You are on page 1of 65

Journal Pre-proof

Rheology of alkali-activated materials: A review

Cuifang Lu, Zuhua Zhang, Caijun Shi, Ning Li, Dengwu Jiao, Qiang Yuan

PII: S0958-9465(21)00130-X
DOI: https://doi.org/10.1016/j.cemconcomp.2021.104061
Reference: CECO 104061

To appear in: Cement and Concrete Composites

Received Date: 21 October 2020


Revised Date: 18 February 2021
Accepted Date: 10 April 2021

Please cite this article as: C. Lu, Z. Zhang, C. Shi, N. Li, D. Jiao, Q. Yuan, Rheology of alkali-
activated materials: A review, Cement and Concrete Composites (2021), doi: https://doi.org/10.1016/
j.cemconcomp.2021.104061.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier Ltd.


1 Rheology of alkali-activated materials: A review

2 Cuifang Lu a, b, Zuhua Zhang a, b*, Caijun Shi a, b*, Ning Li a, b, Dengwu Jiao a, c, Qiang

3 Yuan d
a
4 Key Laboratory for Green & Advanced Civil Engineering Materials and Application Technology of

5 Hunan Province, College of Civil Engineering, Hunan University, 410082, China

6 b International Science and Technology Innovation Cooperation Base for Green Advanced Civil

7 Engineering Materials of Hunan Province, Changsha, 410082, China


c
8 Magnel-Vandepitte Laboratory, Department of Structural Engineering and Building Materials,

f
oo
9 Ghent University, 9052 Ghent, Belgium

r
d
10 National Engineering Laboratory for High Speed Railway Construction, School of Civil

11
-p
Engineering, Central South University, Changsha 410075, China
re
12
lP

13 Abstract: Many alkali-activated materials (AAMs) exhibit poor workability in mixing owing to
na

14 great viscosity of the paste and fast setting. Chemical admixtures commonly used in Portland

15 cement-based materials, such as rheology control additives and superplasticizers, show lower
ur

16 efficiency in AAM systems. Understanding the rheological features of various AAMs and their
Jo

17 control can lead to a valuable guidance for their application in large-scale, especially in the field of

18 pumping and placement, grouting and 3D printing, etc. This paper aims to deliver a comprehensive

19 review of the effects of composition factors including activators, precursors, admixtures, additions,

20 aggregates and fibers on the rheology of AAMs, as well as models for describing their rheological

21 behavior. The current progress has shown that the large knowledge gap needs substantial research

22 from scientific mechanisms to robust formulations of paste, mortar and concrete to meet both

23 workability and mechanical requirements.

24 Keywords: Rheology; Alkali-activated materials; Composition; Models; Workability

25 *Corresponding authors.

26 E-mail addresses: zuhuazhang@hnu.edu.cn (Z. Zhang); cshi@hnu.edu.cn (C. Shi)

27
28 1. Introduction

29 In the past decades, alkali-activated materials (AAMs), including ‘geopolymers’[111], have been

30 accepted by both academic and industrial communities progressively as a family of effective and

31 promising alternatives to Portland cement-based materials (PCMs). These materials can be

32 manufactured by activation of a reactive solid aluminosilicate such as blast furnace slag [1, 2], fly

33 ash (Class-F fly ash in this paper unless otherwise stated) [3, 4], metakaolin [5, 6] and other

34 precursors [7-9] using an alkaline activator, which is usually a concentrated aqueous solution of

35 alkali hydroxide, silicate or carbonate. Properly designed AAMs can have denser microstructure than

f
oo
36 normal PCMs with superior engineering performances [10], particularly in aggressive environment

37 [11, 12]. Moreover, the manufacturing of AAMs avoids the energy consumption and CO2 emission

r
38 -p
with respect to clinker production, endowing this kind of materials great potential to reduce adverse
re
39 environmental impacts by a wide margin [13]. It should be noted that the use of sodium silicate leads

40 to a high CO2 emission and surface water acidification which can reduce the environmental benefit
lP

41 [112]. Hence a better mix design is required. However, the large-scale engineering application of
na

42 AAMs progresses slowly. Regardless the regulation obstacles, there seems more because of the

43 technical challenges. Besides the shrinkage risk of AAMs [14, 15] and the inconsistent raw materials
ur

44 and alkali activators used [8], it is more likely due to the facts such as poor workability [16-18]. In
Jo

45 addition, most commercial chemical admixtures for PCMs are incompatible in AAMs [19, 20].

46 Therefore, the effect of composition on rheological properties of AAMs must be fully understand so

47 as to control their workability.

48 In this paper, the effects of composition factor such as the nature and dosage of activators and

49 precursors on the rheology of AAMs are reviewed. It is expected to provide a comprehensive

50 perspective to understand workability and rheology of AAMs, and a reference for future research

51 towards better control of this new sustainable material in field application.

52 2. Rheological models applied in AAMs

53 Rheology is the science that studies the deformation and flow of matter. Studies have shown that

54 slump value is an indicative but not a sufficient measurement to characterize the fundamental
55 rheological properties of PCMs, namely plastic viscosity and yield stress, which are respectively a

56 measure of the material resistance to flow after the material begins to flow and a measure of the

57 shear stress required to initiate flow [21]. The relationships between shear stress and shear strain rate

58 of fresh PCMs have been developed. The Bingham model, modified Bingham model and

59 Herschel-Bulkley (H-B) model are the most widely accepted for describing the rheological behavior

60 of PCMs [22].

61 The Bingham model has a linear relationship between shear stress (τ) and shear strain rate (𝛾̇ ) as

62 described by Eq. (1) [23]:

63 (1)

f
τ = 𝜏0 + 𝜇𝛾̇

oo
64 where 𝜏0 is the yield stress (Pa) and 𝜇 is the plastic viscosity (Pa·s). This expression is the simplest

r
65 one for non-Newtonian fluids. However, the addition of admixtures in PCMs can lead to the

66
-p
deviation of a linear relationship between shear stress and shear rate. Consequently, the modified
re
67 Bingham model was proposed [24]:
lP

68 τ = 𝜏0 + 𝜇𝛾̇ + 𝑐𝛾̇ 2 (2)

69 Compared to the Bingham model, the modified Bingham model introduced a quadratic term related
na

70 to shear rate and its coefficient 𝑐 is the pseudoplastic constant. Feys et al. [25] applied rheological
ur

71 index 𝑐/𝜇 to depict non-linear behavior of the fluid which undergoes shear thinning when 𝑐/𝜇<0,

72 and shear thickening when 𝑐/μ>0. Besides, the Herschel-Bulkley (H-B) model was proposed in
Jo

73 1926 and its expression is [26]:

74 τ = 𝜏0 + 𝐾𝛾̇ 𝑛 (3)

75 where 𝐾 is the consistency coefficient and 𝑛 is the fluidity index that represents the degree of

76 deviation from the rheological behavior of Newton fluid [27]. When n<1, the fluid exhibits shear

77 thinning behavior, and shear thickening occurs in the case of n>1.

78 Published literatures showed that the Bingham model, modified Bingham model and H-B model

79 are empirically adopted to the description of rheological behavior of AAMs. There is suitability

80 difference for various systems. It is widely accepted that the alkalinity of alkali activators controls

81 the dissolution behavior of precursor particles, and the alkali cations in activators work as charge

82 balancing cations. Moreover, the polymerization status and the adsorption on precursor surface of
83 anionic groups from activator solution, silicate species in particular, could influence particles

84 interaction. These mechanisms determine the effects of activators on the rheological behavior of

85 AAMs. Generally, Bingham model is recommended for NaOH-activated pastes [41, 135], while H-B

86 model suits better the sodium silicate-activated pastes [19, 25]. For example, as active silicates are

87 added in calcium-rich (slag) precursor suspensions, the initial and speedy formation of a primary

88 C,N-A-S-H gel and mixed N-A-S-H makes the alkali-activation reaction more complicated [113].

89 This primary gel may be destroyed and restructured by subsequent shearing, leading to a deviation

90 from Bingham fluid. However, sometimes adding admixtures/additions would change the

91 interactions between particles, so affect the rheological behavior of sodium silicate-based pastes, thus

f
oo
92 conforming to the Bingham model [37-40]. Furthermore, Bingham model seems to suit both NaOH

r
93 and sodium silicate-activated mortar/concrete and this behavior is similar to PC based

94
-p
mortar/concrete [20, 31, 36]. Both Bingham model [41, 128] and H-B model [64, 129] were reported
re
95 to be used in the systems activated by Na2CO3 (single or mixed with other activators). The modified
lP

96 Bingham model not only describes the non-linear behavior deviating from Bingham fluid, but also

97 avoids a variable dimension parameter like in H-B model without mathematical limitation in the low
na

98 shear rate region [22, 25, 126]. However, it is rarely used in describing the rheological behavior of
ur

99 AAMs [49, 63]. The applicability of the modified Bingham model in different AAM systems needs

100 further study.


Jo

101 In addition to the activators used, the origin and composition of the starting aluminosilicates also

102 have great impact on the rheological behavior. Alkali-activated MK systems generally exhibit higher

103 viscosity and apparent yield stress due to their plate-like particles and large specific surface area [45,

104 74], which means that suitable activators and sufficient water are required to ensure good workability

105 [29, 47]. In the very early stage of Na-silicates activated MK pastes, the formation of a gel with

106 Si/Al<4.5, which strongly depends on the chemical composition of mixture, would be accounted for

107 the evolution of storage modulus, instead of colloidal interactions [34]. If Ca-rich precursors are used,

108 the chemistry of the system could be dramatically changed and the colloidal interactions between

109 grains could be enhanced as well [34]. As reported, alkali-activated fly ash pastes behaved like a

110 Bingham fluid [28, 37], while the rheological behavior of the fly ash-slag blended systems was more
111 consistent with H-B model [128, 136]. Besides, the area of thixotropy was significantly affected by

112 fly ash/slag ratio, although there were differences on its influence on thixotropy in different studies

113 [128, 138]. These rheological behaviors could not be separated from the differences of particle size,

114 dissolution behavior and reactivity of different precursor particles, as well as the initial precipitated

115 gels. The effects of composition factors on rheology of AAMs are discussed in detail in the following

116 sections.

117 3. Effect of alkaline activators on rheology of AAMs

118 It is known that alkaline nature and dosage significantly affect the reaction process, microstructure

f
oo
119 and properties of AAMs. Therefore, activator is a key composition controlling the rheological

r
120 behavior of AAMs paste. According to Glukhovsky [50], chemical activators can be classified into

121
-p
six categories in terms of chemical composition: caustic alkalis, non-silicate weak acid salts, silicates,
re
122 aluminates, aluminosilicates, and non-silicate strong acid salts. Among these activators, sodium
lP

123 based-activators are the most widely available and cost effective, such as Na2O·nSiO2, NaOH,

124 Na2CO3, Na2SO4 and a mixture of two or more. In addition, some potassium based-activators have
na

125 been used in laboratory studies [51]. This section discusses the effects of Na/KOH, Na2O·nSiO2 and
ur

126 some unconventional activators on rheology of AAMs.


Jo

127 3.1 Na/KOH

128 3.1.1 Ion nature

129 The most used alkali hydroxides in AAMs are NaOH and KOH. NaOH in particular, is usually

130 prepared into solution in advance in order to avoid the influence of dissolution heat on the reaction.

131 Published literatures have shown that paste suspensions containing NaOH solution usually exhibits

132 higher yield stress and plastic viscosity than KOH-based suspensions [28, 52]. As seen in Fig. 1, the

133 yield stress and plastic viscosity of fly ash paste activated by NaOH solutions are higher than those

134 of activated by KOH with the same dosage in terms of molarity [28]. This is due to the lower charge

135 density of K+, which gives lower ion-dipole forces and viscosity of solution [32]. Similarly,

136 irrespective of activator dosage, the Zeta potential value of KOH-activated slag system is higher than
137 that of NaOH-based system (Fig.2), due to lower concentration of Na+ cations adsorbed on the

138 surface of negatively charged particle [53]. More adsorption of K+ cations might reduce the van der

139 Waals forces and increase the repulsive force of double layer between charged particles, and thus

140 contributes to a decrease in yield stress. By contrast, Na+ cations would rather to combine with water

141 and consume more free water.

142

f
r oo
-p
re
lP

143
na

144 Fig. 1. The yield stress (a) and plastic viscosity (b) of fly ash paste activated by NaOH and KOH solutions of

145 different concentrations [28].


ur

146
Jo

147
148 Fig. 2. Zeta potential of 3 wt.% slag suspensions with various dosages of NaOH, KOH activators [52].

149
150 3.1.2 Concentration

151 Concentration is another important factor that impacts rheology of AAMs. Table 1 summarizes the

152 effects of NaOH concentration on the rheological properties of AAMs prepared with various

153 precursors. Rifaai et al. [49] raised NaOH concentration from 2 to 7 mol/L, and found a rise in yield

154 stress of alkali-activated slag pastes. However, Zhang et al. [54] found that higher NaOH

155 concentration (4.17~6.67 mol/L) of alkali-activated fly ash-slag blended pastes had shown lower

156 yield stress and plastic viscosity, thus increasing the fluidity of pastes (Fig. 3), which was similar to a

157 later study by the same group [135]. For the paste with high NaOH concentration, the value of Zeta

f
oo
158 potential is more negative (−7.28 mV) due to the high dissolution and ionization degree of

159 suspension [54]. The result is somewhat equivocal and unexpected. It is widely accepted that the

r
160 -p
dissolution of slag is faster than fly ash due to its higher reactivity. The ionic bond formed by
re
161 modifier cations (such as Ca2+) and the non bridging oxygen in the calcium aluminosilicate glass

162 structure is more easily dissolved, resulting in the release of Ca2+ ions from slag in the early stage
lP

163 [124]. The Zeta potential become less negative due to the re-adsorption of Ca2+. However, silicate
na

164 and aluminate may be partially replaced by low-charge cations. Therefore, the measured Zeta

165 potential is the result of the mutual compensation between this two mechanisms, which may be
ur

166 related to the dilution of suspension and the reaction time. Further, in the early stage, the
Jo

167 water-binding ability of dominated N-A-S-H in fly ash-slag blended system was weaker than that of

168 C-A-S-H gels in slag system, thus the higher freedom of water in the mixed system [135]. But this

169 effect may be less significant than that of electrostatic repulsion. Therefore, different results were

170 obtained in the similar range of NaOH concentration [49, 54, 135], which may be related to the

171 precursor nature and the water to binder ratio.

172

173 Table 1

174 Effects of NaOH concentration on the rheological properties of AAMs prepared with various precursors.
Critical
w/b or L/S NaOH concentration
Precursors concentration Main features Ref.
(by mass) (mol/L)
(mol/L)
Fly ash
τ0 and μ
(Mortar:
0.5-0.6 8; 10 - increased as concentration [31]
sand/binder
increased
=2:1)
Fly ash τ0 slighthly decreased as
0.25 8; 10 - [55]
concentration increased
τ0 and μ increased as
0.35 4; 8 - [28]
concentration increased
τ0 and G' increased as
concentration (< 7 mol/L)
increased; μ increased as
2; 3.5; 6; 6.5; 7; 7.5; concentration increased
0.4 7 [49]
8; 11; 14 until 3 mol/L, then

f
oo
decreased up to 8 mol/L,
and increased again

r
afterwards
Slag
0.45 1; 2
-p -
τ0 increased as
concentration increased
[52]
re
τ0 decreased as
concentration (< 3.37
lP

0.55 2.57; 3.37; 4.13 3.37 mol/L) increased, then [41]


increased; μ increased as
na

concentration increased
Slag + fly The growth rate of G'
0.83; 1.67; 2.5; 3.33;
ur

ash 0.6 4.17 increased as concentration [56]


4.17; 5; 5.83; 6.67
(< 4.17 mol/L) increased
Jo

0.83; 1.67; 2.5; 3.33; τ0 and μ decreased as


0.6 - [54]
4.17; 5; 5.83; 6.67 concentration increased
τ0 and μ decreased as
0.5 1; 2; 3; 4; 5; 6; 7; 8 - [135]
concentration increased
Viscoelastic transition
2.22; 4.44; 6.67;
Phosphorus was delayed with the
0.45 8.87; 11.11; 13.33; 8.87 [57]
slag increase of concentration
15.56
(< 8.87 mol/L)

175 * w/b = water/binder ratio; L/S = liquid/solid ratio.

176 * In reference [31], τ0 and μ are calculated based on the speed-torque curve.

177
178

f
oo
179 Fig. 3. The yield stress (a) and plastic viscosity (b) of fresh fly ash-slag blended pastes activated by different

180 concentrations of NaOH [54].

r
181 -p
re
182 Importantly, by reviewing the published works, there seems a critical value of NaOH

183 concentration alternating the rheological feature of AAMs, especially the development of
lP

184 viscoelasticity. OH- ions play a role as chemistry driven to stimulate solid aluminosilicate dissolution,
na

185 and Na+ ions are necessary to house in the voids in tetrahedral coordination structure to compensate

186 for the electrical charge due to the substitution of Si4+ by Al3+ [107]. Hence, ion concentrations
ur

187 influence the competition between dissolution and polycondensation kinetics, thus affecting
Jo

188 viscoelasticity [49, 56, 57]. A recent research (Fig. 4) reported that NaOH concentration of 7 mol/L

189 was a critical value, below which the yield stress, storage modulus (𝐺 ′ ) and rigidification rate of

190 alkali-activated fly ash (AAFA) pastes increased as concentration increased, while above which the

191 rheological properties showed a downward trend [49]. A reasonable explanation for this phenomenon

192 is the competition between dissolution and polycondensation. On the one hand, lower NaOH

193 concentration can accelerate both the dissolution and polycondensation kinetics. On the other hand,

194 the dissolution rate is more dominated in extremely alkaline environment, which induces more

195 negatively charged monomers leading to a remarkable repulsion between particles [49]. However,

196 the nature of precursors may also impact the critical value of NaOH concentration. For example, the

197 fly ash-slag blended suspension showed a critical value of 4.17 mol/L (the mass ratio of

198 NaOH/binder was 10%) for NaOH concentration, which might be related to the changed surface
199 charge of particles and initial dissolution behavior [56]. As discussed earlier, the higher CaO content

200 and lower degree of glass polymerization in slag are more conducive to dissolve, thus generating

201 higher reaction rate than FA [124]. This means that the introduction of slag can accelerate the

202 dissolution rate of overall raw materials, compared with the pastes containing FA as single precursor.

203 Hence, the alkalinity needed to achieve the equilibrium of dissolution and precipitation is reduced,

204 that is, the critical NaOH concentration is reduced. Furthermore, although an increase in NaOH

205 concentration promotes the formation of Si-O-Na bond, excessively high concentration is

206 unfavorable due to early hydration product precipitation on surfaces of particles. Xie et al. [57]

207 pointed out that 8.87 mol/L NaOH (the mass ratio of NaOH/phosphorous slag was 16%) was a

f
oo
208 threshold and the influence of NaOH concentration changed near this threshold on viscoelasticity.

r
209 Below this value, the viscoelastic transition (the transition from the viscous behavior to the elastic

210
-p
behavior is the dominant feature) of alkali-activated phosphorous slag (AAPS) was delayed with the
re
211 increase of NaOH concentration. Low hydration rate and good dispersion because of the enhanced
lP

212 electrostatic repulsive force may be responsible for this phenomenon. Above this threshold, the

213 viscoelastic transition accelerated with the increase of NaOH concentration, which was attributed to
na

214 the rapid generation of cross-linked gels and the decreased electrostatic repulsive force [57].
ur

215 Unfortunately, no rheological research on metakaolin activated by NaOH solution has been reported

216 so far.
Jo

217

218
219 Fig. 4. Effects of NaOH concentration on storage modulus and rigidification rate of alkali-activated fly ash pastes.
220 Data from Rifaai et al. [49].

221 3.2 Na/K-silicates

222 3.2.1 Ion nature and ways of adding activators

223 AAMs activated with sodium silicates usually exhibits greater workability, i.e. with lower yield

224 stress, than that of PCMs and alkali hydroxide activated system due to plasticizing effect [31, 52, 55].

225 For example, Kashani et al. [52] reported that the initial Zeta potentials of Na/KOH-slag system and

226 sodium silicate-slag system are both negative due to the deprotonation of slag silanol groups.

f
oo
227 However, as activator concentration increase, the Zeta potential of Na/KOH-slag system changes

228 from negative to positive, which means that repulsive double layer forces decreases and then

r
229 -p
increases. By contrast, the Zeta potential of sodium silicate-slag system becomes more negative [52].
re
230 This is due to that silicate oligomers adsorbing on the surfaces of slag particles, resulting in

231 insufficient dissolved Ca2+ in the solution and greater repulsive double layer forces. As expected,
lP

232 lower yield stress was also obtained in Na-silicate based fly ash pastes instead of NaOH-fly ash [28,
na

233 55]. The overall negative surface charge of fly ash was increased by the adsorption of silicates from

234 activator solution, which enhanced inter-particle repulsive forces and reduce flocculation of particles
ur

235 [55]. The exact electric charge density of the surface of raw materials will depend on the polymer
Jo

236 status of silicate oligomers, being monomers, dimers or trimmers, and the original surface charge.

237 The previous studies of sodium silicates-slag suspensions also showed obvious thixotropy feature

238 [19]. Thixotropy is usually observed in shear thinning fluids because that viscosity decreases with

239 time when a given shear rate is applied to the material at rest, and viscosity would recover once the

240 flow stops [114]. This process corresponds to the destruction and restruction of thixotropic materials,

241 respectively, which is strongly dependent on the flow history and rest time [23]. Thixotropy can be

242 described by a thixotropic hysteresis loop (qualitative) [23] and thixotropic index (quantitative) [114].

243 Fig. 5 shows the thixotropy of slag pastes activated with sodium silicates and NaOH, respectively. It

244 can be seen that sodium silicates-activated slag paste showed a larger area of the hysteresis loop and

245 a higher shear stress than that of NaOH system, especially at the first cycle [19]. This is mainly

246 attributed to the initial and rapid formation of primary C-A-S-H gels, which generate colloidal forces
247 between particles [58, 113]. It is important to note that this colloidal force can be interfered by

248 continuous shearing, thus a prolonged mixing will improve workability of sodium silicate

249 activated-slag concrete [20]. However, neither Na-silicate nor NaOH activated fly ash pastes showed

250 obvious thixotropic behavior [55, 128]. This may be attributed to the slower dissolution rate of fly

251 ash particles and lower precipitation rate geopolymeric gels (slower structure build-up) [128, 136].

252 On the other hand, the spherical particle feature of fly ash could be another important factor.

253 Interestingly, Poulesquen et al. [32, 73] conducted a strain sweep test to investigate the influence of

254 silicate activating solutions synthesized by Na/KOH and two types of silica on the viscoelastic

f
255 parameters (storage modulus G’; loss modulus G’’ and tanδ=G’’/G’) of metakaolin (MK) systems

oo
256 during the polymerization process. Though Na-based activator shows weaker basicity, it can

r
257 unexpectedly ensure a better dissolution of MK and exhibit faster geopolymerization. Smaller Na+
-p
258 with higher charge density strongly attract and retain their hydration layer of water molecules, while
re
259 K+ show a weaker association with water [32, 137]. In other words, K + ions are more likely to
lP

260 combine with negatively charged silicates, exhibiting an interfering effect on the condensation
na

261 process. However, compared to the Na-based system, smaller tanδ in K-based system was observed

262 due to stronger interactions between the dissolved Si and Al tetrahedral monomers, which indicated a
ur

263 more rigid structure [32, 137]. Additionally, the viscoelasticity of MK pastes in their study was not
Jo

264 sensitive to the type of silica in synthetic silicate activation solution, but the activator synthesized by

265 smaller silica particles with NaOH reduced setting time evidently.

266
267
268

f
Fig. 5. Hysteresis cycles for alkali activated slag pastes [19].

oo
269

r
270 It is worth mentioning that activators in this paper are all added in the form of solutions, if not

271
-p
otherwise specified. Recently, there has been particular interest in one-part AAMs, which are
re
272 produced by mixing solid activators and precursors [125]. Compared with liquid sodium silicate,
lP

273 there is a higher yield stress and plastic viscosity observed when solid sodium silicate is used. The

274 dissolution reaction of solid sodium silicate led to less free water acting as lubricant and the
na

275 additional dissolution heat can improve the formation and size of gels in slag-fly ash system [38].
ur

276 Such rheological characteristics seem to make solid sodium silicate more suitable for preparing

277 extrusion-based 3D printing mortars than liquid sodium silicate [99]. Moreover, some studies have
Jo

278 focused on the rheological behavior and fluidity of one-part AAMs with different raw materials [86,

279 106, 126-127]. Unfortunately, these studies did not specifically discussed the differences in rheology

280 properties between one-part AAMs prepared by solid activators and conventional two-part AAMs.

281 More research is therefore needed to explain rheology of one-part AAMs or the effect of adding

282 activators in different states (solid or solution) in the future.

283 3.2.2 Modulus (nSiO2/nNa2O)

284 The addition of NaOH could raise the pH value and reduce modulus (the SiO2/Na2O molar ratio)

285 of sodium silicate activator solution, affecting the distribution of different ions in the solution and

286 thus the rheological properties of AAM suspensions. This is directly related to the immediate
287 viscosity of activator solution. When the alkali content (in oxides form) is constant, adjusting the

288 modulus should consider more about effects of the polymerization degree of silicate oligomers and

289 their adsorption on particle surface on the rheological properties.

290 Mehdizadeh and Kani [35] found that the yield stress of AAPS declined as the modulus of sodium

291 silicate activator increased at constant Na2O/Al2O3 ratio of 1.0, which may be explained by the

292 plasticizing effect of silicates [52], though this molar ratio showed a negligible effect on the

293 shear-thinning parameter. Some similar results were reported in 3D printing pastes based on

294 alkali-activated blast furnace slag/steel slag [63]. It was observed that increasing modulus could

295 decrease the yield stress and its thickening rate, as well as plastic viscosity of fresh pastes. Therefore,

f
oo
296 for example, for 3D printing AAM concrete based on slag, to obtain satisfying thixotropic property,

r
297 reduced modulus is more favorable to ensure the stability between the layers of the printed materials.

298
-p
However, some recent researches reported that increasing the modulus of sodium silicates resulted in
re
299 an increase in apparent viscosity and yield stress of alkali-activated fly ash pastes [28]. Wen et al. [59]
lP

300 found that the viscosity of metakaolin-based AAMs increased dramatically with modulus when the

301 sodium silicates was in SiO2 polymerized zone (the modulus>2.2). In the high-ionized zone (the
na

302 modulus<1.8), however, the viscosity of suspensions showed insignificant responses to modulus
ur

303 variation [59].

304 In slag-fly ash blended pastes (fly ash/slag=5:5), the yield stress and the area of thixotropy
Jo

305 decreased by approximately 76% and 83%, when modulus increased from 0.4 to 1.6 [136]. Similar

306 results were reported in a grout of fly ash-slag (fly ash/slag=9:1) blended system, plastic viscosity

307 decreased as modulus increased from 1.4 to 1.8 [46]. This might be attributed to the high alkalinity

308 and high degree of ionization in suspensions of lower modulus (1.4) at high liquid-solid ratio, which

309 promoted the dissolution and consequent polymerization, thus resulting in higher plastic viscosity.

310 Additionally, Dai et al. [136] found that the effects of low modulus (0<Ms≤0.4) and high modulus

311 (0.4<Ms≤1.4) on evolution of structure build-up were quite different. As shown in Fig. 6, mixtures

312 with lower modulus behaved a higher initial G’, and subsequently structure build-up rate increased

313 slowly, while high modulus led to a delay in structure build-up, followed by a rapid increase [136].

314 Low modulus activator, by contrast, showed higher alkalinity and promoted the early
315 dissolution-precipitation in mixtures, resulting in higher initial G’. Consequently, precipitated gels on

316 the surface of precursor particles might hinder further reactions [136]. Moreover, Hasnaoui et al.

317 [140] reported that increasing modulus (from 1.0 to 2.0) induced a rise of viscosity in alkaline

318 solution, which led to increase of the yield stress and the plastic viscosity of slag-MK mortars, thus

319 reducing workability. This effect is pronounced in the mixture with more MK, which is partly due to

320 the high viscosity of MK-based pastes itself.

f
r oo
-p
re
321
lP

322 Fig. 6. Effect of n(SiO2)/n(Na2O) molar ratio on normal scale(a) and log scale (b) storage modulus evolution of

323
na

alkali-activated slag-fly ash pastes (slag/fly ash=1) using small amplitude oscillatory shear test at strain of 0.005%,

324 frequency of 1 Hz and 20 °C [136].


ur

325
Jo

326 Based on these findings, the relationship between the modulus and yield stress and plastic

327 viscosity of AAMs in different literatures is shown in Fig. 7. In general, the increase of modulus is

328 helpful to reduce yield stress of AAMs thanks to the plasticizing effect when the modulus of

329 activator is below around 1.2 with a relatively high degree of ionization. However, higher modulus

330 (>1.4) seems not conducive to the reduction of yield stress (Fig. 7a). In addition to the strong

331 increase of ion-dipole forces, a rise of modulus also induced more colloidal HO-Si-OM complexes in

332 activator solution forming clusters with several nanometers, and thus the polymerization degree of

333 silicate anions raises. Further, it must consider the effect of early C-(A)-S-H gels precipitate by Ca2+

334 ions and silicate species on rheology in calcium-containing systems [113]. Higher modulus induced

335 more active [SiO4]4- react with dissolved Ca2+ ions, producing more hydration products [46].

336 Therefore, the viscosity of suspension increased [60] and the complex network structure of highly
337 polymerized silicates is not easy to be damaged by shearing [46]. Theoretically, the viscosity of

338 interstitial fluid (silicates) has immediate impact on the viscosity of AAMs. The plastic viscosity

339 would ordinarily increase as modulus, due to more colloidal clusters (Fig. 7b). It should be noted that

340 plastic viscosity is the embodiment of internal friction between substances in suspensions, and also

341 depends on liquid-solid ratio and the structure of gel network (the precipitation of gels or

342 polymerization of silicate oligomers).

343

f
r oo
-p
re
lP
na

344
345 Fig. 7. Effect of n(SiO2)/n(Na2O) molar ratio of sodium silicate activator on yield stress (a) and plastic viscosity
ur

346 (b) of AAMs at the first shear cycle [35, 41, 46, 59, 61, 62, 63, 136, 140].
Jo

347

348 3.2.3 Concentration

349 To provide a guidance for the application requiring better workability and fluidity, some studies

350 have focused on the effect of sodium silicate concentration on the rheology of high-calcium AAMs.

351 At constant modulus of 1.0, Vyšvařil et al. [42] observed that the increase of dosage (higher

352 concentration) of sodium silicates could allow activated-brick powder pastes to flow more easily, but

353 unexpectedly, the long term strengths of hardened pastes were slightly reduced at the highest dosage

354 of activator. Surprisingly, Kashani et al. [52] noticed that higher concentration of sodium silicate

355 induced greater negative Zeta potential, which does not always mean lower yield stress value of slag

356 based suspension, as can be seen in Fig. 8. The yield stress is lower (12.5 Pa-17.5Pa) when the
357 square of Zeta potential is between 40-80 mV2, which means that this sodium silicate concentration

358 was conducive to disperse particles compared with the blank group. Surprisingly, the yield stress

359 increases although the square of Zeta potential are further increased to 150 mV2, which indicates that

360 the attraction and flocculation increases at higher sodium silicate concentration. Fig. 9 well illustrates

361 this phenomenon: in the case of lower concentration of sodium silicates (2.2 and 4.4×10-4 mole

362 Na2O/g slag), the increase of repulsive forces reflects the plasticization and deflocculation effects of

363 [SiO4]4- groups (Fig. 9b). However, a further increase in the amount of Na2O brings about a rise in

364 yield stress (6.6×10-4 mole Na2O/g slag), which indicates that the electric double layer force could

365 exhibit a sign of attractiveness, promoting flocculation and leading to higher yield stress [52, 55]

f
oo
366 (Fig. 9c). The effect of Na2O% on yield stress was complex in sodium silicates-slag systems. With

r
367 relatively low modulus (<0.8) (Fig. 10), the fact that the rise in Na2O% led to an increase in pastes

368
-p
yield stress was consistent with the discussion above. Suspensions, however, did not show the similar
re
369 behavior when increased Na2O% at higher modulus (>1.2) of Na-silicates (Fig. 10). This might be
lP

370 ascribed to the fact that modulus needed for primary C–(A)–S–H gel to form and breakdown in

371 dynamic shearing conditions declined as Na2O% rose [41]. More detailed explanations are needed in
na

372 future studies.


ur

373 For some blended systems and calcium-free systems, the effects of Na/K-silicate concentration

374 (Na/K2O%) on the rheological properties, especially the viscoelasticity (structural build-up kinetics),
Jo

375 have been studied. Dai et al. [136] reported that increasing sodium silicate dosage (3%-6% of N2O%)

376 in slag-fly ash paste caused a delayed structural build-up followed by an abrupt increase, which was

377 related to the longer initial setting time. Interestingly, they observed humps in loss factor evolution

378 curve during the rapid increasing structural build-up stage with higher Na-silicate concentration,

379 which indicated that the loss modulus (G’’) increases faster than the storage modulus (G’) for a given

380 time. This phenomenon could be explained by the release of extra water from the condensation of

381 Si(OH)4 species, leading to an increase of loss factor and a more liquid-like behavior [136]. These

382 humps were also related to the gelation [137]. Furthermore, the dynamic yield stress, thixotropy and

383 the shear-thickening response decreased as Na-silicate dosage increased, and even behaved like a

384 Bingham fluid. In a system of calcined shale (CS) as a metakaolin source, the complex viscosity
385 (obtained by small amplitude oscillatory) for describing hardening rates of pastes activated by

386 different amount of K-silicate was measured [138]. The results showed that increasing activator

387 resulted in a delay of saturation-polycondensation stage and reducing hardening rate and matrix

388 viscosity by the increase of liquid-solid ratio [138].

389

f
r oo
-p
re
lP

390

391
na

Fig. 8. Yield stress of sodium silicate activated slag suspensions with a water-binder ratio of 0.45 as a function of

392 the square of Zeta potential at different activator dosages (0, 2.2, 4.4 and 6.6×10-4 mole Na2O/g slag) [52].
ur
Jo

393
394 Fig. 9. A schematic sketch of electrostatic inter-particle forces with different dosages of sodium silicate in slag

395 paste, adapted from [52].

396
397

f
oo
398 Fig. 10. Effect of Na2O% and n(SiO2)/n(Na2O) molar ratio on 11-min (for the second shear cycle) yield stress in

399 sodium silicate activated slag pastes. Data from [41].

r
400 -p
re
401 3.3 Other activators
lP

402 In addition to the widely used activators discussed above, there are some unconventional
na

403 activators, such as carbonate, sulphate and phosphate. In a calcium-free MK based system, Wang et

404 al. [131] used orthophosphoric acid and liquid monoaluminum phosphate to prepare different Al/P
ur

405 molar ratios (0, 0.1 and 0.3) of phosphate activating solution. It was observed that more aluminum
Jo

406 species from phosphate activating solution favors to reduce the yield stress of MK based pastes due

407 to a faster sol/gel transition. Similarly, in a low-calcium paste with fly ash/slag ratio of 4, the yield

408 stress and viscosity of the control group (without borax) can be effectively reduced by partially

409 replacing borax for sodium silicate solution owing to partial structural breakdown of flocs [139]. In

410 addition, borax composite behaved like a Bingham fluid while the rheological behavior of control

411 group is more consistent with the H-B model. For now, though, more studies on the rheology of

412 high-calcium systems activated by other activators were reported. When using a combination of

413 NaOH, Na2CO3 and waste glass (SiO2, Na2O, CaO, etc.) to activate slag, Torres-Carrasco et al. [64]

414 observed that the rheological behavior of AAS pastes was similar to that of sodium silicates-activated

415 slag pastes, which can be characterized by the H-B model. The other important finding in their work

416 is that “false setting” arose in both systems. This can be attributed to the rapid precipitation of
417 primary C-(A)-S-H gels. It should be noted that longer time is required for activating the solution

418 containing waste glass to generate primary C-(A)-S-H gels because of the difference of silicon

419 source and environment, where the potential retarding effect of the mixed activator bearing waste

420 glass is observed [64].

421 Recently, Mehdizadeh et al. [65] reported the rheological behavior and fluidity of phosphorus slag

422 activated by two mixed activators (Na2CO3+ Ca(OH)2, Ca(OH)2+Na2SO4). The presence of CO32-

423 ions and concomitant formation of AFm and calcium sodium carbonate-type compounds in

424 alkali-activated slag pastes would like to intensify the interactions between flocs and thus generated

425 higher shear stress [41, 129]. In addition, when the ‘designed’ NaOH concentration increases from 3%

f
oo
426 to 5%, the measured shear stress and yield stress values increased. For the phosphorous slag pastes

r
427 activated by Ca(OH)2 and Na2CO3 (PS/CNC), as shown in Fig. 11a, the concentration of main

428
-p
elements, especially Si, increased significantly in the first 35 min, indicating that the reaction
re
429 between Ca(OH)2 and Na2CO3 enhanced the alkalinity of solution and promoted the dissolution of
lP

430 phosphorous slag. Hence, the high yield stress of the PS/CNC was attributed to the formation of

431 more early gels as Ca2+ reacted with dissolved silicates (the yield stress started to increase after 15
na

432 min of mixing in Fig. 12). This can be confirmed by the decreased concentrations of these ions after
ur

433 35 min. Nevertheless, a lower yield stress in the Ca(OH)2 and Na2SO4 activated phosphorous slag

434 pastes (PS/CNS) pastes was detected (Fig. 12), which was probably caused by the weak reaction of
Jo

435 Na2SO4 and Ca(OH)2. As shown in Fig. 11b, concentrations of the leached Si and Al in PS/CNS were

436 very low compared to those in the PS/CNC. Therefore, it can be deduced that lower alkalinity and

437 presence of SO42− mean low activation capability, which also leads to less early gels. This provides

438 an approach for designing a robust formula of AAMs with good fresh performance. For Na2CO3-slag

439 pastes, increasing activator concentration from 6% to 10% showed no significant effect on yield

440 stress and plastic viscosity because of its long setting time [129]. The addition of Ca(OH)2 can

441 accelerate reaction effectively by consuming free CO32− ions. By contrast, it is a good method to

442 replace Na2CO3 with a small amount of NaOH (20 wt.%) by which setting time could be shorten and

443 the rheological parameters yield stress and plastic viscosity only slightly increased [129].

444 Furthermore, Alrefaei et al., designed a slag-fly ash paste activated by one-part Ca(OH)2/Na2SO4
445 exhibited a higher fluidity, but rather the conventional Na2SiO3-anhydrous activator [132]. This

446 one-part Ca(OH)2/Na2SO4 has also been used to activated fiber-reinforced AAS composite, and

447 obtained good workability [106].

448

f
r oo
-p
re
449
450 Fig. 11. Concentration variation of main ions in the leached solutions from (a) Ca(OH)2 + Na2CO3 activated
lP

451 phosphorous slag paste (PS/CNC) and (b) Ca(OH)2 + Na2SO4 activated phosphorous slag paste (PS/CNS) [65].
na

452
ur
Jo

453
454 Fig. 12. The yield stress of phosphorus slag activated by Ca(OH)2 + Na2CO3 (PS/CNC) and Ca(OH)2 + Na2SO4

455 (PS/CNS), with a ‘designed’ NaOH concentration of 3% and 5% [65].

456
457 4. Effect of precursors on rheology of AAMs

458 AAMs cover a wide variety of materials whose differences depend on the physical and chemical

459 properties of the starting aluminosilicates. Extensive researches have been focused on the

460 understanding of the alkali-activation reaction with various precursors under various curing

461 conditions and their products. Inspired by these achievements, a few studies have been carried out

462 about rheological behavior with different precursors, such as slag, fly ash and their blends [40, 65,

463 66]. According to the calcium concentration of precursors [95], AAMs can be divided into three

464 categories: ① low-calcium system, generated from those precursors like metakaolin (CaO%<1%)

f
oo
465 or calcium-poor fly ash (usually Class F, 1%≤CaO<10%), is dominated by highly cross-linked

466 aluminosilicate geopolymeric gels (N-A-S-H); ② high-calcium system based on calcium rich

r
467
-p
precursors such as blast furnace slag (10%≤CaO) with tobermorite-like calcium aluminosilicate
re
468 hydrate gels (C-A-S-H) as the dominant products; ③ alkali-activated blended system with

469 calcium-rich and calcium-poor precursors, containing coexisted amorphous aluminosilicate


lP

470 geopolymeric gels and C-A-S-H products [8, 67, 68]. With a constant Al/Si of 0.5, N-A-S-H were
na

471 found to coexist with C-A-S-H gels when Ca/Si<0.6, but only C-A-S-H gels were formed when

472 Ca/Si≥0.6 [96]. The rheological behavior and parameters of AAMs are substantially different when
ur

473 different precursors are used.


Jo

474 4.1 Chemical and physical properties of precursors

475 For calcium-rich precursors such as blast furnace slag, phosphorous slag and Class C fly ash, they

476 are usually used as a single precursor alone in AAM manufacturing because of their high content of

477 reactive components. Gonzalez et al. [70] used 0.25 M NaOH solutions to activate a modified

478 basic-oxygen-furnace slag incorporated by 13 wt.% Al2O3 (reactive amorphous: 55.5 wt.%) and

479 found that the plastic viscosity of the paste was about 2 times higher than that of the modified slag

480 induced by 11 wt.% Al2O3 and 5 wt.% SiO2 (reactive amorphous: 65.5 wt.%). The results mean that

481 the total amount of reactive component of a specific calcium-rich precursor cannot determine its

482 reactivity. The chemistry of the reactive component seems more important. In a Class C fly

483 ash-waste brick powder blended high-calcium system, it was found that increasing content of
484 calcium-rich waste brick powder tended to increase yield stress and consistency coefficient of pastes

485 activated by Na-silicate [43], due to the fact that more divalent cations (Ca2+) could lead to attractive

486 ion correlation forces and more primary gels [42]. Poor dispersion of brick powder (contained some

487 concrete powder) also led to higher solid volume fraction in the suspension, which induced the

488 increase of rheological responses [43]. Moreover, it was reported that the AAPS pastes (PS was

489 almost completely amorphous, and mainly composed of 45.16 wt.% CaO, 38.39 wt.% SiO2 and 7.66

490 wt.% Al2O3) exhibited as a shear thinning behavior where flow index (n, can be determined by Eq. (3)

491 in section 2) and thinning index (TI, the ratio of the apparent viscosity at low rotational speed to that

492 at rotational speed ten timer higher according to ASTM D2196) were independent upon the chemical

f
oo
493 composition of activator, implying that such shear thinning mechanism was more associated with the

r
494 characteristics of phosphorous slag [35].

495
-p
Class-F Fly ash (FA) is common used as calcium-poor precursor, and their reactivity lies on
re
496 content and composition of vitreous phase(s). In the alkali-activated blended system dominated by
lP

497 low-calcium fly ash, the spherical and fine fly ash particles exhibited beneficial impacts on the

498 rheological properties. Aboulayt et al. [47] studied the alkali-activated metakaolin-fly ash (MK-FA)
na

499 blended system. The maximum packing volume fraction (𝜙𝑀 ) of MK-FA blended AAMs grout
ur

500 increased with the replacement level of MK by FA, which was due to dense packing model of

501 spherical FA particles. Furthermore, the specific surface areas of FA is usually much lower than MK,
Jo

502 and the replacement by FA makes more surface wetting water being free. Thus, in overall, the

503 viscosity (consistency) was tested to be reduced and the setting time as well as the risk of bleeding

504 increased. Yang et al. [40] proposed to use fly ash microspheres (FAM) as a novel inorganic

505 dispersant based on its “ball-bearing” effect. The SEM image in Fig. 13 shows the morphology of

506 nanoscale FAM particles with smooth surface and high sphericity. This morphologic feature

507 facilitates the usage as dispersing agent. When the fly ash replacement ratio by FAM increased, the

508 plastic viscosity for all samples declined, irrespective of the modulus (Ms) of sodium silicate (in

509 Fig.14a). This phenomenon was ascribed to the “ball-bearings” effect, together with the lubricating

510 effect of FAM particles which breaks down the interlocking between flocs and fragmentation [40,

511 71]. Apart from the packing density, the yield stress appeared to be more susceptible to the activator
512 modulus (in Fig.14b). In the Ms=1.4 system , only 40% FAM reduced the yield stress of suspensions,

513 which could be inferred that the ‘ball-bearings’ of FAM destroyed the interlocking between fly ash

514 and slag particles to resist the increased particle packing [40]. By contrast, ‘ball-bearings’ effect in

515 the Ms =1.8 system seems less important, since more colloidal silicate species might react with Ca2+

516 released from slag to produce more primary gels, and thus increasing the attraction between

517 suspended particles. Again, it should be noted that particles with larger specific surface area need to

518 consume more water, therefore, there is a balance between the three effects of packing density,

519 surface water absorption and ‘ball-bearings’.

520

f
r oo
-p
re
lP
na

521
522 Fig. 13. The SEM image of fly ash microspheres used as inorganic dispersant in AAMs [40].
ur

523
Jo

524
525 Fig. 14. The plastic viscosity (a) and yield stress (b) of the alkali-activated fly ash-slag blended pastes prepared by

526 Ms=1.4 and 1.8 sodium silicate solutions. At constant slag content of 40 wt.%, the 60 wt.% fly ash is replaced by 0,
527 10, 20, 30 and 40 wt.% fly ash microspheres, denoted as AM0, AM10, AM20, AM30 and AM40, respectively (data

528 from [40]).

529

530 Effects of fly ash/slag ratios on rheology of alkali activated slag-fly ash composite systems have

531 always been a research hotspot. Introducing different contents of slag directly affects the initial

532 dissolution behavior, and higher calcium content leads to changes of the precipitated gel systems.

533 Compared with FA paste activated by a composite activator (NaCO3 and Na-silicate), a replacement

534 of slag to fly ash (fly ash/slag = 3:1; 4:1) reduced the yield stress of pastes. Similar to this finding,

535 Dai et al. [136] increased the replacement amount of slag to fly ash to 70wt.% in Na-silicate based

f
oo
536 pastes, the yield stress and consistency coefficient were reduced by 51% and 81%, respectively. This

r
537 is attribute to a reduction of flocculation caused by fine FA particles and a faster dissolution of slag

538
-p
resulting in low solid content [128, 136]. Besides, the AAFA pastes behaved like a Bingham fluid,
re
539 while the rheological behavior of the composite pastes was more consistent with H-B model. Dai et
lP

540 al. [136] observed the hysteresis area increased as the slag/fly ash ratio in the composite pastes. The

541 other study, however, reported opposite results in the hysteresis area, and the yield stress of blending
na

542 paste with fly ash/slag ratio of 2 was almost doubled the AAFA [128]. Different w/b ratios and
ur

543 particle sizes may be responsible for these differences. Dai et al. [136] designed the w/b ratios to

544 decrease with the increase of FA content, and the fly ash used was finer than slag. This enhances
Jo

545 attractive inter-particle forces leading to more flocculation structures that are more difficult to be

546 disturbed by shearing. Similarly, Na-silicate based fly ash-slag mortars with copper slag as fine

547 aggregate, higher yield stress and consistency were obtained in the mixture containing more FA due

548 to a higher specific surface area FA [72]. On the contrary, it must be considered that the introduction

549 of more slag means faster dissolution and higher structure build-up rate, which was characterized by

550 shorter percolation time, faster storage modulus growth [136] and flow loss [141]. These results are

551 due to the formation of more early C-A-S-H gels [113].

552 The MK-based AAMs generally have higher rheological response because very high specific

553 surface area of MK [45, 74], which indicates that suitable activators [123], dispersing agents [74] and

554 sufficient water [29] are required to achieve good workability. As a calcium free precursor, sodium
555 silicate-MK pastes have been demonstrated that the formation of primary gels with Si/Al<4.5,

556 instead of colloidal interaction between grains, could be accounted for the evolution of early

557 mechanical properties (storage modulus) [34]. As discussed previously, it is very different from

558 Ca-containing systems: the presence of calcium causes a dramatic change of systemic chemistry and

559 the colloidal interactions between grains could be enhanced as well [34]. Moreover, compared to

560 cement-based suspensions, MK-based pastes usually exhibit high viscosity due to the fact that

561 plate-shape MK grains have low 𝜙𝑀 . In the sodium silicate based-suspension with a volume fraction

562 of MK around 0.25, the small volume of solid grains prevented the contact between grains though

563 the 𝜙𝑀 of MK was small, and thereby the suspension viscosity was nearly exclusively controlled by

f
oo
564 the viscosity of interstitial fluid [33]. Interestingly, the plastic viscosity of alkali-activated MK-based

r
565 suspensions could be reduced by increasing the replacement of MK by waste red brick powder [74].

566
-p
Similar results were also reported in MK-slag mortars activated by sodium silicate solution [64, 140],
re
567 as shown in Table 2. In addition to the negative effect of irregular spiny plate-shape particles of MK,
lP

568 the viscosity-enhancing effect of MK was attributed to the larger specific area, requiring more

569 mixing water for wetting. Moreover, MK with distorted aluminum layer structure had higher
na

570 reactivity than brick powder, hence, the suspensions containing more MK led to faster growth of
ur

571 yield stress and storage modulus due to rapid gelation [74]. Table 2 summarizes the effects of

572 precursor nature on the rheological properties of alkali-activated blended systems. It can be seen that
Jo

573 the physical properties of precursors, morphology in particular, jointly affect the surface water

574 absorption and packing density to alter the rheological properties of AAMs. In PCMs system, many

575 studies have been demonstrated that the combined effects of these factors could be well evaluated by

576 the water (paste or mortar) film thickness [108-110] whereas correlation between these film

577 thickness and rheological properties of AAMs has not been quantitatively proposed. Furthermore,

578 rheological parameters were also affected when precursors with different reactivities was mixed with

579 various proportions, due to changes of polymeric structure and polymer-particle interactions.

580

581 Table 2

582 Effects of precursor nature on the rheological properties of sodium silicate-activated blended systems.
Precursors L/S ratio (by Controlling Rheological parameters Decisive factors Refs.

volume factors 𝜏0 (Pa) K or 𝜇 (Pa*sn/

/mass ) Pa*s)

Metakaolin + L/S MK/FA ratio Higher specific [47]

fly ash (Grout) 0.75 0:10; 2:8; Decreased Increased from surface area of MK;

4:6; 6:4; 8:2 from 0.40 to 0.01 to 0.10 dense packing of FA

0.02

Fly ash + silica

f
w/b SF/FA ratio Higher specific [75]

oo
fume
0.75 0:10; 1:9; Increased Increased from surface area and
(Grout)

r
2:8; 3:7; 4:6 from 1.81 to 0.0105 to 0.066 reactivity of SF
-p
9.51
re
lP

Slag + fly ash + L/S FA/FAM ratio The “ball-bearing” [40]

fly ash 0.5 6:0; 5:1; 4:2; Increased Decreased effect, dense packing
na

microspheres 3:3; 2:4 from 2.51 to from 1.05 to and higher specific
ur

(Paste with 4.47, then 0.11 surface area of FAM

constant slag of decreased to


Jo

40 wt.%) 2.35

Fly ash + slag w/b FA/slag ratio Faster dissolution of [128]

(Paste) 0.45 2:1; 3:1; 4:1; Decreased - slag in low content;

10:0 from 18.6 to the formation of early

7, then C-A-S-H gels in high

increased to content of slag

9.8
Class C fly ash w/b FA/waste brick powder ratio Waste brick powder [43]

+ waste brick contains higher CaO


0.3 4:6; 3:7; 2:8; Decreased Decreased
powder content and shows
1:9; 0:10 from 693 to from 675 to 88
poor solubility
130

Slag + glass w/b Glass powder/slag ratio Lower specific [126]

powder 0.3 1:8; 3:6; 4:5; Decreased Decreased surface area and

(All Pastes 5:4; 6:3; 7:2 from 4.42 to from 7.2 to reactivity of glass

f
containing 10 1.8 2.95 powder

oo
wt.% calcium

r
aluminate

cement)
-p
re
Metakaolin + L/S The brick powder/MK ratio Higher specific [74]
lP

waste red brick 0.71 0:10; no Decreased surface area and

powder 2.5:7.5; 5:5; significant from 2.18 to reactivity of MK


na

(Paste) 7.5:2.5; 10:0 effect 0.09


ur
Jo

Metakaolin + L/B MK/slag ratio Higher specific [62]

slag 0.45 3:7; 5:5; 7:3 Increased Increased from surface area of MK

(Mortar with from 12 to 20 to 76

binder:sand 120

ratio of 0.33)
Metakaolin + L/B MK/slag ratio

slag 1 3:7; 5:5; 7:3 Increased Increased from Higher specific [140]

(Mortar with from 13.3 to 20.5 to 77 surface area of MK

binder:sand 118

ratio of 0.25)

Fly ash + slag L/B FA/slag ratio Higher specific [72]

(Mortar with 0.55 5:5; 8:2 Increased Increased from surface area of FA

binder:sand from 13.5 to 2.5 to 4.5

f
ratio of 0.5 ) 29.5

oo
583 * L/S = liquid/solid ratio; w/b = water/binder ratio; L/B = liquid/binder ratio.

r
584
-p
re
585 4.2 Content of precursors
lP

586 According to the Krieger-Dougherty formula [76] and the“YODEL” model [77], the apparent

587 viscosity (𝜂) and the yield stress (𝜏0 ) of suspension can be evaluated in accordance with Eq. (4) and
na

588 Eq. (5), respectively:


ur

𝜙 −[𝜂]𝜙𝑀
𝜂 = 𝜂0 (1 − ) (4)
Jo

𝜙𝑀
𝜙(𝜙 − 𝜙0 )2
𝜏0 = 𝑚1 (5)
𝜙max (𝜙max − 𝜙)
589 where 𝜂0 is the viscosity of interstitial fluid; 𝜙 is the actual solids packing fraction (by volume);

590 𝜙𝑀 is the maximum packing fraction (by volume); and [𝜂] is the intrinsic viscosity of the

591 suspension, which is related to the effect of individual particles on the viscosity. In Eq. (5), 𝜙0 is the

592 lower limit solid volume fraction (percolation threshold) at which the disperse phase behaves as a

593 three dimensional network structure; 𝑚1 represents the effects independent of solid volume fraction,

594 depending on the particle size distribution and interparticle force parameter. Based on the theory

595 from Eq. (4) and Eq. (5), the viscosity and yield stress of suspensions increase as the solid content

596 increases. It can be explained by the following two aspects: 1) more solid particles will induce the

597 particles-cluster effect; 2) less water added would increase the solids concentration, the inter-particle
598 friction and thereby the mortar is more reluctant to flow [23]. In other words, a high water to binder

599 ratio (w/b) leads to good dispersion of grains and low interparticle friction in the suspension [78],

600 and consequently reducing in shear stress and viscosity. As shown in Fig.15, the yield stress and

601 plastic viscosity of alkali-activated fly ash-silica fume blended grouts increased as w/b decreased [75,

602 79]. Similarly, the yield stress of MK suspensions activated by sodium silicate increased significantly

603 with the increase of precursor content (57 wt.% ~ 62 wt.%) under different curing time [45]. Even

604 more interesting is that the flow index (n) of MK-suspensions was reported to increase with the

605 solid/liquid content, regardless of curing time [45]. This was due to the change of particles

606 distribution (Brownian movement) and inter-particle structure when different shear rates were

f
oo
607 applied.

r
608
-p
re
lP
na
ur
Jo

609
610 Fig. 15. The yield stress (a) and plastic viscosity (b) of alkali-activated fly ash-silica fume blended grouts prepared

611 by various water to binder ratios [75].

612 As expected, Alonso et al. [31] observed that the static yield stress value declined by 96.65% when

613 the liquid to binder (L/B) ratio of sodium silicate-activated slag mortars (the aggregate/binder=2:1)

614 increased from 0.45 to 0.65. This is in line with a report on one-part sodium silicate activated

615 slag/glass powder pastes [126]. Increasing w/b from 0.275 to 0.350, the yield stress and plastic

616 viscosity of pastes were reduced by 64.14% and 91.48%, respectively [126]. Fig. 16 is a typical shear

617 rate-shear stress curve of pastes with different w/b ratios. In the low shear rate region, the measured

618 shear stress of sodium silicate-fly ash/slag (2:1) pastes decreased with the increase of w/b ratios, for
619 suggesting greater fluidity. This is due to strong hydrodynamic forces that seriously disturb the

620 attraction between particles, thus reducing particles agglomeration [128].

621

f
r oo
622
-p
re
623 Fig. 16. Shear rate-shear stress curve of fly ash/slag (2:1) pastes with different w/b ratios [128].

624
lP

625 5. Effects of chemical admixtures on rheology of AAMs


na
ur

626 5.1 Water-reducing admixtures


Jo

627 As is well known, water-reducing admixtures are widely used to improve the rheology and

628 workability of PCMs. Commonly used water-reducing admixtures include 1) lignosulfonate-based

629 type, 2) naphthalene-based, aminosulfonate-based and melamine-based water reducing admixtures,

630 and 3) polycarboxylate-based superplasticizers. The working mechanisms of superplasticizers and

631 their compatibility with Portland cement has been well explained in massive literature, where

632 inter-particle electrostatic repulsion and steric hindrance are induced by the effective adsorption of

633 admixtures on cement particles surface (Fig. 17), and consequently avoiding the formation of flocs

634 [80]. Thus, to some extent, such admixtures can reduce the water demand and improve durability and

635 mechanical properties [81, 82]. Encouraged by their early success of the application in PCMs,

636 researchers attempted to use these chemical admixtures to modify the rheology of AAM systems.

637 Table 3 summarizes and compares effects of various water-reducing admixtures with different
638 dosages on the rheological properties of AAMs in literatures.

639

640
641

f
Fig. 17. The schematic diagram of function mechanisms of superplasticisers on cement particles [83].

oo
642

r
643 Table 3

644
-p
The comparison of effects of various water-reducing admixtures with different dosages on the rheological
re
645 properties of AAMs.
lP

Dosage (by mass


Plasticizers Precursors Activators Main performance Ref.
of binder)
Na-silicates
na

Lignosulphonate (LS);
The minimum μ and
Melamine-derived
𝜏0 from mixtures with M
ur

synthetic polymers
0.80% Fly ash (paste) and M-PCE, respectively; [83]
(M); Modified
segregation in mixtures
Jo

polycarboxylic ethers
with LS
(M-PCE)

PC slightly reduced the


Polycarboxilate (PC);
viscosity; NS not only
naphthalene sulfonate 1.0%~5.0% Fly ash (paste) [130]
increased the viscosity,
(NS)
but also induced flash set
PC showed better
Class F-fly
plasticizing effect for
Naphthalene (N); ash; Class
~2.0% Class-C fly ash than for [119]
PC C-fly ash
Class F fly ash; effects of
(paste)
N can be negligible
PC was the most
effective admixture for
N; melamine (M) and Fly ash + slag
1.0% w/b ≥0.36 while N [127]
PC (paste)
showed better for w/b ≤
0.36 in one-part pastes
Fly ash + slag PC improved workability
N and PC 1.0%~4.0% [141]
(paste) more effectively than N

The minimum μ and 𝜏0


Polycarboxylic ethers Fly ash from mixtures with LS
1%; 1.5% [84]
(PCE); LS (concrete) (undergo segregation with
1.5%)
Sulfonated
melamine-formaldehy
de (SMF); copolymer
of carboxylic acrylic The minimum μ and 𝜏0
0.2%~3.0% Slag (mortar) [86]
acid with ester (CAE); from mixtures with LS
sulfonated

f
naphthalene-formaldeh

oo
yde (SNF); LS; PCE;
PCE; melamine Effects of all

r
formaldehyde Slag (paste and
derivative (M); 0.3%; 0.5%;
-p
mortar)
plasticizer-admixtures are
negligible
re
naphthalene 1.0%; 1.5%; [19]
NaOH
formaldehyde 2.0% Slag (paste and Mixtures with NF
lP

derivative (NF); vinyl mortar) reduced 𝜏0


copolymer (V)
na

Mixtures with 1.26 mg


M; NF; V ~0.4% NF significantly reduced [66]
𝜏0
ur

HPEG PCEs with high


HPEG PCEs with
Jo

anionicity, high molecular


various side chain
0.05% weight and short side [121]
length and anionic
chain length showed a
charge density
Slag (paste) better dispersion
Allylether maleic
anhydride-based PCE
possessing short side
PCE with different
NaOH/ chain showed the best
molecular structure 0.05% [120]
Na2CO3 dispersion in NaOH-slag;
and composition
no PCE showed
effectiveness in
Na2CO3-slag
One
Slag + fly ash
N; M and PC 1.0% part-Ca(OH PC was the most effective [132]
(paste)
)2+Na2SO4
646 5.1.1 Water-reducing admixtures in AAS systems

647 The suitability of commercial water-reducing admixtures to control the rheological properties of

648 AAS have been studied by many researchers. In NaOH-AAS system, proper dosage plasticizers

649 exhibited certain plasticizing effect in the early stage, but they might become unstable afterwards,

650 which was depending on the molecular structure at high pH conditions. Palacios et al. [19]

651 investigated the influences of four different high-range water-reducing admixtures (HRWRAs) on the

652 rheological behavior of NaOH-activated slag pastes/mortars. Only the naphthalene-based HRWRA

653 could lower the yield stress prominently due to its chemical stability in the alkaline environment.

f
oo
654 Though the lower molecular weight fragments of the polycarboxylate-based, melamine-based and

655 vinyl copolymer HRWRAs may be still adsorbed on the particles surface and induce an electrostatic

r
656 -p
repulsion, the steric contribution is not noticeable [19, 87]. It is worth noting that the presence of
re
657 melamine-based superplasticizer caused much higher yield stress in the suspension activated by

658 NaOH and the reason for this phenomenon was not clear [87].
lP

659 The authors also reported similar behavior of these admixtures in their subsequent experiments on
na

660 the 13.6-pH NaOH-AAS pastes [66]. As shown in Fig. 18, the yield stress and plastic viscosity of

661 AAS pastes with different amounts of the melamine-based admixture were all higher than those of
ur

662 the blank control group, when the shearing regime was implemented for at least 10 minutes. The
Jo

663 plasticizing effect of vinyl copolymer seems less prominent in AAS pastes. Only the

664 naphthalene-based superplasticizer with 1.26 mg substantially reduced both yield stress and plastic

665 viscosity in NaOH-AAS pastes (in Fig. 18c-d). This is due to the fact that the adsorption-mediated

666 decrease in size and number of agglomerated particles, which further supports the chemical stability

667 of this admixture. These results revealed the suitability and plasticizing effectiveness of different

668 superplasticizers in alkaline environment of NaOH-AAS pastes.

669
670

f
r oo
-p
re
lP
na
ur

671
Jo

672
673 Fig. 18. The yield stress and plastic viscosity evolution in 13.6-pH NaOH-AAS pastes containing three types of

674 superplasticizers: melamine-based, naphthalene-based and vinyl copolymer at different dosages. Data from
675 Palacios et al. [66].

676

677 In the AAS suspensions activated by sodium silicates, Palacios et al. [19] observed that all

678 water-reducing admixtures were useless. Actually the yield stress even increased in the case of using

679 vinyl copolymer HRWRA. However, Luukkonen et al. [86] pointed out that the first- and

680 second-generation plasticizers were more efficient than the commonly used third-generation

681 superplasticizers, particularly when adding 0.5 wt.% lignosulfonate-based superplasticizer could

682 decrease the yield stress and viscosity of AAS mortars to 56% and 27% at 40 min, respectively.

683 Palacios and Puertas [87, 88] reported functional groups changed (as indicated by the shifts in the

f
oo
684 infrared spectra) when the naphthalene-based HRWRA was mixed in sodium silicates activated slag

r
685 systems. These changes (modified or released) in the sulphonates might be account for the failure of

686
-p
naphthalene-derived admixtures. Nevertheless, they did not provide a detailed description of exact
re
687 structural alterations taking place for this admixture. Interestingly, Ren et al. [89] suggested that the
lP

688 addition of lignosulfonates and sodium silicates separately could reduce the likelihood of failure.

689 Compared to simultaneous addition method, adding the lignosulfonates 3 minutes prior to or after the
na

690 sodium silicates both led to a great reduction on both yield stress and plastic viscosity, thus obtained
ur

691 the increased mini-slump spreads. They also observed shear thickening behaviour when the

692 lignosulfonates and the activator were added separately in the sodium silicate activated slag.
Jo

693 The effective adsorption of admixtures on precursor particles is very important for their

694 plasticizing efficiency. Nevertheless, the results from Palacios et al. [66] showed that the adsorption

695 of the three superplasticizers (melamine-based, naphthalene-based and vinyl copolymer) on slag

696 particles in AAS pastes was much lower than on PC particles, and variation of pH of the alkaline

697 solution seemed less important for the adsorption behavior. This may be partially due to the

698 difference of adsorption mechanisms in the two suspensions, since the initial Zeta potential of the

699 11.7-pH NaOH AAS paste was negative (about −2 mV) owing to deprotonation whereas that of PC

700 paste was positive (about +0.5 mV). Generally, with the increasing adsorption of anionic

701 superplasticizer moleculars (admixtures containing sulfonic groups) on the particles surface, the Zeta

702 potential values of blast furnace slag and fly ash suspensions gradually decrease (more negative),
703 which indicated that electrostatic repulsion may play a role in the dispersion mechanism [53]. The

704 working mechanism of these superplasticizers in NaOH-AAS were further studied by Palacios et al.

705 [115] later and they turned out that that steric hindrance was dominant in the working mechanisms of

706 both the sulfonate and polycarboxylate superplasticizers in NaOH-AAS. Melamine admixtures

707 induced the highest contribution (only around 26%) to the electrostatic repulsion, and the

708 naphthalene derivative admixture with the lowest molecular weight presented the smallest range of

709 steric repulsion. Compared to the melamine derivative and vinyl copolymer, polycarboxylate

710 superplasticizer with lower molecular weight showed the highest steric hindrance, which might be

711 attributed to different adsorption conformations related to molecular structures [115-116]. In addition,

f
oo
712 Kashani et al. [30] noticed that the addition of polycarboxylates in sodium silicates-AAS pastes

r
713 reduced the flow index in the H–B model with shear thinning behavior. A dramatic yield stress

714
-p
reduction was found in the mixture with higher density of long side chains in the lower molecular
re
715 weight polymers. Side chains were assumed to work as a steric effect supplier to disperse particles
lP

716 after polycarboxylate ethers molecular was adsorbed on particles surface. Because of this, a bridging

717 effect may occur to provide partial protection of the backbone charges against attachment of one
na

718 admixture molecule to two or more slag particles, thereby enhancing the plasticizing efficiency.
ur

719 Interestingly, the plasticizing mechanisms are similar despite the different charge signs of molecules,

720 which indicates that both anionic and cationic adsorption sites are available on slag particle surface
Jo

721 in the alkaline environment [30].

722 In addition to the chemical stability, the solubility in alkaline media [120-121] and the competitive

723 adsorption between the anionic species of activators and anionic superplasticizers [121-122] also

724 affect the adsorption efficiency and dispersion effect of these superplasticizers in AAMs. Conte et al.

725 [120] reported that the solubility difference of PCEs with different molecular structures in NaOH and

726 Na2CO3 solutions induced different dispersing efficiency. All the PCEs tested in the study had no

727 important impact on improving the flow of Na2CO3-slag paste, due to the extremely low solubility in

728 Na2CO3 solution. Moreover, allyl ether based PCEs with a high molecular weight (Mw > 30000 Da)

729 and a short side chain (nEO = 7) exhibited the best dispersing performance in NaOH-slag system,

730 which might be attributed to their good solubility in NaOH solution and stronger chelating ability
731 with Ca2+ ions. It should be noted that better solubility of these polymers in alkaline solutions does

732 not always mean superior dispersing ability. Lei et al. [121] synthesized a series of PCEs with

733 different anionic charge density or side chain lengths, based on α-methallyl-ω-hydroxy poly

734 (ethylene glycol) ether (HPEG). As shown in Fig. 19, even 23HPEG15 (Mw = 42000 Da; nEO = 23)

735 showed higher solubility (4.50 mol/L) in NaOH solutions than 23HPEG7(a) (Mw = 410000 Da; nEO

736 = 23; 2.96 mol/L NaOH), it presented poor dispersing performance in NaOH-slag pastes. Therefore,

737 the solubility of PCEs is a limit factor of dispersing efficiency of such polymers in AAMs, but not

738 the decisive factor.

739

f
r oo
-p
re
lP
na
ur

740
Jo

741 Fig. 19. NaOH concentrations (a) needed to achieve the cloud point of different HPEG PCEs, and

742 spread flow (b) of NaOH-slag pastes containing 0.05 wt% of different HPEG PCEs [121].

743

744 It is well known that many commercial anionic superplasticizers improve adsorption efficiency by

745 the strong chelation with Ca2+ ions. Due to the deprotonation of hydroxyl, the negatively charged

746 slag particles attract some Ca2+ ions in alkali-activated slag system, which provides a possibility for

747 the adsorption of anionic superplasticizers on slag particles surface. However, current research shows

748 that the adsorption capacity of superplasticizers on slag surface is still far less than that in PCMs. To

749 some extent, the low adsorption efficiency might be generically grouped under competitive

750 adsorption between such polymer molecules and anionic species in activators [120-122]. Marchon et

751 al. [122] considered that the fast flow loss of PCEs in alkali-activated system was probably due to a
752 competitive adsorption between PCE molecules with the hydroxyls. As previously discussed, the low

753 adsorption performance of PCEs in Na2CO3-slag pastes could also be accounted for the almost

754 complete removal of Ca2+ ions in solutions resulting from precipitation by the CO32- ions [120]. It

755 could lead to an insufficient adsorption of Ca2+ ions on slag particles, lowing adsorption efficiency of

756 PCEs. Similarly, Palacios et al. [88] found that all the superplasticizers tested in sodium silicate-slag

757 pastes failed to provide any dispersing effect. To a certain extent, this could be attributed to the

758 preferential adsorption of silicate species with Ca2+ ions dissolved from slag.

759 5.1.2 Water-reducing admixtures in AAFA systems

f
oo
760 The effects on rheology of the three generations of water-reducing agents also have also been

r
761 studied in AAFA pastes and concretes. Compared to the blank control pastes, the AAFA containing

762
-p
the tested water-reducing admixtures (including lignosulphonates, melamines and polycarboxylates)
re
763 all showed reduced yield stress and plastic viscosity after mixing and the effectiveness was only in
lP

764 the early age [83, 84]. As discussed earlier, the chemical stability and solubility of superplasticizers

765 in alkali solution, and the adsorption degree on particles should be considered as possible reasons for
na

766 low plasticizing efficiency. Interestingly, Laskar and Bhattacharjee [84] pointed out that the
ur

767 concentration of 4 M for NaOH solution was a critical value, below which increasing dosages of

768 both lignosulphonate-based and polycarboxylate-based admixtures could improve workability of fly
Jo

769 ash paste activated by sodium silicate. However, it still unclear why the addition of

770 lignosulphonate-based admixture is more likely to induce segregation in fly ash based-suspensions

771 [83, 84]. Furthermore, unlike in AAS suspensions, Criado et al. [83] reported that polycarboxylate

772 superplasticizers in AAFA systems showed the most effective plasticizing properties among these

773 three types of water-reducing agents. Most admixtures including the third-generation

774 supersuperplasticizer are designed to form complexes with the dissolved Ca2+ in the early hydration

775 of Portland cement [85]. Slag and high-calcium (Class C) fly ash can also release large amounts of

776 Ca2+, yet hardly any calcium is dissolved in low-calcium fly ash (Class F) systems. Therefore, the

777 steric-hindrance effect is prohibited or not trigged in Class-F fly ash systems. As reported by Montes

778 et al., the polycarboxylate-based admixture only slightly reduced the viscosity of fly ash paste
779 activated by Na-silicate [130]. In another study, however, polycarboxylates were more effective to

780 induce apparent drop of shear stress and viscosity in sodium silicate-Class C fly ash, and this higher

781 adsorption efficiency was attributed to the presence of positively charged Ca2+ [119]. In contrast to

782 PC, the drop in the plastic viscosity and yield stress of AAFA pastes containing plasticizers is not

783 always accompanied by the improvement of fluidization performance. There was hardly correlation

784 between slump and the rheological parameters in the research by Criado et al. [83]. This is adverse to

785 the report from Laskar and Bhattacharjee [84] who concluded that good correlation between

786 rheological parameters and slump test results in AAFA concretes containing different plasticizers. It

787 is evident that the properties of fly ash, the interaction between water-reducing admixtures and fly

f
oo
788 ash, and the type of activator and the addition method should be considered in the future research for

r
789 different AAM systems.
-p
790 5.2 Other chemical admixtures
re
lP

791 In addition to water-reducing admixtures, other commonly used admixtures, such as shrinkage

792 reducing agents, retarders, deflocculants, etc., have also been studied for their impacts on rheological
na

793 behavior of AAMs. The rheological modification effect of a polypropyleneglycol derivative


ur

794 shrinkage reducing agent on AAS was investigated by Palacios et al. [19]. Although these authors

795 have proved that this admixture was stable in highly alkaline solutions [88], it did not improve the
Jo

796 fluidity of NaOH-activated slag pastes, and even increased the yield stress of sodium silicates-slag

797 pastes. A recent publication showed that fluidity of AAS pastes with polypropylene glycol

798 (PPG)-based shrinkage reducing agent could be increased (PPG1000), reduced (PPG400), or

799 maintained (PPG2000) at a similar level (Fig. 20), which depended on the dosages and molecular

800 weight of PPG [118]. The difference in fluidity can be attributed to the effect of PPG on the viscosity

801 of suspensions and the dissolution process of slag grains [118]. Hence the chemical stability and

802 rheological properties of shrinkage reducing agents with different structure and molecular weight in

803 high alkaline-AAMs need more research. Xu et al. [90] found that citric acid can work not only as a

804 setting retarder, but also a potential plasticizer in AAMs. As shown in Table 4, the NaOH activated

805 fly ash (Class C)-lime kiln dust paste (marked as AAFA-LKD) with addition of 3 wt.% citric acid
806 showed a low plastic viscosity and a comparable yield stress to Portland cement (PC), and these

807 rheological parameters could be maintained at viable levels in the first 1h. The introduction of citric

808 acid reduces the alkalinity of suspension and thus delay the dissolution process. In addition, the

809 formation of carboxylate complexes between the carboxylic acid groups of citric acid and Ca2+ ions

810 released from calcium-containing precursors would also affect the gelation process, then reducing

811 yield stress. Similarly, Wang et al. [61] also reported that the introduction of a retarder was effective

812 in improving the flowability of the alkali-activated carbonatite-slag grouting materials.

f
r oo
-p
re
lP

813
na

814 Fig. 20. Fluidity of the AAS pastes after adding different dosages of polypropylene glycol (PPG)-based shrinkage
ur

815 reducing agent with different molecular weights [118].

816
Jo

Table 4

817 Plastic viscosity and yield stress of Portland cement (PC) paste and alkali-activated fly ash (Class-C)-lime kiln dust

818 (marked as AAFA-LKD) pastes without and with 3 wt% citric acid at different times after reaction [90].

Pastes PC AAFA-LKD AAFA-LKD + 3% citric acid


Time (min) 10 5 10 5 10 60
Plastic viscosity (Pa·s) 12.4 12.6 13.6 4.4 3.9 6.5
Yield Stress (Pa) 105 69.3 432 91.8 72.5 64.7

819 * In the table of Ref [90], the unit of yield stress is (MPa). This might be wrong, so here it is

820 corrected here as (Pa).

821 Sodium polyacrilate with low molecular weight (Mw=2000 au) was used to work as a deflocculant

822 to reduce the yield stress and shear viscosity of aqueous suspensions of kaolinite [45,117]. However,

823 the sodium polyacrylate has limited effect on reducing yield stress and viscosity of metakaolin-based
824 AAMs, or even for the aqueous suspensions of metakaolinite [45]. Therefore, the admixture in this

825 case behaves more like a flocculant than deflocculant. Grout materials usually require a good balance

826 between flowability and stability. Viscosity modifying agents or stabilizers are usually added to

827 prevent segregation and bleeding. Aboulayt et al. [47] applied polysaccharide-based stabilizer

828 (xanthan gum) in metakaolin-fly ash blended AAM grouts and found the beneficial effect on the

829 stability. For a given fly ash ratio, both the yield stress and plastic viscosity at different shear rates

830 increased with the increase of stabilizer dosage. Similarly, excellent pumpability at high temperature

831 is vital for drilling fluid in oil well cementing engineering. A novel self-degradable temporary sealing

832 material based on sodium carboxymethyl cellulose-modified alkali-slag-fly ash (Class C) material

f
oo
833 was developed, and the addition of 2 wt.% polylactic acid could enhance the pumpability and reduce

r
834 the viscosity [91].
-p
835 6. Effects of mineral additions on rheology of AAMs
re
lP

836 Whether inert or latent hydraulic mineral additions with small particle size, they all have physical

837 filling effect in the hydration process, not only in PCMs but also AAMs. The high surface area of the
na

838 fines can provide abundant nucleation sites for the hydration. Accordingly, both the fresh and
ur

839 hardened properties will be affected to varying degrees.


Jo

840 6.1 Reactive mineral additions

841 Silica fume (SF) is a frequently-used pozzolanic addition, and it also can be used as a precursor in

842 AAMs to mix with slag or fly ash. Here, in order to better explain the influence of SF, it will be

843 discussed in this part. The rheological modification of AAMs by SF is closely linked with the water

844 to binder ratio (w/b) and its physicochemical properties. The amount of SF that caused significant

845 increase in the yield stress and plastic viscosity of AAFA grout decreased as w/b ratio (Fig.15). For

846 example, the critical content of SF in the suspension with w/b of 0.75 is 20%, while 50% is the

847 critical content of SF in the suspension with w/b of 1.25 (Fig.15). Beyond this critical content, it was

848 observed that the yield stress and plastic viscosity of AAFA grout increased with the increasing of SF

849 proportion [75]. This is similar to the impact of adding SF in the conventional grouts [92, 93]. The
850 influence of higher specific surface (21080 m2/kg for SF, in comparison with 379 m2/kg for FA) [75]

851 and reactivity of SF under the interaction of alkaline activators must be taken into account [92, 94,

852 126]. The use of SF above the critical content means that there is insufficient water to wet and react,

853 leading to a pronounced increase in the yield stress and plastic viscosity. On the other hand, the “ball

854 bearing” effect of this spherical ultra-fine powder (the passing rate of 0.01mm is 93% for SF, while

855 18% passing for FA) could play a role as lubricators [40, 75]. Below this critical content, the

856 lubricating effect of SF was dominant due to spherical particles. In a recent study, ultra-high

857 performance concrete based on alkali-activated slag with a very low w/b ratio of 0.175 retained good

858 workability by incorporating with 12.5 wt.% SF [97]. Similarly, Liu et al. [98] reported that the

f
oo
859 workability of alkali-activated ultra-high strength concrete increased when used less than 10 wt.% SF,

r
860 since the lubricating effect of SF was more important in this case.

861
-p
It is well understood that a high thixotropy characterized by shear thinning and large viscosity
re
862 recovery is more favorable for printable materials. Fig. 21 illustrates the viscosity recovery of AAS
lP

863 system with different amounts of thermal treated nanoclay (NC) [99]. It shows that NC content has

864 little effect on the initial apparent viscosity, but more NC addition yields larger viscosity recovery as
na

865 a result of a more rigid AAS through the flocculation effect of clay. As shown in Fig. 22, the initial
ur

866 static yield stress of mixtures with 0.2% NC was around triple that of NC free mixtures. This

867 indicates that the flocculation effect of NC is far greater than that of the plasticizing and
Jo

868 deflocculating effects induced by silicate particles, so NC modified 3D printing-AAS shows

869 significant static yield stress. It is important to note that structural rebuilding rate (the rate of change

870 of static yield stress with time) is not sensitive to the presence of NC. Steel slag (SS) is a

871 metallurgical slag containing a large number of low reactive or inert ingredients and its pozzolanic

872 reactivity is low [100]. Increasing the replacement ratio of SS in AAS-based 3D printing resulted in

873 the reduction of thixotropy [48]. Raising SS content reduces the degree of hydration, and therefore

874 the connection between particles is weakened due to fewer intial gels formed after mixing, thus

875 reducing the reversible connection of particles [48]. Besides, the addition of SS reduced the plastic

876 viscosity of AAM due to the smaller specific surface area of SS [101].

877
878

f
oo
879 Fig. 21. Effect of the nanoclay (NC) on the viscosity recovery of 3D printing-AAS mixtures [99].

r
-p
re
lP
na
ur

880
Jo

881 Fig. 22. Effect of the nanoclay (NC) dosage on the apparent viscosity and static yield stress of 3D printing-AAS

882 mixtures [99].

883

884 6.2 Inert mineral additions

885 The addition of a proper amount of inert mineral additions, such as limestone (LS) powder, can

886 improve rheological properties by working as fillers to optimize the particle packing model of blend

887 system, and more water for lubrication. This can compensate for the negative effects of high specific

888 surface area and nucleation effect. According to the study by Xiang et al. [38], the incorporation of 5

889 wt.%-20 wt.% LS powder in sodium silicate-activated slag-fly ash blended grout did not alter the

890 rheology model (Bingham model), but showed a positive effect on reducing yield stress and plastic
891 viscosity (Fig. 23), especially for the grouts activated by solid sodium silicate. As shown in Fig. 24,

892 the particle size range and d50 of LS are lower than those of slag and fly ash. To a certain extent, LS

893 can therefore play a filling role and reduce the wetting water demand. Similarly, to modify the very

894 sticky AAFA binders for 3D printing purpose, Alghamdi et al. [102] used 15 wt.%-30 wt.%

895 limestone powder (d50 =1.5 μm) to replace fly ash (d50 =17.9 μm). As expected, a partial

896 replacement by limestone powder can effectively address this issue and improve print quality, due to

897 the consistency and water retention. Moreover, it was [29] also found that a similar positive effect

898 could be achieved in sodium silicate-activated MK grout by replacing MK with different percentages

899 of Fuller-fine sand powders (by10, 20, 30 and 40 wt.%). For example, in the sample group

f
oo
900 synthesized with 70 wt.% MK (specific surface area=16646 m2/kg) and 30 wt.% Fuller-fine sand

r
901 (specific surface area=8886 m2/kg), the yield stress was reduced by 73% at the first shear cycle.

902
-p
However, it should be noted that adding different volume fraction (10%, 30% and 60 wt.%) of fine
re
903 silica sand as filler to Na-silicate based MK pastes caused an increase in viscosity, but less than if the
lP

904 water content of pastes was reduced to obtain the same solid volume fraction [134].

905
na
ur
Jo

906
f
oo
907

r
908 Fig. 23. The 27 min yield stress (a) and plastic viscosity (b) of the alkali-activated slag-fly ash (AASF) grouts

909
-p
using liquid waterglass and solid waterglass as activator, after adding 0-20 wt.% limestone (LS) [38].
re
lP
na
ur
Jo

910
911 Fig. 24. Particle size distribution of slag, fly ash and limestone powder [38].

912

913 7. Effect of aggregates on rheology of AAMs

914 Currently, most research efforts are being focused on the rheology AAM paste while only very few

915 studies regarding the rheology of AAM mortar/concretes have been reported. For AAMs paste,

916 Bingham model is normally recommended to describe the rheological behavior of NaOH

917 activated-pastes, and H-B model is more suitable for sodium silicate-activated pastes [19, 41].
918 However, for alkali-activated mortar/concrete, Bingham model suits both two systems and this

919 behavior is similar to PC based mortar/concrete [20, 31]. Furthermore, compared with PC mortars, it

920 was reported that the workability of both alkali-activated fly ash and alkali-activated slag mortars

921 was more sensitive to changes of liquid/solid ratio [31]. You et al. [72] found that using copper slag

922 (CS) as fine aggregate to replace natural sand (NS) could alter rheological characteristic of the

923 alkali-activated fly ash-slag mortar. It was observed that a low replacement by 20% and 40% (by

924 volume) CS resulted in higher yield stress and consistency, due to the characterics of CS by more

925 irregular shape and containing higher amounts of fine powders [103]. Higher replacement (60%) led

926 to undesiable rheological responses, such as undergoing segregation and bleeding, which might

f
oo
927 thanks to a relatively higher density of CS than NS [104].

r
928 8. Effect of fibers on rheology of AAMs -p
re
929 Though AAMs have been shown to possess more compact microstructure than PCMs in numerous
lP

930 literature, and good mechanical properties, the large shrinkage property, particularly for sodium

931 silicate activated systems, remains a big challenge, particularly for potential application in aggressive
na

932 environments. This is because excessive shrinkage can cause microcracks. Therefore, in order to
ur

933 address this problem, adding fibers to AAMs becomes an approach and this leads to recent

934 innovation of high-performance and ultra-high-performance AAM concrete (UHPC) have been
Jo

935 conducted in recent years. As expected, the flowability of AAM-based UHPC decreased as steel fiber

936 volume fractions increased from 1%, 2% and 3% [98, 105]. The mixtures with hooked-end fiber

937 showed the best fluidity at high fiber content while the effect of fiber shapes was negligible at low

938 fiber content (Fig. 25). Moreover, Liu et al. [105] investigated effects of various aspect ratios

939 (length/diameter: 6/0.12, 8/0.12,13/0.12, 13/0.2) of straight steel fiber on the flowability. They

940 reported that the aspect ratio of 6/0.12 is the most optimal choice to ensure good flowability when

941 the steel fiber volume fraction increased to 3% (Fig. 26). The increase of the polyvinyl alcohol (PVA)

942 fiber volume also can give rise to high viscosity and yield stress of the AAS composite due to its

943 poor dispersibility. Even so, Choi et al. [106] have successfully produced AAS based-composite with

944 high ductility, low plastic viscosity and yield stress, at the proportion of w/b of 0.4, superplasticizer
945 of 0.27% and PVA fibers of 1.3 vol.%.

946

f
oo
947

r
948 -p
Fig. 25. Flowability of fresh AAM-based UHPC with different steel fibers [105]. Note: “D&L” means “diameter
re
949 and length”.
lP
na
ur
Jo

950
951 Fig. 26. Flowability of fresh AAM-based UHPC with different aspect ratios of straight steel fibers [105].

952

953 9. Conclusions and perspectives

954 9.1 Conclusions

955 This paper has reviewed the progresses in the the rheology of AAMs influenced by composition
956 factors including activators, precursors, admixtures and additions. Based on these discussions, the

957 following conclusions can be drawn:

958 1) Current models for PCMs can also describe the rheological behavior of AAMs though there

959 is suitability difference for various systems. The type of activator seems to be the

960 predominant factor that control the rheological model. Generally, Bingham model is

961 recommended for NaOH-activated pastes, while H-B model suits better the sodium

962 silicate-activated pastes. And other component factors such as precursors and chemical

963 admixtures also have important effects on rheological parameters, for example, Na-silicate

964 activated fly ash pastes behaved like a Bingham fluid in some reports, while the rheological

f
oo
965 behavior of a fly ash-slag composite was more consistent with H-B model.

r
966 2) The polymerization status of silicate species from activator solution and their adsorption

967
-p
mechanism on precursor surface plays an important role affecting rheological properties of
re
968 AAMs. Furthermore, due to the different ion sizes, Na-based activators usually induce
lP

969 higher yield stress, higher viscosity and faster viscoelastic evolution of suspensions

970 compared to K-based activators. Silicates-based activators at proper doses can have an
na

971 excellent plasticizing effect. In general, there is a critical concentration of activators


ur

972 (modulus and Na2O%), which can reverse the trend of rheological properties by affecting

973 the ion distribution and activation.


Jo

974 3) The physical properties of precursors affect their surface water absorption and packing

975 density, whereas the chemical properties change dissolution behavior of precursor particles

976 and types of precipitated gels, both altering rheology of AAMs. In alkali-activated blended

977 systems, the content of precursor with high specific surface area (such as MK) is usually the

978 predominant factor that increases the yield stress and plastic viscosity.

979 4) Nowadays most of the common chemical admixtures for PCMs show lower dispersion

980 efficiency in AAMs, which can be explained by three aspects: chemical stability in the

981 extremely alkaline environment, the solubility in alkaline media, and the competitive

982 adsorption between the anionic species of activators and anionic chemical admixtures. All

983 these aspects depend on the rational design of polymers molecular structure and charge. A
984 few candidates show effectiveness for some specific systems: HPEG PCEs with high

985 anionicity, high molecular weight and short side chain length showed a good dispersion in

986 high-calcium systems.

987 5) The addition of inert mineral additions, such as fine sand and limestone power, have

988 exhibited a positive effect on improving rheological properties due to better particle dense

989 packing model, which can compensate for the negative effects of high specific surface area

990 and nucleation effect. On the other hand, the impact of reactive mineral additions on the

991 rheology of AAMs is complex, depending on their physicochemical characteristics and the

992 water to binder ratio of suspension.

f
oo
993 6) Bingham model could well describe the rheology behavior of AAMs mortar/concrete.

r
994 Furthermore, the flowability of AAMs mortars is proportional to liquid/solid ratio and

995
-p
significantly affected by the type of aggregate. For fibre reinforced AAMs, like in PCMs,
re
996 the shape and aspect ratio of fiber are two vital factors.
lP

997 9.2 Perspectives


na

998 Although a great effort has been made in the field, there is still a large gap between recent progress
ur

999 and full understanding of rheology of AAMs and facilitation of engineering application. In future

1000 research, the following concepts (concerns) can be considered:


Jo

1001 1) More in-depth studies on the relationship between the composition factors and inter-particle

1002 forces, as well as hydration kinetics in the AAMs. This will provide a deeper understanding

1003 of the interaction mechanisms and a full picture of effects of various components.

1004 2) More efforts should be paid with regard to some research gaps and limited research,

1005 including the way of adding activator, NaOH-MK based systems or systems applying a

1006 greener activator (such as Na2CO3 or a mixture of them) and so on.

1007 3) Robust formulations and relevant chemicals to meet rheological and mechanical

1008 requirements of AAMs. For instance, the development of organic plasticizers tailored for

1009 AAMs or the appropriate addition of inorganic plasticizers to improve the poor workability

1010 of AAMs are required.


1011 4) The rheology studies on AAM concrete. There are some works on paste and mortar, but

1012 concrete study is limited, especially the effect of fibers and additives on rheology.

1013 Establishing empirical formulas to quantitatively describe the relationship between these

1014 materials and the rheological properties of AAMs are necessary, and thus promotes

1015 large-scale commercial applications of AAMs.

1016

1017 Acknowledgements

1018 The authors are grateful to the financial support by the National Natural Science Foundation of

f
oo
1019 China projects 51878263, U 2001225 and 51638008.

r
1020
-p
1021 References
re
lP

1022 [1] A.O. Purdon, 1940, The action of alkalis on blast-furnace slag, J. Soc. Chem. Indus. 59 (1940)

1023 191-202.
na

1024 [2] C. Shi, D. Roy, P. Krivenko, Alkali-activated cements and concretes, CRC press, 2003.
ur

1025 [3] P. Chindaprasirt, T. Chareerat, V. Sirivivatnanon, Workability and strength of coarse high calcium

1026 fly ash geopolymer, Cem. Concr. Compos. 29 (2007) 224-229.


Jo

1027 [4] T. Yang, X. Yao, Z. Zhang, H. Wang, Mechanical property and structure of alkali-activated fly ash

1028 and slag blends, J. Sustain. Cement-Based Mater. 1 (2012) 167-178.

1029 [5] J.L. Provis, P. Duxson, J.S.J. Van Deventer, G.C. Lukey, The role of mathematical modelling and

1030 gel chemistry in advancing geopolymer technology, Chem. Eng. Res. Des. 83 (2005) 853-860.

1031 [6] D.L. Kong, J.G. Sanjayan, K. Sagoe-Crentsil, Comparative performance of geopolymers made

1032 with metakaolin and fly ash after exposure to elevated temperatures. Cem. Concr. Res. 37 (2007)

1033 1583-1589.

1034 [7] C. Shi, Characteristics and cementitious properties of ladle slag fines from steel production, Cem.

1035 Concr. Res. 32 (2002) 459-462.

1036 [8] J.L. Provis, A. Palomo, C. Shi, Advances in understanding alkali-activated materials, Cem. Concr.
1037 Res. 78 (2015) 110-125.

1038 [9] Z. Zhang, Y. Zhu, T. Yang, L. Li, H. Zhu, H. Wang, Conversion of local industrial wastes into

1039 greener cement through geopolymer technology: A case study of high-magnesium nickel slag, J.

1040 Clean. Prod. 141 (2017) 463-471.

1041 [10] D.M. Roy, W. Jiang, M.R. Silsbee, Chloride diffusion in ordinary, blended, and alkali-activated

1042 cement pastes and its relation to other properties, Cem. Concr. Res. 30 (2000) 1879-1884.

1043 [11] E. Douglas, A. Bilodeau, V.M. Malhotra, Properties and durability of alkali-activated slag

1044 concrete, Mater. J. 89 (1992) 509-516.

1045 [12] T. Williamson, M.C.G. Juenger, The role of activating solution concentration on alkali–silica

f
oo
1046 reaction in alkali-activated fly ash concrete, Cem. Concr. Res. 83 (2016), 124-130.

r
1047 [13] P. Duxson, J.L. Provis, G.C. Lukey, J.S. Van Deventer, The role of inorganic polymer

1048
-p
technology in the development of ‘green concrete’, Cem. Concr. Res. 37 (2007) 1590-1597.
re
1049 [14] A.A.M. Neto, M.A. Cincotto, W. Repette, Drying and autogenous shrinkage of pastes and
lP

1050 mortars with activated slag cement, Cem. Concr. Res. 38 (2008) 565-574.

1051 [15] N.K. Lee, J.G. Jang, H.K. Lee, Shrinkage characteristics of alkali-activated fly ash/slag paste
na

1052 and mortar at early ages, Cem. Concr. Compos. 53 (2014) 239-248.
ur

1053 [16] P. Chindaprasirt, P. De Silva, K. Sagoe-Crentsil, S. Hanjitsuwan, Effect of SiO2 and Al2O3 on

1054 the setting and hardening of high calcium fly ash-based geopolymer systems, J. Mater. Sci. 47 (2012)
Jo

1055 4876-4883.

1056 [17] W. Rakngan, T. Williamson, R.D. Ferron, G. Sant, M.C. Juenger, Controlling workability in

1057 alkali-activated Class C fly ash, Constr. Build. Mater. 183 (2018) 226-233.

1058 [18] S. Choi, K. Lee, Influence of Na2O Content and Ms (SiO2/Na2O) of Alkaline Activator on

1059 Workability and Setting of Alkali-Activated Slag Paste, Mater. 12 (2019) 2072.

1060 [19] M. Palacios, P.F. Banfill, F. Puertas, Rheology and setting of alkali-activated slag pastes and

1061 mortars: effect of organic admixture, ACI Mater. J. 105 (2008) 140.

1062 [20] F. Puertas, B. González-Fonteboa, I. González-Taboada, M.M. Alonso, M. Torres-Carrasco, G.

1063 Rojo, F. Martínez-Abella, Alkali-activated slag concrete: Fresh and hardened behavior, Cem. Concr.

1064 Compos. 85 (2018) 22-31.


1065 [21] N. Roussel, Understanding the rheology of concrete, Woodhead Publishing, 2011.

1066 [22] D. Jiao, C. Shi, Q. Yuan, X. An, Y. Liu, H. Li, Effect of constituents on rheological properties of

1067 fresh concrete-A review, Cem. Concr. Compos. 83 (2017) 146-159.

1068 [23] A. Yahia, K.H. Khayat, Analytical models for estimating yield stress of high performance

1069 pseudoplastic grout, Cem. Conc. Res. 31 (2001) 731-738.

1070 [24] A. Yahia, K.H. Khayat, Analytical models for estimating yield stress of high-performance

1071 pseudoplastic grout, Cem. Concr. Res. 31 (2001) 731-738.

1072 [25] D. Feys, R. Verhoeven, G. De Schutter, Evaluation of time independent rheological models

1073 applicable to fresh self-compacting concrete, Appl. Rheol. 17 (2007) 56241-56244.

f
oo
1074 [26] W.H. Herschel, R. Bulkley, Konsistenzmessungen von gummi-benzollösungen,

r
1075 Kolloid-Zeitschrift 39 (1926) 291-300.

1076
-p
[27] Y. Liu, C. Shi, D. Jiao, X. An, Rheological properties, models and measurements for fresh
re
1077 cementitious materials – A short review, J. Chinese Ceram. Soc. 45 (2017) 708-716 (in Chinese with
lP

1078 English abstract).

1079 [28] K. Vance, A. Dakhane, G. Sant, N. Neithalath, Observations on the rheological response of
na

1080 alkali activated fly ash suspensions: the role of activator type and concentration, Rheol. Acta 53
ur

1081 (2014) 843-855.

1082 [29] J. Xiang, L. Liu, X. Cui, Y. He, G. Zheng, C. Shi, Effect of Fuller-fine sand on rheological,
Jo

1083 drying shrinkage, and microstructural properties of metakaolin-based geopolymer grouting materials,

1084 Cem. Concr. Compos. 104 (2019) 103381.

1085 [30] A. Kashani, J.L. Provis, J. Xu, A.R. Kilcullen, G.G. Qiao, J.S. van Deventer, Effect of molecular

1086 architecture of polycarboxylate ethers on plasticizing performance in alkali-activated slag paste, J.

1087 Mater. Sci. 49 (2014) 2761-2772.

1088 [31] M.M. Alonso, S. Gismera, M.T. Blanco, M. Lanzón, F. Puertas, Alkali-activated mortars:

1089 Workability and rheological behavior, Constr. Build. Mater. 145 (2017) 576-587.

1090 [32] A. Poulesquen, F. Frizon, D. Lambertin, Rheological behavior of alkali-activated metakaolin

1091 during geopolymerization, J. Non-Cryst. Solid. 357 (2011) 3565-3571.

1092 [33] A. Favier, J. Hot, G. Habert, N. Roussel, J.D. De Lacaillerie, Flow properties of MK-based
1093 geopolymer pastes. A comparative study with standard Portland cement pastes, Soft Matter 10 (2014)

1094 1134-1141.

1095 [34] A. Favier, G. Habert, J.D.E. de Lacaillerie, N. Roussel, Mechanical properties and

1096 compositional heterogeneities of fresh geopolymer pastes, Cem. Concr. Res. 48 (2013) 9-16.

1097 [35] H. Mehdizadeh, E.N. Kani, Rheology and apparent activation energy of alkali activated

1098 phosphorous slag, Constr. Build. Mater. 171 (2018) 197-204.

1099 [36] A.I. Laskar, R. Bhattacharjee, Rheology of fly-ash-based geopolymer concrete, ACI Mater. J.

1100 108 (2011) 536-542.

1101 [37] M. Criado, A. Palomo, A. Fernández-Jiménez, P.F.G. Banfill, Alkali activated fly ash: effect of

f
oo
1102 admixtures on paste rheology, Rheol. Acta 48 (2009) 447-455.

r
1103 [38] J. Xiang, L. Liu, X. Cui, Y. He, G. Zheng, C. Shi, Effect of limestone on rheological, shrinkage

1104
-p
and mechanical properties of alkali–activated slag/fly ash grouting materials, Constr. Build. Mater.
re
1105 191 (2018) 1285-1292.
lP

1106 [40] T. Yang, H. Zhu, Z. Zhang, X. Gao, C. Zhang, Q. Wu, Effect of fly ash microsphere on the

1107 rheology and microstructure of alkali-activated fly ash/slag pastes, Cem. Concr. Res. 109 (2018)
na

1108 198-207.
ur

1109 [41] F. Puertas, C. Varga, M.M. Alonso, Rheology of alkali-activated slag pastes. Effect of the nature

1110 and concentration of the activating solution, Cem. Concr. Compos. 53 (2014) 279-288.
Jo

1111 [42] M. Vyšvařil, E. Vejmelková, P. Rovnaníková, Rheological and mechanical properties of

1112 alkali-activated brick powder based pastes: effect of amount of alkali activator, In IOP Conference

1113 Series: Mater. Sci. Eng. 379 (2018) 012011.

1114 [43] X. Guo, H. Shi, W. Hu, F. Meng, Setting time and rheological properties of solid waste-based

1115 composite geopolymers, J. Tongji Univ. (Natural Science), 44 (2016) 1066-1070 (in Chinese with

1116 English abstract).

1117 [44] R.A. Santa, J. C. Kessler, C. Soares, H.G. Riella, Microstructural evaluation of initial

1118 dissolution of aluminosilicate particles and formation of geopolymer material, Particuology 41 (2018)

1119 101-111.

1120 [45] M. Romagnoli, C. Leonelli, E. Kamse, M.L. Gualtieri, Rheology of geopolymer by DOE
1121 approach, Constr. Build. Mater. 36 (2012) 251-258.

1122 [46] S. Yin, H. Guan, J. Hu, H. Huang, Q. Yu, Rheology and fluidity of alkali activated fly ash-slag

1123 grouting materials, J. South China Univ. Technol. (Natural Science), 47 (2019) 120-128 (in Chinese

1124 with English abstract).

1125 [47] A. Aboulayt, R. Jaafri, H. Samouh, A.C. El Idrissi, E. Roziere, R. Moussa, A. Loukili, Stability

1126 of a new geopolymer grout: rheological and mechanical performances of metakaolin-fly ash binary

1127 mixtures, Constr. Build. Mater. 181 (2018) 420-436.

1128 [48] D. Zhang, D. Wang, C. Pu, X. Lin, T. Zhang J. Xia, Influence of steel slag content on

1129 rheological properties of 3D printing geopolymer fresh pastes, J. Basic Sci. Eng. 26 (2018) 596-604

f
oo
1130 (in Chinese with English abstract).

r
1131 [49] Y. Rifaai, A. Yahia, A. Mostafa, S. Aggoun, E.H. Kadri, Rheology of fly ash-based geopolymer:

1132
-p
Effect of NaOH concentration, Constr. Build. Mater. 223 (2019) 583-594.
re
1133 [50] V.D. Glukhovsky, Soil silicate articles and structures, Budivel’nyk Publish, Russian, 1967.
lP

1134 [51] C. Shi, F. He, Types and characteristics of alkali activated cement, J. Chinese Ceram. Soc. 40

1135 (2012) 69-75 (in Chinese with English abstract).


na

1136 [52] A. Kashani, J.L. Provis, G.G. Qiao, J.S. van Deventer, The interrelationship between surface
ur

1137 chemistry and rheology in alkali activated slag paste, Constr. Build. Mater. 65 (2014) 583-591.

1138 [53] E. Nägele, U. Schneider, The zeta-potential of blast furnace slag and fly ash, Cem. Concr. Res.
Jo

1139 19 (1989) 811-820.

1140 [54] D. Zhang, D. Wang, Z. Liu, F. Xie, Rheology, agglomerate structure, and particle shape of fresh

1141 geopolymer pastes with different NaOH activators content, Constr. Build. Mater. 187 (2018)

1142 674-680.

1143 [55] M. Palacios, M.M. Alonso, C. Varga, F. Puertas, Influence of the alkaline solution and

1144 temperature on the rheology and reactivity of alkali-activated fly ash pastes, Cem. Concr. Compos.

1145 95 (2019) 277-284.

1146 [56] D. Zhang, D. Wang, F. Xie, Microrheology of fresh geopolymer pastes with different NaOH

1147 amounts at room temperature, Constr. Build. Mater. 207 (2019) 284-290.

1148 [57] F. Xie, Z. Liu, D. Zhang, J. Wang, D. Wang, Understanding the acting mechanism of NaOH
1149 adjusting the transformation of viscoelastic properties of alkali activated phosphorus slag, Constr.

1150 Build. Mater. 257 (2020) 119488.

1151 [58] H. Zhou, X. Wu, Z. Xu, M. Tang, Kinetic study on hydration of alkali-activated slag, Cem.

1152 Concr. Res. 23 (1993) 1253-1258.

1153 [59] Z. Wen, J. Yan, S. Yin, Rheological characteristics of sodium silicate solution and its effects on

1154 the performance of geopolymer slurry, J. Build. Mater. 14 (2011) 723-729 (in Chinese with English

1155 abstract).

1156 [60] J.F. Stebbins, I. Farnan, X. Xue, The structure and dynamics of alkali silicate liquids: A view

1157 from NMR spectroscopy, Chem. Geol. 96 (1992) 371-385.

f
oo
1158 [61] X. Wang, S. Yin, S. Zhao, Q. Wu, Q. Yu, Z. Wen, Study on rheological properties of alkali

r
1159 activated carbonatite slag composite grouting material, J. Yangtze River Sci. Res. Ins. 23 (2006)

1160 80-83 (in Chinese with English abstract).


-p
re
1161 [62] B. Fu, Z. Cheng, J. Han, Y. Hu, Strength and fresh properties of alkali activated metakaolin slag
lP

1162 geopolymer mortar, Bull. Chin. Ceram. Soc. 38 (2019) 4013-4020 (in Chinese with English abstract).

1163 [63] D. Zhang, D. Wang, X. Lin, T. Zhang, The study of the structure rebuilding and yield stress of
na

1164 3D printing geopolymer pastes, Constr. Build. Mater. 184 (2018) 575-580.
ur

1165 [64] M. Torres-Carrasco, C. Rodríguez-Puertas, M. del Mar Alonso, F. Puertas, Alkali activated slag

1166 cements using waste glass as alternative activators. Rheological behavior, Boletín de la sociedad
Jo

1167 española de Ceramicay Vìdrio 54 (2015) 45-57.

1168 [65] H. Mehdizadeh, E.N. Kani, A.P. Sanchez, A. Fernandez-Jimenez, Rheology of activated

1169 phosphorus slag with lime and alkaline salts, Cem. Concr. Res. 113 (2018) 121-129.

1170 [66] M. Palacios, T.F. Houst, P. Bowen, F. Puertas, Adsorption of superplasticizer admixtures on

1171 alkali-activated slag pastes, Cem. Concr. Res. 39 (2009) 670-677.

1172 [67] J.L. Provis, G.C. Lukey, J.S. van Deventer, Do geopolymers actually contain nanocrystalline

1173 zeolites? A reexamination of existing results, Chem. Mater. 17 (2005) 3075-3085.

1174 [68] E.M. Gartner, D.E. Macphee, A physico-chemical basis for novel cementitious binders, Cem.

1175 Concr. Res. 41 (2011) 736-749.

1176 [70] P.L.L. Gonzalez, R.M. Novais, J.A. Labrincha, B. Blanpain, Y. Pontikes, Modifications of
1177 basic-oxygen-furnace slag microstructure and their effect on the rheology and the strength of

1178 alkali-activated binders, Cem. Concr. Compos. 97 (2019) 143-153.

1179 [71] K. Vance, A. Kumar, G. Sant, N. Neithalath, The rheological properties of ternary binders

1180 containing Portland cement, limestone, and metakaolin or fly ash, Cem. Concr. Res. 52 (2013)

1181 196-207.

1182 [72] N. You, Y. Liu, D. Gu, T. Ozbakkaloglu, J. Pan, Y. Zhang, Rheology, shrinkage and pore

1183 structure of alkali-activated slag-fly ash mortar incorporating copper slag as fine aggregate, Constr.

1184 Build. Mater. 242 (2020) 118029.

1185 [73] J. Rouyer, A, Poulesquen, Evidence of a fractal percolating network during geopolymerization, J.

f
oo
1186 Am. Ceram. Soc, 98 (2015) 1580-1587.

r
1187 [74] P. Rovnaník, P. Rovnanikova, M. Vysvařil, S. Grzeszczyk, E. Janowskarenkas, Rheological

1188
-p
properties and microstructure of binary waste red brick powder/metakaolin geopolymer, Constr.
re
1189 Build. Mater. 188 (2018) 924-933.
lP

1190 [75] H. Güllü, A. Cevik, K.M. Al-Ezzi, M.E. Gülsan, On the rheology of using geopolymer for

1191 grouting: A comparative study with cement-based grout included fly ash and cold bonded fly ash,
na

1192 Constr. Build. Mater. 196 (2019) 594-610.


ur

1193 [76] I.M. Krieger, T.J. Dougherty, A mechanism for non‐Newtonian flow in suspensions of rigid

1194 spheres, Trans. Soc. Rheol. 3 (1959) 137-152.


Jo

1195 [77] R.J. Flatt, P. Bowen, Yodel: a yield stress model for suspensions, J. Am. Ceram. Soc. 89 (2006)

1196 1244-1256.

1197 [78] K.H. Khayat, A. Yahia, Effect of welan gum-high-range water reducer combinations on

1198 rheology of cement grout, Mater. J. 94 (1997) 365-372.

1199 [79] E. Ghafari, H. Costa, E. Julio, A. Portugal, L. Duraes, The effect of nanosilica addition on

1200 flowability, strength and transport properties of ultra high performance concrete, Mater. Des. 59

1201 (2014) 1-9.

1202 [80] M.M. Alonso, M. Palacios, F. Puertas, A.G. De la Torre, M.A.G. Aranda, Effect of

1203 polycarboxylate admixture structure on cement paste rheology, Mater. Constr. 57 (2007) 65-81.

1204 [81] J. Björnström, S. Chandra, Effect of superplasticizers on the rheological properties of cements,
1205 Mater. Struct. 36 (2003) 685-692.

1206 [82] F. Puertas, H. Santos, M. Palacios, S. Martínez-Ramírez, Polycarboxylate superplasticiser

1207 admixtures: effect on hydration, microstructure and rheological behavior in cement pastes, Adv. Cem.

1208 Res. 17 (2005) 77-89.

1209 [83] M. Criado, A. Palomo, A. Fernández-Jiménez, P.F.G. Banfill, Alkali activated fly ash: effect of

1210 admixtures on paste rheology, Rheol. Acta 48 (2009) 447-455.

1211 [84] A.I. Laskar, R. Bhattacharjee, Effect of plasticizer and superplasticizer on rheology of

1212 fly-ash-based geopolymer concrete, ACI Mater. J. 110 (2013) 513-518.

1213 [85] H. Uchikawa, D. Sawaki, S. Hanehara, Influence of kind and added timing of organic admixture

f
oo
1214 on the composition, structure and property of fresh cement paste, Cem. Concr. Res. 25 (1995)

r
1215 353-364.

1216
-p
[86] T. Luukkonen, Z. Abdollahnejad, K. Ohenoja, P. Kinnunen, M. Illikainen, Suitability of
re
1217 commercial superplasticizers for one-part alkali-activated blast-furnace slag mortar, J. Sustain.
lP

1218 Cement-Based Mater. 8 (2019) 244-257.

1219 [87] M. Palacios, F. Puertas, Stability of superplasticizer and shrinkage-reducing admixtures Stability
na

1220 of superplasticizer and shrinkage-reducing admixtures in high basic media, Mater. Constr. 54 (2004)
ur

1221 65-86.

1222 [88] M. Palacios, F. Puertas, Effect of superplasticizer and shrinkage-reducing admixtures on


Jo

1223 alkali-activated slag pastes and mortars, Cem. Concr. Res. 35 (2005) 1358-1367.

1224 [89] J. Ren, Y. Bai, M.J. Earle, C.H. Yang, A preliminary study on the effect of separate addition of

1225 lignosulfonate superplasticiser and sodium silicates on the rheological behavior of alkali-activated

1226 slags, The Third International Conference on Sustainable Construction Materials & Technologies,

1227 (2013) 1-11.

1228 [90] L. Xu, F. Matalkah, P. Soroushian, N. Darsanasiri, S. Hamadneh, W. Wu, Effects of citric acid

1229 on the rheology, hydration and strength development of alkali aluminosilicate cement, Adv. Cem.

1230 Res. 30 (2018) 75-82.

1231 [91] H. Tan, X. Zheng, C. Duan, B. Xia, Polylactic acid improves the rheological properties, and

1232 promotes the degradation of sodium carboxymethyl cellulose-modified alkali-activated cement,


1233 Energy, 9 (2016) 823.

1234 [92] C.K. Park, M.H. Noh, T.H. Park, Rheological properties of cementitious materials containing

1235 mineral admixtures, Cem. Concr. Res. 35 (2005) 842-849.

1236 [93] N.J. Wagner, J.F. Brady, Shear thickening in colloidal dispersions, Phys. Today 62 (2009) 27-32.

1237 [94] Z. Zhang, J.L. Provis, J. Zou, A. Reid, H. Wang, Toward an indexing approach to evaluate fly

1238 ashes for geopolymer manufacture, Cem. Concr. Res. 85 (2016) 163-173.

1239 [95] Z. Zhang, Preparation, properties and reaction mechanism of metakaolin based inorganic

1240 polymer, Diss. Nanjing University of technology 2010 66-70.

1241 [96] L. Qin, B. Q, C. Shi, Z. Zhang, Effect of Ca/Si ratio on the formation and characteristics of

f
oo
1242 synthetic aluminosilicate hydrate gels, Mater. Rept. 34 (2020) 61-67.

r
1243 [97] A. Wetzel, B. Middendorf, Influence of silica fume on properties of fresh and hardened

1244
-p
ultra-high performance concrete based on alkali-activated slag, Cem. Concr. Compos. 100 (2019)
re
1245 53-59.
lP

1246 [98] Y. Liu, C. Shi, Z. Zhang, N. Li, D. Shi, Mechanical and fracture properties of ultra-high

1247 performance geopolymer concrete: effects of steel fiber and silica fume, Cem. Concr. Compos. 112
na

1248 (2020) 103665.


ur

1249 [99] B. Panda, S. Ruan, C. Unluer, M.J. Tan, Investigation of the properties of alkali-activated slag

1250 mixes involving the use of nanoclay and nucleation seeds for 3D printing, Compos. Part B: Eng. 186
Jo

1251 (2020) 107826.

1252 [100] M.P. Luxán, R. Sotolongo, F. Dorrego, E. Herrero, Characteristics of the slags produced in the

1253 fusion of scrap steel by electric arc furnace, Cem. Concr. Res. 30 (2000) 517-519.

1254 [101] A. Papo, L. Piani, R. Ricceri, Rheological properties of very high-strength Portland cement

1255 pastes: influence of very effective superplasticizers, Inter. J. Chem. Eng. (2010) 1-7.

1256 [102] H. Alghamdi, S.A. Nair, N. Neithalath, Insights into material design, extrusion rheology, and

1257 properties of 3D-printable alkali-activated fly ash-based binders, Mater. Des. 167 (2019) 107634.

1258 [103] M. Westerholm, B. Lagerblad, J. Silfwerbrand, E. Forssberg, Influence of fine aggregate

1259 characteristics on the rheological properties of mortars, Cem. Concr. Compos. 30 (2008) 274-282.

1260 [104] S. Chithra, S.S. Kumar, K. Chinnaraju, The effect of colloidal nano-silica on workability,
1261 mechanical and durability properties of high performance concrete with copper slag as partial fine

1262 aggregate, Constr. Build. Mater. 113 (2016) 794-804.

1263 [105] Y. Liu, Z. Zhang, C. Shi, D. Zhu, Y. Deng, Development of ultra-high performance geopolymer

1264 concrete (UHPC): influence of steel fiber on mechanical properties, Cem. Concr. Compos. (2020)

1265 103670.

1266 [106] S.J. Choi, J.L. Choi, J.K. Song, B.Y. Lee, Rheological and mechanical properties of

1267 fiber-reinforced alkali-activated composite, Constr. Build. Mater. 96 (2015) 112-118.

1268 [107] C. Shi, A.F. Jiménez, A. Palomo, New cements for the 21st century: The pursuit of an

1269 alternative to Portland cement, Cem. Concr Res. 41 (2011) 750-763.

f
oo
1270 [108] L.G. Li, A. K. H. Kwan, Mortar design based on water film thickness, Constr. Build. Mater. 25

r
1271 (2011) 2381-2390.

1272
-p
[109] A. K. H. Kwan, L.G. Li, Combined effects of water film thickness and paste film thickness on
re
1273 rheology of mortar, Mater. Struct. 45 (2012) 1359-1374.
lP

1274 [110] A. K. H. Kwan, L.G. Li, Concrete mix design based on water film thickness and paste film

1275 thickness, Cem. Concr. Compos. 39 (2013) 33-42.


na

1276 [111] J. Davidovits, Geopolymers, J. Therm. Anal. 37 (1991) 1633-1656.


ur

1277 [112] G. Habert, J.B. d’Espinose de Lacaillerie, N. Roussel, An environmental evaluation of

1278 geopolymer based concrete production: reviewing current research trends, J. Clean. Prod. 19 (2011)
Jo

1279 1229-1238.

1280 [113] M. Palacios, S. Gismera, M.M. Alonso, J.B. d’Espinose de Lacaillerie, B. Lothenbach, A.

1281 Favier, C. Brumaud, F. Puertas, Early reactivity of sodium silicate-activated slag pastes and its

1282 impact on rheological properties, Cem. Concr. Res. 140 (2021) 106302.

1283 [114] P. Hou, T. R. Muzenda, Q. Li, H. Chen, X. Cheng, Mechanisms dominating thixotropy in

1284 limestone calcined clay cement (LC3), Cem. Concr. Res. 140 (2021) 106316.

1285 [115] M. Palacios, M. Kappl, M. Stuer, U. Aschauer, P. Bowen, H. J. Butt, Repulsion forces of

1286 superplasticizers on ground granulated blast furnace slag in alkaline media, from afm measurements

1287 to rheological properties, Materiales de Construcción, 62 (2012) 489-513.

1288 [116] A. Zingg, F. Winnefeld, L. Holzer, J. Pakusch, S. Becker, L. Gauckler, Adsorption of


1289 polyelectrolytes and its influence on the rheology, zeta potential, and microstructure of various

1290 cement and hydrate phases, J. Coll. Interface Sci. 323 (2008) 301-312.

1291 [117] M. Romagnoli, F. Andreola, Mixture of deflocculants: a systematic approach, J. Eur. Ceram.

1292 Soc. 27 (2007) 1871-4.

1293 [118] H. Ye, C. Fu, A. Lei, Mitigating shrinkage of alkali-activated slag by polypropylene glycol

1294 with different molecular weights, Constr. Build. Mater. 245 (2020) 118478.

1295 [119] J. Xie, O. Kayali, Effect of superplasticiser on workability enhancement of class F and class C

1296 fly ash-based geopolymers, Constr. Build. Mater. 122 (2016) 36-42.

1297 [120] T. Conte, J. Plank, Impact of molecular structure and composition of polycarboxylate comb

f
oo
1298 polymers on the flow properties of alkali-activated slag, Cem. Concr. Res. 116 (2019) 95-101.

r
1299 [121] L. Lei, H. K. Chan, Investigation into the molecular design and plasticizing effectiveness of

1300
-p
hpeg-based polycarboxylate superplasticizers in alkali-activated slag, Cem. Concr. Res. 136 (2020)
re
1301 106150.
lP

1302 [122] D. Marchon, U. Sulser, A. Eberhardt, R.J. Flatt, Molecular design of comb-shaped

1303 polycarboxylate dispersants for environmentally friendly concrete, Soft Matter 9 (2013) 10719–
na

1304 10728.
ur

1305 [123] G. Landrou, C. Brumaud, F. Winnefeld, R. J. Flatt, G. Habert, Lime as an anti-plasticizer for

1306 self-compacting clay concrete, Materials 9 (2016) 330.


Jo

1307 [124] KC. Newlands, M. Foss, T. Matchei, J. Skibsted, DE. Macphee, Early stage dissolution

1308 characteristics of aluminosilicate glasses with a blast furnace and fly-ash-like composition, J Am

1309 Ceram Soc. 100 (2017) 1941-1955.

1310 [125] T. Luukkonen, Z. Abdollahnejad, J. Yliniemi, P. Kinnunen, M. Illikainen, One-part

1311 alkali-activated materials: a review, Cem. Concr. Res. 103 (2018) 21-34.

1312 [126] L. Li, J. X. Lu, B. Zhang, C. S. Poon, Rheology behavior of one-part alkali activated slag/glass

1313 powder (AASG) pastes, Constr. Build. Mater. 258 (2020) 120381.

1314 [127] Y. Alrefaei, Y. S. Wang, J. G. Dai, The effectiveness of different superplasticizers in ambient

1315 cured one-part alkali activated pastes, Cem. Concr. Compos. 97 (2019) 166-174.

1316 [128] I. G, B. Singh, S. Deshwal, S.K. Bhattacharyyac, Effect of sodium carbonate/sodium silicate
1317 activator on the rheology, geopolymerization and strength of fly ash/slag geopolymer pastes, Cem.

1318 Concr. Compos. 97 (2019) 226-238.

1319 [129] B. Akturk, A. B. Kizilkanat, N. Kabay, Effect of calcium hydroxide on fresh state behavior of

1320 sodium carbonate activated blast furnace slag pastes, Constr. Build. Mater. 212 (2019) 388-399.

1321 [130] C. Montes, D. Zang, E. N. Allouche, Rheological behavior of fly ash-based geopolymers with

1322 the addition of superplasticizers, J. Sustain. Cement-Based Mater. 1 (2012) 179-185.

1323 [131] Y. Wang, J. L. Provis, J. Dai, Role of soluble aluminum species in the activating solution for

1324 synthesis of silico-aluminophosphate geopolymers, Cem. Concr. Compos. 93 (2018) 186-195.

1325 [132] Y. Alrefaei, Y. Wang, J. Dai, Q. Xu, Effect of superplasticizers on properties of one-part

f
oo
1326 Ca(OH)2/Na2SO4 activated geopolymer pastes, Constr. Build. Mater. 241 (2020) 117990.

r
1327 [134] C. Kuenzel, L. Li, L. Vandeperre, A. R. Boccaccini, C. R. Cheeseman, Influence of sand on the

1328
-p
mechanical properties of metakaolin geopolymers, Constr. Build. Mater. 66 (2014) 442-446.
re
1329 [135] D. Zhang, K. Zhao, F. Xie, H. Li, D. Wang, Effect of water-binding ability of amorphous gel
lP

1330 on the rheology of geopolymer fresh pastes with the different NaOH content at the early age, Constr.

1331 Build. Mater. 261 (2020) 120529.


na

1332 [136] X. Dai, S. Aydına, M. Y. Yardımcıa, K. Lesage, G. D. Schuttera, Effects of activator properties
ur

1333 and GGBFS/FA ratio on the structural build-up and rheology of AAC, Cem. Concr. Res. 138 (2020)

1334 106253.
Jo

1335 [137] P. Steins, A. Poulesquen, O. Diat, F. Frizon, Structural evolution during geopolymerization

1336 from an early age to consolidated material, Langmuir 28 (22) (2012) 8502–8510.

1337 [138] D. Rieger, J. Kadlec, M. Pola, T. Kovářík, P. Franče, Mechanical properties of non-woven glass

1338 fiber geopolymer composites, IOP Conf. 175 (2017) 012054.

1339 [139] T. Revathi, R. Jeyalakshmi, Fly ash–GGBS geopolymer in boron environment: A study on

1340 rheology and microstructure by ATR FT-IR and MAS NMR, Constr. Build. Mater. 267 (2021)

1341 120965.

1342 [140] A. Hasnaoui, E. Ghorbel, G. Wardeh, Optimization approach of granulated blast furnace slag

1343 and metakaolin based geopolymer mortars, Constr. Build. Mater. 198 (2019) 10–26.

1344 [141] J.G. Jang, N.K. Lee, H.K. Lee, Fresh and hardened properties of alkali-activated fly ash/slag
1345 pastes with superplasticizers, Constr. Build. Mater. 50 (2014) 169–176.

f
r oo
-p
re
lP
na
ur
Jo
Statement of Conflict of Interest
We have no conflict financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

f
r oo
-p
re
lP
na
ur
Jo

You might also like