEA Greenhous Gas R&D Programme, Report number PH 3/7. Utrecht (The Netherlands): Ecofys; 1999. [11] Stern N. Stern review on economics of climate change. Cambridge University Press; 2006. [12] Caijun Shi, Jueshi Qian. High performance cementing materials from industrial slags a review. Resour Conserv Recycl 2000;29:195207. [13] Rashad AM, Bai Y, Basheer PAM, Milestone NB, Colier NC. Hydration and properties of sodium sulfate activated slag. Cem Concr Compos 2013;37:209. [14] 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. Construct Build Mater 2011;25:3098107. [15] Taylor HFW. Cement chemistry. London: Thomas Telford; 1997. [16] Hendriks CA, Worrell E, Jager D, Blok K, Riemer P. Emission reduction of greenhouse gases from the cement industry. In: Proceedings of the 7th international conference on greenhouse gas control technologies, IEA GHG R&D Probram, Vancouver, Canada, 2004. [17] Sakulich Aaron Richard, Miller Sean, Barsoum Michel W. Chemical and microstructural characterization of 20-month-old alkali-activated slag cements. J Am Soc 2010;93(6):17418. [18] Morsy MS, Rashad AM, Shebl SS. Effect of elevated temperature on compressive strength of blended cement mortar. Build Res J, BRJ 2008;56(23):17385 [Institute of Construction and Architecture, Slovak Academy of Science]. [19] Morsy MS, Shebl SS, Rashad AM. Effect of re on microstructure and mechanical properties of blended cement pastes containing metakaolin and silica fume. Silic Ind 2009;74(34):5964.

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 [20] Seleem Hosam El-Din H, Rashad Alaa M, Elsokary Tarek. Effect of elevated temperature on physico-mechanical properties of blended cement concrete. Construct Build Mater 2011;25:100917. [21] Rashad Alaa M, Seleem Hosan El-Din H, Yousry Khalid M. Compressive strength of concrete mixtures with binary and ternary cement blends. Build Res J, BRJ 2009;57(2):10730 [Institute of Construction and Architecture, Slovak Academy of Science]. [22] Rashad AM. Effect of incorporating blended cement on fresh and hardened properties of concrete. Silic Ind 2009;74(78):24554. [23] Rashad AM. Chloride ion permeability of plain and blended cement concretes. Build Res J, BRJ 2011;59(12):3952 [Institute of Construction and Architecture, Slovak Academy of Science]. [24] Seleem Hosam El-Din H, Rashad Alaa M, El-Sabbagh Basil. Durability and strength evaluation of high-performance concrete in marine structures. Construct Build Mater 2010;24:87884. [25] Rashad Alaa M, Zeedan Sayieda R. A preliminary study of blended pastes of cement and quartz powder under the effect of elevated temperature. Construct Build Mater 2012;29:67281. [26] Gazquez MJ, Bolivar JP, Vaca F, Garca-Tenorio R, Caparros A. Evaluation of the use of TiO2 industry red gypsum waste in cement production. Cem Concr Compos 2013;37:7681. [27] Uchikawa H, Hanehara Shunsuke. Recycling of waste as an alternative raw material and fuel in cement manufacturing. Waste Mater Used Concr Manufact 1996:430553. [28] Chen Ying-Liang, Chang Juu-En, Shih Pai-Haung, Ko Ming-Sheng, Chang YiKuo, Chiang Li-Choung. Reusing pretreated desulfurization slag to improve clinkerization and clinker grindability for energy conservation in cement manufacture. J Environ Manage 2010;91:18927. [29] Pan Jill R, Huang Chihpin, Kuo Jung-Jen, Lin Sheng-Huan. Recycling MSWI bottom and y ash as raw materials for Portland cement. Waste Manage 2008;28:11138. [30] Tan Y, Zhou Q, Wei R. The initial study on the use of copper slag as a additional materials for cement production. J Xinjiang Inst Technol 2000;21(3):2369. [31] Huang K. Use of copper slag in cement production. Sichuan Cement 2001;4:257. [32] Puertas F, Garca-Daz I, Barba A, Gazulla MF, Palacios M, Gmez MP, et al. Ceramic wastes as alternative raw materials for Portland cement clinker production. Cem Concr Compos 2008;30:798805. _ Deveci H, Sngn H. Utilization of otation wastes of copper slag as raw [33] Alp I, material in cement production. J Hazard Mater 2008;159:3905. _ Isabel Snchez de Rojas M. Total and soluble chromium, nickel [34] Moiss FrIas, and cobalt content in the main materials used in the manufacturing of Spanish commercial cements. Cem Concr Res 2002;32:43540. [35] Davidovits J. Geopolymer properties and chemistry. In: 1st European conference on soft mineralurgy, geopolymer 88 Compiegne France, 1998. p. 2548. [36] Weil M, Dombrowski K, Buchawald A. Life-cycle analysis of geopolymers. In: Provis J. Van Deventer J, editors. Geopolymers, structure, processing, properties and applications. Abington Hall (Cambridge, UK): Woodhead Publishing Limited; 2009. p. 194210. ISBN-13:97818456914494. [37] Shi C, Krivenko PV, Roy D. Alkali-activated cements and concretes. USA and Canada: Taylor and Francis; 2006. [38] Bakharev T, Sanjayan JG, Cheng YB. Alkali activation of Australian slag cements. Cem Concr Res 1999;29:11320. [39] Fernndez-Jimnez A, Palomo JG, Puertas F. Alkali-activated slag mortars mechanical strength behavior. Cem Concr Res 1999;29:131321. [40] Chen W, Brouwers HJH. The hydration of slag. Part 1: Reaction models for alkali-activated slag. J Mater Sci 2007;42:42843. [41] Bakharev T, Sanjayan JG, Cheng Y-B. Sulfate attack on alkali-activated slag concrete. Cem Concr Res 2002;32:2116. [42] Bakharev T, Sanjayan JG, Cheng Y-B. Resistance of alkali-activated slag concrete to acid attack. Cem Concr Res 2003;33:160711. [43] Puertas F, Amat T, Fernndez-Jimnez A, Vazquez T. Mechanical and durable behaviour of alkaline cement mortars reinforced with polypropylene bers. Cem Concr Res 2003;33:20316. [44] Rashad Alaa M, Khalil Mervat H. A preliminary study of alkali-activated slag blended with silica fume under the effect of thermal loads and thermal shock cycles. Construct Build Mater 2013;40:52232. [45] Rashad AM, Bai Y, Basheer PAM, Collier NC, Milestone NB. Chemical and mechanical stability of sodium sulfate activated slag after exposure to elevated temperature. Cem Concr Res 2012;42:33343. [46] Roy Della M, Jiang Weimin, Silsbee MR. Chloride diffusion in ordinary, blended, and alkali-activated cement pastes and its relation to other properties. Cem Concr Res 2000;30:187984. [47] Shi C, Xie P. Interface between cement paste and quartz sand in alkaliactivated slag mortars. Cem Concr Res 1998;28:88796. [48] Brough AR, Atkinson A. Automated identication of the aggregate-paste interfacial transition zone in mortars of silica sand Portland or alkaliactivated slag cement paste. Cem Concr Res 2000;30:84954. ivica V. Effects of type and dosage of alkaline activator and temperature on [49] Z the properties of alkali-activated mixtures. Construct Build Mater 2007;21:14639. [50] Collins PG, Sanjayan JG. Workability and mechanical properties of alkali activated slag concrete. Cem Concr Res 1999;29:4558.

53

[51] Bakharev T, Sanjayan JG, Cheng YB. Resistance of alkali-activated slag concrete to carbonation. Cem Concr Res 2001;31:127783. [52] Hakkinen T. The inuence of slag content on the microstructure, permeability, and mechanical properties. Cem Concr Res 1993;23:40721. [53] Collins F, Sanjayan JG. Cracking tendency of alkali activated slag subjected to restrained shrinkage. Cem Concr Res 2000;30:7918. [54] Shi CJ, Krivenko PV, Roy DM. Alkali-activated cements and concretes. Taylor & Francis; 2005. p. 1376. [55] Li Z, Ding Z, Zhang Y. Development of sustainable cementitious materials. In: Proceedings of international workshop on sustainable development and concrete technology, Beijing, Chain, 2004. p. 5576. [56] Swamy RN. Concrete technology and design, vol. 3. Cement replacement materials, 1st ed. London: Surrey University Press; 1986. [57] CSN EN 1971 Cement Part1: Composition, specications and conformity esky institute, 2001. normalizac criteria for common cements. Praha: C [58] Roy DM. Alkali activated cements opportunities and challenges. Cem Concr Res 1999;29(2):24954. [59] Glukhovsky VD. Soil silicates. Russia: Gosstroiizdat Kiev; 1959 [in Russian]. [60] Purdon AO. The action of alkalis on blast-furnace slag. J Soc Chem Ind 1940;59:191202. [61] Wang SD. Review of recent research on alkali-activated concrete in China. Mag Concr Res 1991;43(254):2935. [62] Shi C, Wu Xueqan, Tang Mingshu. Research on alkali-activated cementitious systems in China: a review. Adv Cem Res 1993;5(17):17. [63] Smith MA, Osborne GJ. Slag/y ash cements. World Cem Technol 1977:22333. [64] Fross B. F-cement: a new low-porosity slag cement. Silic Ind 1983;48(3):7982. [65] Talling B. Effect of curing conditions on alkali-activated slags. In: Proc. 3rd int. conf. on y ash, silica fume, slag and natural pozzolans in concrete, Trondheim, 2, SP114-72, 1989. p. 14851500. [66] Roy DM, Silsbee MR. Alkali activated cementitious materials: an overview. Mater Rese Soc Symp Proc 1992;245:15364. [67] Clarke WJ, Helal M. Alkali activated slag and Portland/slag ultrane cement. Mater Res Soc Symp Proc 1991;179:21932. [68] Shi Caijun, Fernndez-Jimnez A. Stabilization/solidication of hazardous and radioactive wastes with alkali-activated cements. J Hazard Mater B 2006;137:165663. [69] Kovalchuk G, Fernndez-Jimnez A, Palomo A. Alkali-activated y ash: effect of thermal curing conditions on mechanical and microstructural developmentPart II. Fuel 2007;86:31522. [70] Khale Divya, Chaudhary Rubina. Mechanism of geopolymerization and factors inuencing its development: a review. J Mater Sci 2007;42:72946. [71] Al Bakri Mohd Mustara, Mohammed H, Kamarudin H, Khairul Niza I, Zarina Y. Review on y ash-based geopolymer concrete without Portland cement. J Eng Technol Res 2011;3(1):14. [72] Al Bakri Mustafa AM, Kamaradin H, Bulussain M, Khairul Nizar I, Mastura WIW. Mechanism and chemical reaction of y ash geopolymer cement a review. J Asian Sci Res 2011;1(5):24753. [73] Komnitsas Kostas, Zaharaki Dimitra. Geopolymerisation: a review and prospects for the minerals industry. Miner Eng 2007;20:126177. [74] Obonyo Esther, Kameeu Elie, Melo Uphie C, Leonelli Cristina. Advancing the use secondary inputs in geopolymer binders for sustainable cementitious composites: a review. Sustainability 2011;3:41023. [75] Li Chao, Sun Henghu, Li Longtu. A review: the comparison between alkaliactivated slag (Si + Ca) and metakaolin (Si + Al) cements. Cem Concr Res 2010;40:13419. [76] Heah CY, Kamarudin H, Mustafa Al Bakri AM, Luqman M, Khairul Nizar I, Liew YM. Potential application of kaolin without calcine as Greener concrete: a review. Aust J Basic Appl Sci 2011;5(7):102635. [77] Liew YM, Kamarudin H, Mustafa Al Bakri AM, Luqman M, Khairul Nizar I, Heah CY. Investigating the possibility of utilization of kaolin and the potential of metakaolin to produce green cement for construction purposes a review. Aust J Basic Appl Sci 2011;5(9):4419. [78] Rashad Alaa M. Alkali-activated metakaolin: a short guide for Civil Engineer an overview. Construct Build Mater 2013;41:75165. [79] Fernando Pacheco-Torgal, Joo Casto-Gomes, Said Jalali. Alkali-activated binders: a review Part 1. Historical background, terminology, reaction mechanisms and hydration products. Construct Build Mater 2008;22:130514. [80] Fernando Pacheco-Torgal, Joo Casto-Gomes, Said Jalali. Alkali-activated binders: A review. Part 2. About materials and binders manufacture. Construct Build Mater 2008;22:131522. [81] Richard Sakulich Aaron. Reinforced geopolymer composites for enhanced material greenness and durability. Sustain Cities Soc 2011;1:195210. [82] Pacheco-Torgal F, Abdollahnejad Z, Cames AF, Jamshidi M, Ding Y. Durability of alkali-activated binders: a clear advantage over Portland cement or an unproven issue? Construct Build Mater 2012;30:4005. [83] Pacheco-Togal F, Ding Y, Miraldo S, Abdollahnejad Z, Labrincha JA. Are geopolymers more suitable than Portland cement to produce high volume recycled aggregates HPC? Construct Build Mater 2012;36:104852. [84] Abdullah A, Jaafar MS, Tauq-yap YH, Alhpzaimy A, Al-Negheimish A, Noorzaei J. The effect of various chemical activators on pozzolanic reactivity: a review. Sci Res Essays 2012;7(7):71929.

54

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 [116] Yang Keun-Hyeok, Song Jin-Kyu. Workability loss and compressive strength development of cementless mortars activated by combination of sodium silicate and sodium hydroxide. J Mater Civil Eng, ASCE 2009:11927. [117] Yang Keun-Hyeok, Song Jin-Kyu, Lee Kang-Seok, Ashrour Ashraf F. Flow and compressive strength of alkali-activated mortars. ACI Mater J 2009:508. [118] Yang Keun-Hyeok, Song Jin-Kyu, Ashour Ashraf F, Lee Eun-Taik. Properties of cementless mortars activated by sodium silicate. Construct Build Mater 2008;22:19819. [119] Talling B, Brandstetr J. Present state and future of alkali-activated slag concretes. In: 3rd Int. con. on y-ash, silica fume, slag and natural pozzolanas in concrete, vol. 2, Norway. SP 11474, 1989. p. 151946. [120] Rashad Alaa M. Utilizing of alkali activation y ash concrete blended with slag. Iran J Mater Sci Eng (IJMSE), 2013;10(1):5764. [121] Parameswaran PS, Chatterjee AK. Alkali activation of Indian blast furnace slags. In: 8th int concr chem cem, Rio de Janeiro, vol. 4; 1986. p. 8691. [122] Sugama T, Brothers LE, Van de Putte TR. Acid-resistant cements for geothermal wells: sodium silicate activated slag/y ash blends. Adv Cem Res 2005;17(2):6575. [123] Kumar Sanjay, Kumar Rakesh, Mehrotra SP. Inuence of granulated blast furnace slag on the reaction, structure and properties of y ash based geopolymer. J Mater Sci 2010;45:60715. [124] Puertas F, Amat T, Fernndez-Jimnez A, Vzquez T. Mechanical and durable behaviour of alkaline cement mortars reinforced with polypropylene bers. Cem Concr Res 2003;33:20316. [125] Shi C, Day RL. Early strength development and hydration of alkali-activated blast furnace slag/y ash blends. Adv Cem Res 1999;11(4):18996. [126] Puertas F, Fernndez-Jimnez A. Mineralogical and microstructural characterization of alkali-activated y ash/slag pastes. Cem Concr Compos 2003;25:28792. [127] Lu C. The preliminary research of the y ash-slag-alkali concrete. In: Proceedings of the 2nd Beijing, vol. 2; 1989. p. 2329. [128] Smith MA, Osborne GJ. BFS/y ash cements. World Cem Technol 1977;8:22333. [129] Bijen J, Waltje H. Alkali activated slag-y ash cements. In: Proceedings of the 3rd international conference on the Use of y ash, silica fume, slag & natural pozzolans in, concrete, SP-114, 1989. p. 156578. [130] Dai L, Cheng J. An investigation on BFS-y ash-alkali systems. Bull Chin Silic Soc 1988;16:2332 [in Chinese]. 9 a Escalante, Campos-Venegas K, Gorokhovsky A, Fernndez A. [131] Garci Cementitious composites of pulverized fuel ash and blast furnace slag activated by sodium silicate: effect of Na2O concentration and modulus. Adv Appl Ceram 2006;105(4):2018. [132] Weiguo Shen, Wang Yiheng, Zhang Tao, Zhou Mingkai, Li Jiasheng, Cui Xiaoyn. Magnesia modication of alkali-activated slag y ash cement. J Wuhan Univ Technol Mater Sci Ed 2011:1215. [133] Kim Hyunjung, Kim Yooteak. Characteristics of the geopolymer using y ash and blast furnace slag with alkaline activators. In: 4th international conference on chemical, biological and environmental engineering, IPCBEE, Singapore, 2012. p. 1549. http://dx.doi.org/10.7763/IPCBEE [43, 32] [134] Puertas F, Martnez-Ramrez S, Alonso S, Vzquez T. Alkali-activated y ash/ slag cement strength behavior and hydration products. Cem Concr Res 2000;30:162532. [135] Zhang Dajie, Liu Wenshi, Hou Haobo, He Xighua. Strength, leachability and microstructure characterization of Na2SiO3-activated ground granulated blast-furnace slag solidied MSWI y ash. Waste Manage Res 2007;25:4027. [136] Guerrieri Maurice, Sanajayan Jay G. Behavior of combined y ash/slag-based geopolymers when exposed to high temperatures. Fire Mater 2009. http:// dx.doi.org/10.1002/fam.1014. [137] Chi Maochieh, Huang Ran. Binding mechanism and properties of alkaliactivated y ash/slag mortars. Construct Build Mater 2013;40:2918. [138] Goretta KC, Chen Nan, Gutierrez-Mora F, Roubort JL, Lukey GC, van Deventer JSJ. Solid-particle erosion of a geopolymer containing y ash and blastfurnace slag. Wear 2004;256:7149. [139] Ismail Idawati, Bernal Susan A, Provis John L, Hamdan Sinin, van Deventer Jannie SJ. Microstructural changes in alkali activated y ash/slag geopolymers with sulfate exposure. Mater Struct 2012. http://dx.doi.org/10.1617/s11527012-9906-2. [140] Izquierdo Maria, Querol Xavier, Davidovits Joseph, Antenucci Diano, Nugteren Henk, Fernndez-Pereira Constantino. Coal y ash-slag-based geopolymer: microstructure and metal leaching. J Hazard Mater 2009;166:5616. [141] Bai Yun, Milestone Neil B, Yang Changhui. Sodium sulphate activated GGBS/ PFA and its potential as a nuclear waste immobilization matrix. Mater Res Symp Proc Mater Res Soc 2006;932. [142] Bernal Susan A, Provis John L, Rose Volker, de Gutierrez Ruby M. Evolution of binder structure in sodium silicate-activated slag-metakaolin blends. Cem Concr Compos 2011;33:4654. [143] Cheng TW, Chiu JP. Fire-resistant geopolymer produced by granulated blast furnace slag. Miner Eng 2003;16:20510. -Wczelik Wieslawa. Heat evolution in alkali activated synthetic slag[144] Nocun metakaolin mixtures. J Thermal Anal Calorim 2006;86(3):73943. [145] Buchwald Anja, Tatarin R, Stephan D. Reaction progress of alkaline-activated metakaolin-ground granulated blast furnace slag blends. J Mater Sci 2009;44:560917.

[85] Shi Caijun, Jimnez A Fernndez, Palomo Angel. New cement for the 21st century: the pursuit of an alternative to Portland cement. Cem Concr Res 2011;41:75063. [86] Bernal Susan, De Gutierrez Ruby, Delvasto Silvio, Rodriguez Erich. Performance of an alkali-activated slag concrete reinforced with steel bers. Construct Build Mater 2010;24:20814. [87] Aydn Serdar, Baradan Blent. The effect of ber properties on high performance alkali-activated slag/silica fume mortars. Composites: Part B 2013;45:639. [88] Bernal Susan, De Gutierrez Ruby, Delvasto Silvio, Rodriguez Erich. Performance of geopolymeric concrete reinforced with steel bers. In: 10th int. inorganic-bonded ber composites conference, IIBCC 2006, So Paulo Brazil, Universidade de So Paulo & University of Idaho, So Paulo, October 1518 2006. p. 15667. [89] Lee Bang Yeon, Cho Chang-Geun, Lim Hyun-Jin, Song Jin-Kyu, Yang KeunHyeok, Li Victor C. Strain hardening ber reinforced alkali-activated mortar a feasibility study. Construct Build Mater 2012;37:1520. [90] Alcaide JS, Alcocel EG, Puertas F, Lapuente R, Garces P. Carbon berreinforced, alkali-activated slag mortars. Mater Constr 2007;57:3348. [91] Puertas F, Gil-Maroto A, Palacios M, Amat T. Alkali-activated slag mortars reinforced with glass ber, performance and properties. Mater Constr 2006;56:7990. [92] Li W, Xu J. Impact characterization of basalt ber reinforced geopolymeric concrete using a 100-mm-diameter split Hopkinson pressure bar. Mater Sci Eng A 2009;513514:14553. [93] Li W, Xu J. Mechanical properties of basalt ber reinforced geopolymeric concrete under impact loading. Mater Sci Eng A 2009;505:17886. [94] Silva FJ, Thaumaturgo C. Fiber reinforcement and fracture response in geopolymeric mortars. Fatigue Fract Eng Mater Struct 2003;26:16772. [95] Natali A, Manzi S, Bignozzi MC. Novel ber-reinforced composite materials based on sustainable geopolymer matrix. Proc Eng 2011;21:112431. [96] Douglas E, Brandstetr J. A preliminary study on the alkali activation of ground granulated blast-furnace slag. Cem Concr Res 1990;20:74656. [97] Wang Shao-Dong, Scrivener Karen L, Pratt PL. Factors affecting the strength of alkali-activated slag. Cem Concr Res 1994;24(6):103343. [98] Puertas F, Palomo A, Fernndez-jimnez A, Izquierdo JD, Granizo ML. Effect of superplasticisers on the behaviour and properties of alkaline cements. Adv Cem Res 2003;15:238. [99] Palacios M, Puertas F. Effect of superplasticizer and shrinkage-reducing admixtures on alkali-activated slag pastes and mortars. Cem Concr Res 2005;35:135867. [100] Palacios M, Houst YF, Bowen P, Puertas F. Adsorption of superplasticizer admixtures on alkali-activated slag pastes. Cem Concr Res 2009;39:6707. [101] Palacios Marta, Banll Phillip FG, Puertas Francisca. Rheology and setting of alkali-activated slag pastes and mortars: effect of organic admixtures. ACI Mater J 2008:1408. [102] Bakharev T, Sanjayan JG, Cheng Y-B. Effect of admixtures on properties of alkali-activated slag concrete. Cem Concr Res 2000;30:136774. _ Serhan. Inuence of [103] Bilim Cahit, Karahan Okan, Atis Cengiz Duran, Ilentapar admixtures on the properties of alkali-activated slag mortars subjected to different curing conditions. Mater Des 2013;44:5407. [104] Palacios M, Puertas F. Effect of shrinkage-reducing admixtures on the properties of alkai-activated slag mortars and pastes. Cem Concr Res 2007;37:691702. [105] Collins Frank, Sanjayan JG. Effects of ultra-ne materials on workability and strength of concrete containing alkali-activated slag as the binder. Cem Concr Res 1999;29:45962. [106] Rashad Alaa M, Zeedan Sayieda R, Hassan Hassan A. A preliminary study of autoclaved alkali-activated slag blended with quartz powder. Construct Build Mater 2012;33:707. [107] Escalante-Garca Jose Ivan, Palacios-Villanueva Victor M, Gorokhovsky Alexander V, Mendoza-Surez Guillermo, Fuentes Antonio F. Characteristics of a NaOH-activated blast furnace slag blended with a ne particle silica waste. J Am Ceram Soc 2002;7(85):178892. [108] Escalante-Garca JI, Gorokhovsky AV, Mendoza G, Fuentes AF. Effect of geothermal waste on strength and microstructure of alkali-activated slag cement mortars. Cem Concr Res 2003;33:156774. [109] Aydn Serdar. A ternary optimization of mineral additives of alkali activated cement mortars. Construct Build Mater 2013;43:1318. ivica V. Silica fume-basic blast furnace slag systems [110] Rousekov I, Bajza A, z activated by an alkali silica fume activator. Cem Concr Res 1997; 27(12):18258. [111] Brew DRM, Mackenzie KJD. Geopolymer synthesis using silica fume and sodium aluminate. J Mater Sci 2007;42:39903. [112] 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. http://dx.doi.org/10.1007/s12649-011-9093-3. ivica Vladimr. Acidic resistance of materials based on the novel use of silica [113] Z fume in concrete. Construct Build Mater 1999;13:2639. ivica Vladimr. High effective silica fume alkali activator. Bull Mater Sci [114] Z 2004;27(2):17982. ivica Vladimr. Effectiveness of new silica fume alkali activator. Cem Concr [115] Z Compos 2006;28:215.

A.M. Rashad / Construction and Building Materials 47 (2013) 2955 [146] Yip CK, Lukey GC, van Deventer JSJ. The coexistence of geopolymeric gel and calcium silicate hydrate at the early stage of alkaline activation. Cem Concr Res 2005;35:168897. 9 guez Erich D, de Gutirrez Ruby Meji 9 a, Gordillo Marisol, [147] Bernal Susan A, Rodri Provis Johan L. Mechanical and thermal characterisation of geopolymers based on silicate-activated metakaolin/slag blends. J Mater Sci 2011. http:// dx.doi.org/10.1007/s10853-011-5490-z. 9 guez Erich D, de Gutirrez Ruby Mejia, Provis John L, [148] Bernal Susan A, Rodri Delvasto Silvio. Activation of metakaolin/slag blends using alkaline solutions based on chemically modied silica fume and rice husk ash. Waste Biomass Valor 2011. http://dx.doi.org/10.1007/s12649-011-9093-3. [149] Bernal Susan A, de Gutierrez Ruby Meja, Provis Johan L, Rose Volker. Effect of silicate modulus and metakaolin incorporation on the carbonation of alkali silicate-activated slags. Cem Concr Res 2010;40:898907. [150] Chen Song, Wu Mengqiang, Zhang Shuren. Mineral phases and properties of alkali-activated metakoalin-slag hydroceramics for a disposal of simulated highly-alkaline wastes. J Nucl Mater 2010;402:1738. [151] Buchwald A, Hilbig H, Kaps Ch. Alkali-activated metakaolin-slag blendsperformance and structure in dependence of their composition. J Mater Sci 2007;42:302432. [152] Wang Jin, Wu Xiu-Ling, Wang Jun-Xia, Liu Chang-Zhen, Lai Yuan-Ming, Hong Zun-Ke. Hydrothermal synthesis and characterization of alkali-activated slag-y ash-metakaolin cementitious materials. Micropor Mesopor Mater 2012;155:18691. 9 az Oswaldo, Escalante-Garci 9 a Jose Ivan, Arellano Ral, [153] 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. [154] Bernal Susan A, de Gutirrez Ruby Mejia, Provis John L. Engineering and durability properties of concretes based on alkali-activated granulated blast furnace slag/metakaolin blends. Construct Build Mater 2012;33:99108. [155] Yip Christina K, Lukey Grant C, Provis John L, van Deventer Jannie SJ. Effect of calcium silicate sources on geopolymerisation. Cem Concr Res 2008;38:55464. [156] 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. [157] 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. [158] 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. [159] Guangren Qian, Facheng YI, Shi Rongming. Improvement of metakaolin on radioactive Sr and Cs immobilization of alkali-activated slag matrix. J Hazard Mater 2002;B92:289300. [160] Wu Xuequan, Jiang Weimin, Roy DM. Early activation and properties of slag cement. Cem Concr Res 1990;22:96174. [161] Fu-Sheng Wang, Rui-Lian Sun, Ying-Jing Cui. Study on modication of the high-strength slag cement material. Cem Concr Res 2005;35:13448. [162] Roy S, Chanda S, Bandopadhyay SK, Ghosh SN. Investigation of Portland slag cement activated by waterglass. Cem Concr Res 1998;28(7):104956. [163] Bilim Cahit, Atis Cengiz Duran. Alkali activation of mortars containing different replacement levels of ground granulated blast furnace slag. Construct Build Mater 2012;28:70812. ivica Vladimr. Alkali-silicate admixture for cement composites [164] Z incorporating pozzolan or blast furnace slag. Cem Concr Res 1993;23:121522. [165] Acevedo-Martinez E, Gomez-Zamorano LY, Escalante-Garcia JI. Portland cement-blast furnace slag mortars activated using waterglass Part 1: Effect of slag replacement and alkali concentration. Construct Build Mater 2012;37:4629. [166] Singh N, Rai S, Singh NB. Effect of sodium sulfate on the hydration of granulated blast furnace slag blended Portland cement. Indian J Eng Mater Sci (IJEMS) 2001;8(2):1103. [167] Veiga KK, Gastaldini ALG. Sulfate attack on a white Portland cement with activated slag. Construct Build Mater 2012;34:494503.

55

[168] Pu XC, Gan CC, Wang SD, Yang CH. Research reports on alkali-activated slag cement and concrete, vol. 6. China: Chongqing Institute of Architecture and Engineering; 1988 [in Chinese]. [169] Douglas E, Biodeau A, Brandster J, Malhotra VM. Alkali activated ground granulated blast furnace slag concrete: preliminary investigation. Cem Concr Res 1991;21(1):1018. [170] Cheng QH, Tagnit-Hamou A, Sarkar SL. Strength and microstructural properties of waterglass activated slag. Mater Soc Symp Proc 1992;245:4954. [171] Yang Keun-Hyeok, Cho Ah-Ram, Song Jin-Kyu, Nam Sang-Ho. Hydration products and strength development of calcium hydroxide-based alkaliactivated mortars. Construct Build Mater 2012;29:4109. [172] Yang Keun-Hyeok, Song Jin-Kyu. Empirical equations for mechanical properties of Ca(OH)2-based alkali-activated slag concrete. ACI Mater J 2012:43140. [173] Yang Keun-Hyeok, Cho Ah-Ram, Song Jin-Kyu. Effect of water-binder ratio on the mechanical properties of calcium hydroxide-based alkali-activated slag concrete. Construct Build Mater 2012;29:50411. [174] Li Zongjin, Liu Sifeng. Inuence of slag as additives on compressive strength of y ash-based geopolymer. J Mater Civil Eng, ASCE 2007:4704. [175] Zhang Yao Jun, Li Sheng, Xu De Long, Wang Bao Qiang, Xu Guo Ming, Yang Dong Feng. A novel method for preparation of organic resins reinforced geopolymer composites. J Mater Sci 2010;45:118992. [176] Zhang Yao Jun, Wang Ya Cho, Xu De Long, Li Sheng. Mechanical performance and hydration mechanism of geopolymer composite reinforced by resin. Mater Sci Eng A 2010;527:657480. [177] Zhang Yao Jun, Li Sheng, Wang Ya Cho, Xu De Long. Microstructural and strength evolutions of geopolymer composite reinforced by resin exposed to elevated temperature. J Non-Cryst Solids 2012;358:6204. [178] Sanyin Zhao et al. Setting and strength characteristics of alkali-activated carbonatite cementitious materials with ground slag replacement. J Wuhan Univ Technol Mater Sci Ed 2006;21(1):1258. [179] Fang Yong-Hao, Liu Jun-Feng, Chen Yi-Qun. Effect of magnesia on properties and microstructure of alkali-activated slag cement. Water Sci Eng 2011;4(4):4639. [180] Brough AR, Holloway M, Sykes J, Atkinson A. Sodium silicate-based alkaliactivated slag mortars: Part II. The retarding effect of additions of sodium chloride or malic acid. Cem Concr Res 2000;30:13759. ek V, Tomkov V, Babkov P, Vavro M. Alkali-activated composites based [181] Vlc on slags from iron and steel metallurgy. Metalurgija 2009;48(4):2237. [182] El-Didamony Hamdy, Amer Ahmed A, El-Sokkary Tarek M, Abd-El-Aziz Hamdy. Effect of substitution of granulated slag by air-cooled slag on the properties of alkali activated slag. Ceram Int 2013;39:17181. [183] Allahverdi Ali, Kani Ebrahim Naja, Yazdanpour Mahshad. Effects of blastfurnace slag on natural pozzolan-based geopolymer cement. Ceram-Silikty 2011;55(1):6878. [184] Kani Ebrahim Naja, Allahverdi Ali, Provis John L. Eforescence control in geopolymer binder based on natural pozzolan. Cem Concr Compos 2012;34:2533. [185] Chang Jiang Jhy, Yeih Weichung, Hung Chi Che. Effects of gypsum and phosphoric acid on the properties of sodium silicate-based alkali-activated slag pastes. Cem Concr Compos 2005;27:8591. [186] Dongxu Li, Xuequan Wu, Jinlin Shen, Yujiang Wang. The inuence of compound admixtures on the properties of high-content slag cement. Cem Concr Res 2000;30:4550. [187] Khater HM. Effect of cement kiln dust on geopolymer composition and its resistance to sulfate attack. Green Mater 2012:111. [188] Yongde Li, Yao Sun. Preliminary study on combined-alkali-slag paste materials. Cem Concr Res 2000;30:9636. niowski Rafa, [189] Brylicki Witold, Malolepszy Jan, Stryczek Stanislaw, Wis Kotwica Lukasz. Effects of modication of alkali activated slag cementing slurries with natural clinoptilolite. Gospodarka Surowcami Mineralnymi 2009;25:6175. [190] Esmaily H, Nuranian H. Non-autoclaved high strength cellular concrete from alkali activated slag. Construct Build Mater 2012;26:2006. [191] Rakhimova NR, Rakhimov RZ. Alkali-activated slag-blended cements with silica supplementary materials. Neorganicheskie Mater 2012;48(9):10838.

You might also like