You are on page 1of 41

Journal Pre-proof

The role of limestone and calcined clay on the rheological properties of LC3

Tafadzwa Ronald Muzenda, Pengkun Hou, Shiho Kawashima, Tongbo Sui, Xin
Cheng

PII: S0958-9465(20)30007-X
DOI: https://doi.org/10.1016/j.cemconcomp.2020.103516
Reference: CECO 103516

To appear in: Cement and Concrete Composites

Received Date: 30 May 2019


Revised Date: 6 January 2020
Accepted Date: 7 January 2020

Please cite this article as: T.R. Muzenda, P. Hou, S. Kawashima, T. Sui, X. Cheng, The role of limestone
and calcined clay on the rheological properties of LC3, Cement and Concrete Composites, https://
doi.org/10.1016/j.cemconcomp.2020.103516.

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.

© 2020 Elsevier Ltd. All rights reserved.


1 The role of limestone and calcined clay on the rheological properties of LC3

2 Tafadzwa Ronald Muzenda1, Pengkun Hou1※, Shiho Kawashima2※, Tongbo Sui3, Xin Cheng1※

3 1. Shandong Provincial Key Lab for Preparation and Measurement of Building Materials,

4 University of Jinan, Jinan, 250022, China

5 2. Civil Engineering and Engineering Mechanics, Columbia University, New York, 10027, USA

6 3. SINOMA International Engineering Co. Ltd., Beijing, 100102, China

8 Abstract: Understanding the rheological properties of cementitious materials is important for

9 controlling their fresh properties and for improving the practical applications. In this study, the

10 rheological properties of pastes made from ordinary Portland cement blended with different

11 amounts of calcined clay and limestone were investigated in order to understand the combined

12 and independent effects of limestone and calcined clay on the rheological properties of limestone

13 calcined clay cement (LC3). Large Amplitude Oscillation Strain was applied and the subsequent

14 harmonic distortion was used to evaluate nonlinear response for the first time for cementitious

15 materials. The results showed that calcined clay leads to increased static and dynamic yield stress,

16 initial thixotropic index, plastic viscosity and cohesion, as well as decreased harmonic distortion,

17 while limestone has an opposite effect. It is believed that these results will aid in understanding

18 the viscoelasticity of this blended cement.

19 Key words: calcined clay, limestone, cement, rheological properties, nonlinear behavior

Email addresses: pkhou@163.com (P.K. Hou), sk2294@columbia.edu (S. Kawashima), chengxin@ujn.edu.cn (X. Cheng).

1
20 1. Introduction

21 Clinker substitution using supplementary cementitious materials (SCMs) is becoming

22 increasingly prominent with the demand to cut down on CO2 emissions in the cement industry

23 [1-4]. In order for clinker substitution to remain practical and effective, it is imperative that we

24 use SCMs with the potential of meeting high demand while yielding desired chemical and

25 mechanical properties. Clay containing kaolinite is abundantly available in the earth’s crust and

26 when calcined and used together with limestone, it yields comparable physiochemical properties

27 and better durability when compared with pure ordinary Portland cement (OPC) even at 50%

28 clinker substitution [5-9]. This renders limestone calcined clay cement (LC3) a viable alternative

29 for the future of green cement.

30 One of the main differences between LC3 and OPC is rheological behavior, which greatly

31 influences practical application. Rheology plays a key role in the application of cementitious

32 materials and in achieving successful concrete performance. As noted by Ferraris et al. [10],

33 rheological properties of concrete need to be carefully balanced to achieve specific goals such as

34 limiting segregation, producing a good surface finish, minimizing pumping pressure, or

35 controlling formwork pressure. Hence, making LC3 cement suitable for practical application

36 requires a thorough understanding of its rheological properties [11,12].

37 The rheological properties of OPC systems containing clay have been studied to some extent

38 in the past [13-17] but not much focus has been afforded to systems with calcined clay and the

39 LC3 system with high clinker substitution. Kawashima et al., evaluated the influence of small

40 additions of highly-purified attapulgite clays [14]. They concluded that clay increases the

41 cohesion and viscosity of the pastes, as shown by the trend of the peak force of the tack test. In

2
42 another study, Kawashima et al. reported that purified attapulgite clays significantly accelerate

43 the rate of recovery of pastes, especially at early ages, and that this accelerating effect diminishes

44 at longer resting times as hydration of cement begins to dominate [13].

45 It has been reported that limestone filler alone is not able to prevent segregation or bleeding,

46 and that there is no difference between cement pastes mixed with limestone filler and pure OPC

47 pastes in terms of rheology [15]. On the other hand, if one needs low slump and low spread, the

48 use of metakaolin is recommended because this material creates a strong interconnected network

49 in the paste, increasing the yield stress and viscosity of the cement paste. Tregger et al., after

50 comparing the rheological effect of different clay admixtures and other SCMs like fly ash,

51 concluded that clay significantly increased the shape stability of cement pastes [16].

52 One of the key in-depth studies on the rheology of LC3 systems was conducted by Ferreiro

53 et al. [17]. The authors investigated the workability and strength performance of blended

54 cements with different replacement levels (up to 65%) for two calcined clay mineral additions

55 (1:1 and 2:1 type) and mixtures with limestone coupled with the effect of fly ash addition and

56 different water-to-cement ratios. They reported that workability is strongly affected by the

57 calcined clay content, particularly for the 1:1 clay.

58 To the best of the authors’ knowledge, not much focus has been afforded to the independent

59 and combined effect of limestone and calcined clay on the rheological properties of OPC pastes

60 with high clinker substitution of up to 50%. In fact, the aim to utilize LC3 at 50% clinker

61 substitution makes studying the effect of high clinker substitution on the rheological properties

62 necessary research. This motivated the present study, which aims to provide more insight into the

63 rheological properties of the LC3 system, as well as the independent effect of limestone (LS) and

64 calcined clay (CC).

3
65 2. Materials and methods

66 2.1 Materials

67 The materials used were clinker, calcium sulphate, limestone and calcined clay. Table 1

68 shows the chemical composition of the materials. The clinker and limestone were ground

69 separately, after which they were sieved with a 149 µm sieve. Commercially available calcined

70 clay from Maoming, China was used after further grinding. The particle size distribution of the

71 clinker, limestone (LS) and calcined clay (CC) as well as the REF and the 7 blends involving LS

72 and CC (Table 2) were analyzed by laser diffraction using the LS Particle Size Analyzer and they

73 are shown in Fig. 1. The specific surface area (Table 2) was analyzed using the BET method

74 according to the internal laboratory procedure. The SEM images of the clinker, calcined clay and

75 limestone are shown in Fig. 2.

76
4
77 Fig. 1. Particle size distribution of clinker, limestone calcined clay and the 8 samples used in this

78 study.

79 Table 1. Chemical compositions of raw materials.

Raw materials Clinker (%) Limestone (%) Calcined Clay (%)

Ca0 65.62 53.59 0.45

Al2O3 4.01 0.84 44.13

SiO2 20.10 2.18 51.27

SO3 0.47 0.15 0.08

Fe2O3 3.39 0.86 1.10

MgO 2.57 1.14 0.49

TiO2 0.25 0.07 0.12

K2O 0.71 0.45 2.09

Na2O 0.10 0.12 0.17

Cl 0.05 0.07 0.04

SrO 0.03 0.04 -

P2O5 - 0.05 0.03

LOI 2.70 40.44 0.03

Specific surface area 316 536 573

(m2/kg)

80

5
81
82 (a) (b)

83
84 (c)
85 Fig. 2. Particle shapes for (a) clinker, (b) calcined clay and (c) limestone at a magnification of
86 20,000.
87

88 2.2 Binder formulations

89 Up to 8 different binders were prepared for this study (Table 2). Gypsum was kept

90 constant at 5% while the amount of clinker varied depending on the amount of CC and LS. The

6
91 ratio of CC to LS was maintained at 2:1 (the ratio currently in use for practical LC3 applications)

92 in the first four blends, which were employed to investigate the variation in rheological

93 properties with the increase in both LS and CC. The last four blends contained either calcined

94 clay or limestone only, with either 80 or 65% clinker. These were employed to study the

95 independent effects of LS and CC. All binders were prepared with a water-to-binder ratio of 0.45

96 to avoid segregation in the control during shearing. The pastes were prepared in a cement paste

97 mixer according to the following mixing protocol: slow mixing for 120 s, 15 s rest and fast

98 mixing for 120 s. The tests were carried out after a total of 7 minutes from the time of initial

99 contact of the binder with water depending on the protocol employed (Table 3).

100 Table 2. Binder formulations (SCM blend abbreviation: clinker content (C_)/CC content

101 (Cc_)/LS content (L_)).

Clinker (%) Calcined clay (%) Limestone (%) Gypsum (%)

REF 95 0 0 5

C80/Cc10/L5 80 10 5 5

C65/Cc20/L10 65 20 10 5

C50/Cc30/L15 50 30 15 5

C80/Cc15/L0 80 15 0 5

C80/Cc0/L15 80 0 15 5

C65/Cc30/L0 65 30 0 5

C65/Cc0/L30 65 0 30 5

102

103 2.3 Rheological protocols

7
104 The Kinexus Lab+ rotational rheometer (Malvern Panalytical Ltd) was used for all the

105 tests in this study (Fig. 3a). The rheometer operates on direct strain control, shear rate control or

106 shear stress control, with a maximum torque of 200 mNm. Vane-in-cup, and parallel-plate

107 geometries (Fig. 3b) with the inner diameter of the cup and the top plate both being 40 mm and

108 the vane blades having a diameter of 25 mm were used. Different protocols were employed in

109 order to carefully analyze different rheological properties and, as an illustration, the parallel plate

110 setup is shown in Fig. 2c and the general equipment setup is shown in Fig. 3d.

111

112 (a) (b)

8
113

114 (c) (d)

115 Fig. 3. Rheometer equipment and experimental setup.

116 2.3.1 Logarithmic stress ramp

117 Cement paste, mortar and concrete are yield stress materials, that is to say some minimal

118 shear stress is required to initiate flow. This stress is referred to as static yield stress and it arises

119 from the ability of the microstructure of the cementitious materials to resist flow up to a certain

120 limit, after which it is broken down. The yield stress of a cement paste originates from the

121 microstructure of a particle-particle network through colloidal interaction or direct contact [26].

122 Several methods can be employed to study the static yield stress of a cement paste; for example,

123 the creep cycles [27] and stress- or strain-controlled ramp [28]. In this study, the logarithmic

124 stress ramp (LSR), a form of the stress-controlled test, was employed with a gap of 1 mm

125 between the cup surface and the vane.

126 The protocol is described below:

9
127 • After preparation(a total of about 7 minutes from initial contact of water with the binder),

128 the paste was pre-sheared from 0 to 100 s-1 in 60 s.

129 • After preshearing, the LSR was run after the respective resting times until the shear strain

130 was about 0.1 s-1 (based on other work [29]). The stress corresponding to 0.1 s-1 was

131 considered to be the static yield stress.

132 • To study the effect of resting time, the paste was allowed to rest for 2, 30 and 60 min

133 before the logarithmic stress ramp was performed. The samples were covered using the

134 rheometer geometry cover in between the time intervals to avoid water loss.

135

136 2.3.2 Three step flow test

137 The common three step flow test protocol (Fig. 4a) [30-32] was employed to study

138 dynamic yield stress and plastic viscosity. It was performed with the vane-in-cup geometry at a

139 gap of 1 mm. The surface of the cup was not smooth, which helped to avoid slippage [33].

140 Dynamic yield stress is a measure of the stress required to terminate flow. It corresponds to a

141 broken microstructure. Different rheological models can be employed for cementitious materials,

142 the most common being the Bingham model and the Hershel-Bulkley model. Where relative

143 linearity is observed, the Bingham model is used because of its simplicity and easy-to-determine

144 parameters, and such was the case with the results in this study. The results also did not show

145 any shear thinning or thickening, which would make the curve nonlinear and necessitate the use

146 of the Hershel-Bulkley model [33]. Hence the Bingham (linear) model (Fig. 4b) was utilized to

147 determine dynamic yield stress and plastic viscosity from the plot of shear stress vs. shear strain

148 for the down curve (Fig. 4a).

10
149

150

151 Fig. 4. a) Three step flow test protocol and b) illustration of Bingham model and its parameters.

152 2.3.3 Tack test

11
153 The tack test is used to evaluate the cohesion and adhesion of the cement pastes. The test,

154 as illustrated in [14], works with the parallel plate geometry and it involves fixing a certain

155 velocity at which the top plate pulls away from the bottom plate. In the current study, the test

156 was performed with an initial vertical lift velocity of 0.1 µms-1 (minimum possible velocity) with

157 a fixed linear acceleration of 1 mm/s2. The normal force evolution was monitored and the peak

158 force reached was considered to be a measure of adhesive properties of the samples. The initial

159 gap was 1 mm and each paste sample was subjected to a pre-shear of 50 s-1 before the tack test to

160 ensure that all samples start at the same reference state after preparation. The bottom and upper

161 plates were covered by a 150-grit sand paper to avoid slippage during the test (Fig. 3d).

162 2.3.4 Amplitude oscillation rheometry

163 A large amplitude oscillation strain (LAOS) protocol in which an increasing logarithmic

164 strain ramp is applied to the sample was used to evaluate nonlinear response. It is known that in

165 practical applications large deformation and rapid flow are very common and this brings about a

166 need to quantify the resultant nonlinear response [18,19]. This has led to renewed interest in

167 large amplitude oscillatory shear/strain (LAOS), whereas in the recent past most viscoelasticity

168 studies have focused on small amplitude oscillatory shear/strain (SAOS) [20-22]. SAOS can

169 provide important information on the influence of hydration on the developing rigid structure (i.e.

170 stiffness) of fresh cement-based systems [23]. However, it somewhat comes short of reality, as

171 most processing operations involve deformations that can be large and rapid, hence, nonlinear

172 material properties control the system response [18]. LAOS tests are substantially more complex

173 rheological probes than SAOS tests because of the non-sinusoidal and nonlinear system response.

174 This necessitates a relatively simple approach to quantify non-linear response and one such

175 approach is provided in the parameter termed harmonic distortion. Harmonic distortion is a

12
176 common phenomenon in electrical engineering where it is used to quantify unwanted nonlinear

177 conduction of electricity [24,25]. Here, this parameter is being employed for the first time for

178 cementitious materials and it appears to be a very handy tool in evaluating nonlinear response for

179 many practical applications.

180 Non-linear behavior can be understood to result in differences between the sinusoidal

181 input and the resulting output, and this can be quantified through harmonic distortion. This

182 parameter can be defined as a measure of the non-linearity of the sample, that is how alike the

183 input and output waves are—high numbers mean that the output (measured sine wave) is not the

184 same shape as the input sine wave. In the linear viscoelastic region, the input strain and output

185 stress are both sinusoidal, which means that the harmonic distortion is low (<< 1%). Outside the

186 linear region, the output will display higher harmonics than the fundamental (input) frequency.

187 Harmonic distortion can help in evaluating the distortions of the sinusoidal input in the output,

188 which corresponds to the distortion of the well-connected microstructure of the sample.

189 In oscillatory measurements, if the strain amplitude is sufficiently small, the output strain

190 and rate of strain are also sinusoidal and respond with the same frequency, but generally with a

191 phase lag [19]. At larger amplitudes, higher harmonics might occur, whereby only the odd

192 harmonics are caused by the rheological response of the sample. The occurrence of higher

193 harmonic contributions can be used as a sensitive tool to characterize crossover from linear to

194 non-linear behavior in an amplitude sweep, and as soon as higher harmonics are occurring, the

195 values of the moduli will change as well [35]. The Fourier transform was proposed in 2008 by

196 Edwolt et al. as a means to analyze the odd harmonics resulting from nonlinear response and in

197 turn the viscoelastic properties of the material using the MITlaos software [36]. This approach

198 was utilized for a cement system for the first time in 2016 [19]. As an alternative approach, we

13
199 are considering the use of harmonic distortion (HD) to quantify the nonlinear response. The HD

200 value is the root mean squared of the odd harmonics 3-13 and it is given in the output of the

201 Kinexus rheometer. HD adequately covers the major odd harmonics that occur in rheology.

202 Although the strain, which varies from about 0.007 to 100%, is in fact the independent

203 variable, a plot of the elastic modulus (or amplitude of viscosity) can be built as a function of the

204 applied/corresponding shear stress. It is also worth noting, from Fig. 5, that for all samples the

205 elastic modulus is greater than the viscous modulus until a certain point in the nonlinear response

206 region where the viscous modulus becomes greater. This represents a strain-induced transition

207 from a viscoelastic semi-solid to a viscous liquid [34], and a breakdown of the microstructure of

208 the paste.

209

210 Fig. 5. Linear to non-linear transition of C80/Cc10/L5.

14
211 Table 3. Summary of sample preparation and testing protocols.

Protocol Rheological Sample Preshearing Protocol details Time periods for

name property preparation and measurements

geometry

Logarithmic Static yield Slow 100 s-1; Logarithmic increment of 2 min (after

stress ramp stress mixing for Vane-in- shear stress until a strain preshearing);

120 s cup of 0.1 s-1. 30 min;

60 min

Up-down Dynamic 100 s-1; Fig. 3; 2 min;


Rest for 15
flow curve yield stress Vane-in- Preshearing at every time 30 min;
s
cup interval. 60 min

Tack Adhesion Fast 50 s-1; Normal force applied on Following preshearing.

and mixing for Parallel top plate in the upward The test lasted for a

cohesion 120 s plates direction. most a few minutes.

Large Harmonic 100 s-1; Temperature stabilization Following preshearing.

amplitude distortion Vane-in- for 5 min; The protocol required

oscillation and yield cup Logarithmic strain ramp about 20 minutes.

strain stress from 0.007 s-1 to 100 s-1.

212

213

15
214 3. Results and discussion

215 3.1 Static yield stress

216 3.1.1 Increasing limestone and calcined clay

217 The influence of increasing the combined substitution of limestone (LS) and calcined

218 clay (CC) from 0 to 45% on the static yield stress (SYS) at 2 min resting time can be seen in Fig.

219 6a. There is a clear increase in the static yield stress (SYS) from the REF to C50/Cc30/L15, with

220 the SYS of C50/Cc30/L15 being over 15 times that of the REF. SYS increased with resting time

221 for all samples (Fig. 6a) but at different rates which reflects the different mechanisms at work in

222 the different systems. Overall, the rate of increase was consistent with the exception of

223 C80/Cc10/L5 and REF at 60 min resting time, where the SYS of REF went on to exceed that of

224 C80/Cc10/L5. Increase in clinker substitution by the less reactive calcined clay and limestone for

225 LC3 samples results in a dilution effect which leads to relatively lower SYS increase with resting

226 time as compared to REF. This phenomenon is further illustrated in the thixotropy index

227 discussed in the next section. Overall, the results correspond well with past studies [13,17,37],

228 where increase in resting time and the combined addition of CC and LS increases the SYS for

229 Portland cement systems. A further study on the link between the rheological results presented in

230 this paper and the pore structure, volume fraction and hydration mechanisms has been conducted.

231

232 3.1.2 Independent effect of limestone and calcined clay

233 In order to better understand the independent effect of each SCM on the SYS of LC3

234 systems, we refer to Fig. 6. In generalized terms, the SYS of C80/Cc10/L5 and C65/Cc20/L10

235 were greater than that of C80/Cc0/L15 and C65/Cc0/L30 but less than that of C80/Cc15/L0 and

16
236 C65/Cc30/L0 at all resting times, respectively. Therefore, in general, it is calcined clay that is

237 responsible for the increase in SYS with increase in SCM content in the LC3 system. However,

238 there was an exception when C80/Cc10/L5 was greater than C80/Cc15/L0 at 60 minutes resting

239 time. This is likely because the combined effect (synergy) of CC and LS on cement hydration

240 supersedes that of calcined clay and limestone when employed separately [5,38], resulting in

241 greater increase in SYS for C80/Cc10/L5. Apart from its contribution to hydration, limestone

242 also moderately affects SYS increase over time due to dissolution-precipitation processes leading

243 to the formation of links between grains, as illustrated in limestone paste [39].

244 A case in point is that only C65/Cc20/L10 and C65/Cc0/L30 are shown in Fig. 6c

245 because the torque required to reach the SYS for C65/Cc30/L0, even after a short resting time (2

246 min), exceeded the rheometer’s maximum limit of 200 mNm. This highlights the marked effect

247 of CC on SYS. C50/Cc30/L15 has the same content of CC as C65/Cc30/L0 (30%) but the former

248 flowed while the latter did not. This can be attributed to the 15% LS content in C50/Cc30/L15.

249 At 2 min resting time, C80/Cc15/L0 yielded a 67% increase in SYS while C80/Cc0/L15 yielded

250 a 69% decrease in comparison to C80/Cc10/L5. Therefore, it could be concluded that limestone

251 significantly reduces SYS, while calcined clay has an opposite effect. One reason for this

252 observation is the difference in the particle shapes of the two SCMs. Limestone has a somewhat

253 round or circular shape that provides a ‘bearing effect’ in the paste and thus increasing

254 flowability whereas calcined clay has a layered structure which contributes to its high water

255 demand. Similar results were observed by other researchers in different OPC systems [17,39].

17
256

257

18
258

259 Fig. 6. (a) The variation of SYS with resting time, and comparison of the effect of limestone and

260 calcined clay on SYS at (b) 80wt.% substitution and (c) 65wt.% substitution (C65/Cc30/L0 did

261 not flow at the rheometer’s maximum torque of 200mNm).

262 3.2 Dynamic yield stress and plastic viscosity

263 Flow curves of different LC3 and REF samples at different resting times were obtained

264 and used to determine the Bingham model parameters, namely dynamic yield stress (DYS) and

265 plastic viscosity (Fig. 7). Results show that DYS and plastic viscosity both increase with increase

266 in SCM content and resting time. Clays are well known to act as viscosity modifiers to enhance

267 flocculation strength [20], therefore an increase in calcined clay content is expected to increase

268 both DYS and plastic viscosity. And although limestone can act as a filler to increase viscosity at

269 certain addition levels, from the results of the previous section on SYS, it is expected that the

19
270 effect of calcined clay governs here. Also, the increase over time for all mixes is consistent with

271 REF, as this is primarily tied to the progression of cement hydration.

272 To gain further insight into how the SCMs are influencing structural build-up and

273 breakdown over time, SYS and DYS are compared directly. Qian and Kawashima studied the

274 relationship between static and dynamic yield stress for mortar and supported that the reason

275 why static yield stress is higher than dynamic yield stress is because SYS corresponds to an

276 undisturbed, well-connected microstructure while DYS corresponds to a broken-down

277 microstructure [26]. Thixotropy is assumed to play a big role in the discrepancy. Table 4 shows

278 the difference between SYS and DYS for all samples and resting times, which corresponds to the

279 thixotropy of each sample at the given resting time. A thixotropy index is utilized to quantify the

280 relationship between SYS and DYS, as follows:

281 Thixotropy index = (SYS – DYS)/DYS

282 A higher thixotropy index indicates higher degree of thixotropy, i.e. structural breakdown and

283 recovery.

284 At very early ages, thixotropy generally increases with the increase in SCM content,

285 indicated by the higher thixotropy index values at 2 min for the LC3 mixes compared to REF. As

286 noted earlier, calcined clay is responsible for the increase in SYS with increase in SCM content.

287 In addition, clays can exhibit significant shear thinning. Hence, calcined clay leads to an increase

288 in the thixotropy of LC3. Parallel conclusions were reached in other studies involving clay

289 [13,15,17], and this was largely attributed to flocculation and high-water demand. Interestingly,

290 it can also be seen from Table 4 that the effect of calcined clay on thixotropy is most pronounced

291 at the earlier stages (2 min) but diminished at 30 and 60 min compared to the REF, as hydration

20
292 mechanisms became dominant [13]. This can mainly be attributed to the significant increase in

293 SYS of REF over time – 27 Pa at 2 min to 559 Pa at 60 min – as the influence of hydration in the

294 early age will be greatest for REF. This is in contrast to DYS of REF – 21.6 Pa at 2 min to 35.1

295 Pa at 60 min – which is obtained under shearing conditions (i.e. flow curve) and early hydrates

296 can be broken if sufficient energy is introduced.

297 It is interesting to note the sharp increase in plastic viscosity for REF as compared to the

298 other three blends (Fig. 7b), similar to SYS and DYS. Plastic viscosity changes by 183%, 136%,

299 129% and 125% from 2 min to 60 min for REF, C80/Cc10/L5, C65/Cc20/L10 and

300 C50/Cc30/L15, respectively. Even though the plastic viscosity of REF at 60 min (0.3242) was

301 still far lower than that of C50/Cc30/L15 (1.7934), the increase from 2 min for each of the REF

302 shows that, for early ages, its hydration process is faster than that of the LC3 system, which

303 corresponds well with the results of SYS and thixotropy index. This is due to the relative

304 amounts of clinker in each sample and the fact that calcined clay is not very reactive in the early

305 stages of cement hydration. It has been reported that LC3 blends usually have a low 1- or 3-day

306 compressive strength than that of REF in several studies [5,38] – this is also mirrored in the

307 lower effect of resting time (and hence the hydration mechanism) on the plastic viscosity for

308 high clinker substitutions. This reflects that the increasing effect of calcined clay on static and

309 dynamic yield stress, as well as plastic viscosity, does not primarily hinge on their hydration

310 effect but instead on their high-water adsorption, or possible interaction between the clays and

311 hydration products like ettringite [13,40,41].

21
312

313

314 Fig. 7. Effect of resting time on (a) dynamic yield stress and b) plastic viscosity, as determined

315 from the flow curves.

22
316 Table 4. The discrepancy between SYS and DYS and the thixotropic index.

Sample Resting Static yield Dynamic yield Thixotropy index

time stress (Pa) stress (Pa) [(SYS-DYS)/DYS]

REF 2 min 27.0 21.6 0.25

30 min 181.6 21.6 7.41

60 min 559.4 35.1 14.94

C80/Cc10/L5 2 min 102.1 53.1 0.92

30 min 255.8 59.2 3.32

60 min 443.7 75.8 4.85

C65/Cc20/L10 2 min 191.8 108.4 0.77

30 min 296.8 120.7 1.46

60 min 893.5 138.8 5.44

C50/Cc30/L15 2 min 448.4 222.8 1.01

30 min 772.3 258.1 1.99

60 min 1120 274.5 3.08

317

318 3.3 Adhesive properties

319 The tack test was used to measure the adhesive properties of the samples (Fig. 8a).

320 Adhesive properties are a function of material cohesion and adherence to the top plate [42]. By

321 definition, cohesion is the property of molecules of the same substance to stick to each other due

322 to mutual attraction and adherence refers to the ability of the paste to cling/stick to the upper

323 plate. This leads to two potential failure mechanisms: separation within the material vs.

23
324 separation at the plate surface. Here, separation within the material was primarily observed so the

325 discussion is focused on cohesion.

326 The results showed that a combined increase of calcined clay and limestone leads to an

327 increase in cohesion as shown by the peak force reached. A clear trend exists in the results –

328 C50/Cc30/L15 had the highest peak force and REF had the lowest. The adhesive properties

329 become more pronounced with increase in SCM content. In [14] a second peak at low plate

330 velocities was observed after about 10 s of running the test and it was attributed to a high rate of

331 structural rebuilding. However, in this study, this was only observed for C50/Cc30/L15. Apart

332 from the difference in rheometers and the smaller diameter of the top plate used in this study, the

333 differences in protocols may also have had a contribution. A protocol with a small acceleration

334 was employed in this study while in [42] variations of fixed velocities were used.

24
335

336

25
337 Fig. 8. (a) The variation of normal force during the separation of the plates with increase in SCM

338 content and (b) comparison of the effect of limestone and calcined clay on the normal force

339 during the separation of the plates at 80 wt.% substitution.

340 The independent effects of limestone and calcined clay on cohesion were analyzed by

341 comparing C80/Cc15/L0, C80/Cc0/L15 and C80/Cc10/L5 (Fig. 8b). Comparison shows that

342 C80/Cc10/L5 provides an almost perfect ‘average’ to the cohesive properties of C80/Cc0/L15

343 and C80/Cc15/L0. It can be seen from comparison of Fig. 8a and b that increase in calcined clay

344 content significantly increased cohesion while LS led to a decrease. This agrees well with the

345 results of SYS.

346 Since limestone was slightly finer than calcined clay (Fig. 1) this could have been

347 expected to contribute to properties like adhesion yet this was not the case. With as little as 10%

348 calcined clay in C80/Cc10/L5, the peak force doubles in comparison to that of REF, indicating a

349 significant increase in cohesion. This was attributed to the increase in viscosity caused by the

350 calcined clay, which originates from flocculation [43] and is also related to the high water

351 demand of calcined clay due to its layered particle structure (Fig. 2) and higher specific surface

352 area (Table 1).

353 3.4 Non-linear viscoelastic properties

354 3.4.1 Oscillatory yield stress

355 Further insight into the static yield stress of the samples is offered from part of the large

356 amplitude strain sweep results. Fig. 9 shows the minor linear viscoelastic (LVE) region and

357 transition zone for all the samples. The protocol employed here only had a minor LVE region.

358 Indeed, Fig. 9 shows an increase in the elastic modulus with an increase in SCM content, thus the

26
359 elastic portion in LC3 increases and, as established in earlier studies [27,38], the elastic modulus

360 for all the samples has a sharp drop beyond a certain critical strain which signifies transition

361 from linear to nonlinear response. According to Ref. [44], the linear region can be defined as the

362 point beyond which the measured value of complex modulus decreases to 95% of its maximum

363 value. This defining point is also referred to as the critical strain (Fig. 9). It is apparent that the

364 critical strain did not change between the mixes. The similarity in critical strain can be attributed

365 to similar hydration mechanisms between all the systems, i.e. early formation of calcium silicate

366 hydrate.

367 Irrespective of the fact that the plots in Fig. 10a and b are based on the relationship

368 between the elastic modulus and the amplitude of shear stress the trend of the oscillatory static

369 yield stress (OSYS) is similar to that of the static yield stress (SYS) as measured by the

370 logarithmic stress ramp (section 3.1). It is considered that the sample underwent permanent

371 deformation (the rigid structure was irreversibly broken) where the function G’ = f (σ) has a

372 point of inflection. Thus, the OSYS can be easily obtained for each paste if the necessary torque

373 is not greater than the maximum torque of the rheometer [15]. The results presented here vary

374 slightly from Ref. [15] in that Fig. 10a shows a valley beyond the point of deflection, which was

375 tied to oscillatory dynamic yield stress (ODYS). A summary of the oscillatory yield stresses

376 obtained is presented in Table 5. Similar to the results of SYS and DYS from before, with

377 increasing SCM content there is an increase in OSYS and ODYS.

27
378 Fig. 9.

379 Linear elastic region, part of the nonlinear transition zone and the critical strain for all the

380 samples.

28
381

382

383 Fig. 10. a) The variation of oscillatory yield stress with increase in SCM content and b)

384 Comparison of the effect of LS and CC on oscillatory yield stress at 80 wt.% substitution.

29
385 The trend in Fig. 10a is supported by that in Fig. 10b, which shows the independent effect

386 of LS and CC on the amplitude of shear stress and in turn the OSYS. The results from this

387 protocol do not show any significant difference in oscillatory yield stress resulting from the CC,

388 i.e. C80/Cc10/L5 vs C80/Cc15/L0. Recalling the results of SYS (Fig. 6) C80/Cc15/L0 exhibited

389 higher SYS at 2 min but then C80/Cc10/L5 went on to exhibit higher SYS over time up to 60

390 min. As aforementioned, the aging effect on yield stress may be attributed to the dissolution-

391 precipitation processes by LS, which cannot be precisely resolved here since the oscillatory

392 protocol employed took only 20 min. On the other hand, the increase in flowability and decrease

393 in yield stress due to the addition of LS was very apparent.

394 Table 5. The difference in OSYS and ODYS as obtained from the LAOS (strain) protocol.

Sample OSYS (Pa) ODYS (Pa) Thixotropy index

[(SYS-DYS)/DYS]

REF 13.9 11.0 0.26

C80/Cc10/L5 26.1 23.4 0.12

C65/Cc20/L10 56.5 46.8 0.21

C50/Cc30/L15 84.6 60.5 0.40

395

396 3.4.2 Harmonic distortion

397 As explained in section 2.3.4, harmonic distortion (HD) is a measure of the non-linear

398 response of the specimen. Since a large amplitude oscillation strain (LAOS) protocol was used,

399 the HD for all samples were mostly greater than 1%, implying that the output wave is not

30
400 perfectly sinusoidal and that there was apparent nonlinear response. The peak HD for all the

401 samples occurred between 60 to 80% strain (Fig. 11 (a)). Such is the case for REF. On the other

402 hand, LC3-50 displayed far less harmonic distortion, which corresponds to viscoelastic semi-

403 solid behavior.

404 The shape of the HD plots is linked to the evolution of the elastic and viscous modulus

405 and their corresponding slope (Fig. 12). It can be seen that there are generally four segments that

406 appear in Fig. 12 (and Fig. 11). Segment 1: Fast HD acceleration which corresponds to a huge

407 drop in elastic and viscous modulus, up to about 10% strain; Segment 2: Slow HD acceleration

408 which corresponds to the increasingly small difference between the elastic and viscous modulus,

409 up to about 60%; Segment 3: Peak HD, the region in which the viscous modulus becomes

410 dominant, between 60 to 80%; Segment 4: HD deceleration, up to 100% strain, viscous modulus

411 increasingly becomes larger than elastic modulus. The significance of the viscoelastic moduli

412 even in non-linear response is here illustrated despite the fact that in non-linear response the

413 viscoelastic moduli may not be defined uniquely [18,19]. It is especially interesting to note that

414 after the peak HD is reached and the viscous (loss) modulus becomes dominant, the HD begins

415 to decelerate. As stated earlier this change signifies a transition from a viscoelastic semi-solid to

416 a viscous liquid [15,46], and a breakdown of the microstructure of the paste.

417 Fig. 11b shows that limestone increases HD contrary to the effect of calcined clay.

418 C80/Cc10/L5 and C80/Cc15/L0 both had their peak distortion occurring at about 63% strain,

419 unlike C80/Cc0/L15 which continues to experience an increase in distortion even at 100% strain.

420 The HD behavior of C65/Cc20/L10 was similar to that of C80/Cc15/L0, as both showed an

421 unexpected acceleration in HD between 20 and 30% shear strain, this phenomenon may require

422 further research to be adequately explained. Clearly, limestone largely contributed to viscous

31
423 liquid behavior while calcined clay was responsible for the elastic solid behavior. In other words,

424 calcined clay increased resistance to shear deformation and hence, a relatively small distortion of

425 the sinusoidal input occurred in the output. The positive contribution of calcined clay on shape

426 stability was evident from these results. This property is very much desirable for applications

427 such as in 3-D printing [28,45].

428

32
429

430 Fig. 11. Harmonic distortion variations for a) different SCM content and b) comparison of the

431 effect of limestone and calcined clay, 80 wt.% substitution.

432

33
433 Fig. 12. Segments of harmonic distortion graphs linked with the evolution of shear modulus.

434 4. Conclusions

435 In this study, the rheology of LC3 was studied through the logarithmic stress ramp, three step

436 up-down flow analysis, the tack test and nonlinear harmonic distortion. The main conclusions

437 that can be drawn from this work are as follows:

438 • The rheology of LC3 paste differs widely from that of OPC paste. With an increase in

439 SCM content, there was an increase in plastic viscosity, static and dynamic yield stress,

440 cohesion and adhesion and a decrease in harmonic distortion.

441 • In all systems, increase in resting time led to increase in viscosity, dynamic yield stress

442 and static yield stress. REF exhibited higher plastic viscosity and dynamic yield stress

443 percentage increase from 2 minutes to 60 minutes, which is related to its faster hydration

444 mechanism owing to higher clinker content, whereas the LC3 pastes exhibited some

445 dilution effect.

446 • Through an analysis of pastes with only calcined clay or limestone, calcined clay is

447 understood to be the main contributing factor for increased plastic viscosity, static and

448 dynamic yield stress, initial thixotropy index, and cohesion, as well as reduced harmonic

449 distortion. Whereas limestone, despite its smaller proportion, slightly reduces these

450 parameters and therefore can help tailor the workability of LC3. The effect of calcined

451 clay emanates from its intrinsic properties such as high specific surface area and layered

452 particle structure which primarily lead to a high water demand and greater flocculation.

453 • Harmonic distortion is closely tied to the evolution of the viscous and elastic modulus

454 and it offers a relatively simple and effective way to analyze the somewhat complex large

455 amplitude oscillation strain protocol and the corresponding nonlinear behavior. It is

34
456 believed that this parameter opens an interesting pathway for further study on the

457 nonlinear response of fresh cement-based systems, while the protocol allows for the

458 characterization of both the linear and the nonlinear response at once.

459

460 Acknowledgement

461 The authors gratefully acknowledge the financial support from National Key R&D Plan

462 (2016YFE0206100), Natural Science Foundation of China (Grants No.51672107, and

463 51850410509). Support by Taishan Scholars Program, Case-by-Case Project for Top

464 Outstanding Talents of Jinan and the 111 Project of International Corporation on Advanced

465 Cement-based Materials (No. D17001) is also acknowledged. We would like to also thank the

466 anonymous reviewers for comments that helped to improve the quality of the manuscript.

467
468
469 References
470 [1] M.C.G. Juenger, F. Winnefeld, J.L. Provis, J. H. Ideker, Advances in alternative cementitious
471 binders, Cem. Concr. Res. 41 (2011) 1232-1243. https://doi.org/10.1016/j.cemconres.2010.11.012.
472 [2] M. Schneider, M. Romer, M. Tschudin, H. Bolio, Sustainable cement production present and
473 future, Cem. Concr. Res. 41 (2011) 642–650.
474 http://dx.doi.org/10.1016/j.cemconres.2011.03.019.
475 [3] WBCSD, Cement Sustainability Initiative, Getting the numbers right, http://www.
476 wbcsdcement.org/index.php/key-issues/climate-protection/gnr-database (accessed 29 April
477 2019).
478 [4] R. Snellings, Assessing, understanding and unlocking supplementary cementitious materials.
479 RILEM Technical Letters, [S.l.], v. 1, Aug. 2016, pp. 50–55 ISSN 2518-0231. Available at:
480 https://letters.rilem.net/index.php/rilem/article/view/12 (accessed 11 March 2019, doi:
481 https://doi.org/10.21809/rilemtechlett. 2016.12).
482 [5] K. Scrivener, M. Fernando, B. Shashank, M. Soumen, Calcined clay limestone cements
483 (LC3), Cem. Conr. Res. 114 (2017) 49-56. https://doi.org/10.1016/j.cemconres.2017.08.017.

484 [6] C. He, B. Osbaeck, E. Makovicky, Pozzolanic reactions of six principal clay minerals:
485 activation, reactivity assessments and technological effects, Cem. Conr. Res. 25 (8) (1995)
486 1691–1702. https://doi.org/10.1016/0008-8846(95)00165-4.
487 [7] A. Alujas, R. Fernández, R. Quintana, K.L. Scrivener, F. Martirena, Pozzolanic reactivity of

35
488 low grade kaolinitic clays: influence of calcination temperature and impact of calcination
489 products on OPC hydration, Appl. Clay Sci. 108 (2015) 94–101.
490 https://doi.org/10.1016/j.clay.2015.01.028.
491 [8] F. Avet, R. Snellings, A. Alujas Diaz, M. Ben Haha, K. Scrivener, Development of a new
492 rapid, relevant and reliable (R3) test method to evaluate the pozzolanic reactivity of calcined
493 kaolinitic clays, Cem. Concr. Res. 85 (2016) 1–11.
494 https://doi.org/10.1016/j.cemconres.2016.02.015.
495 [9] F. Avet, K. L. Scrivener, Investigation of the calcined kaolinite content on the hydration of
496 Limestone Calcined Clay Cement (LC3), Cem. Concr. Res. 107 (2018) 124-135.
497 https://doi.org/10.1016/j.cemconres.2018.02.016.
498[10] ACI Committee 238, Workability of Fresh Concrete, Role of rheology in achieving
499 successful concrete performance, American Concrete Institute, Farmington Hills, MI, 2017.
500 https://academiccommons.columbia.edu/doi/10.7916/D8GQ86X7/download (accessed 29
501 April 2019).
502[11] P.F.G. Banfill, Rheology of fresh cement and concrete, Rheol. Rev. (2006) 61 - 130.
503 http://www.schleibinger.com/k2003/banfill/banfill.pdf (accessed 29 April 2019).
504[12] R.J. Flatt, Towards a prediction of superplasticized concrete rheology, Mater. Struct. 37
505 (2004) 289-300. http://dx.doi.org/10.1007/BF02481674.
506[13] S. Kawashima, M. Chaouche, D.J. Corr, S.P. Shah, Rate of thixotropic rebuilding of cement
507 pastes modified with highly purified attapulgite clays, Cem. Concr. Res. 53 (2013) 112–118.
508 https://doi.org/10.1016/j.cemconres.2013.05.019.
509[14] S. Kawashima, M. Chaouche, D.J. Corr, S.P. Shah, Influence of purified attapulgite clays on
510 the adhesive properties of cement pastes as measured by the tack test, Cem. Concr. Compos.
511 48 (2014) 35–41. https://doi.org/10.1016/j.cemconcomp.2014.01.005.
512[15] F.N. Santos, S.R. Gomes de Sousa, A.J.F. Bombard, S.L. Vieira, Rheological study of
513 cement paste with metakaolin and/or limestone filler using Mixture Design of Experiments,
514 Constr. Build. Mater. 143 (2017) 92–103. https://doi.org/10.1016/j.conbuildmat.2017.03.001.
515[16] N.A. Tregger, M.E. Pakula, S.P. Shah, Influence of clays on the rheology of cement pastes,
516 Cem. Concr. Res. 40 (2010) 384–391. http://dx.doi.org/10.1016/j.cemconres.2009.11.001.
517[17] S. Ferreiro, D. Herfort, J.S. Damtoft, Effect of raw clay type, fineness, water-to-cement ratio
518 and fly ash addition on workability and strength performance of calcined clay – Limestone
519 Portland cements, Cem. Concr. Res. 101 (2017) 1–12.
520 https://doi.org/10.1016/j.cemconres.2017.08.003.
521[18] K. Hyun, M. Wilhelm, C.O. Klein, K.S. Cho, J.G. Nam, K.H. Ahn, S.J. Lee, R.H. Ewoldt,
522 G.H. McKinley, A review of nonlinear oscillatory shear tests: Analysis and application
523 of large amplitude oscillatory shear (LAOS), Prog. Polym. Sci. 36 (2011) 1697–1753.
524 https://doi.org/10.1016/j.progpolymsci.2011.02.002.
525[19] T. Conte, M. Chaouche, Rheological behavior of cement pastes under Large Amplitude
526 Oscillatory Shear, Cem. Concr. Res. 89 (2016) 332–344.
527 https://doi.org/10.1016/j.cemconres.2016.07.014.

36
528[20] Y. Qian, S. Ma, S. Kawashima, G. De Schutter, Rheological characterization of the
529 viscoelastic solid-like properties of fresh cement pastes with nanoclay addition, Theoretical
530 and Applied Fracture Mechanics 103 (2019) 102262.
531 https://doi.org/10.1016/j.tafmec.2019.102262
532[21] M. Choi, K. Park, T. Oh, Viscoelastic properties of fresh cement paste to study the flow
533 behavior, Inter. J. Concr. Struct. Mater. 10 (3) (2016) S65–S74.
534 http://dx.doi.org/10.1007/s40069-016-0158-3.
535[22] Q. Yuan, D. Zhou, K.H. Khayat, D. Feys, C. Shi, On the measurement of evolution of
536 structural build-up of cement paste with time by static yield stress test vs. small amplitude
537 oscillatory shear test, Cem. Concr. Res. 99 (2017) 183–189.
538 http://dx.doi.org/10.1016/j.cemconres.2017.05.014
539[23] N. Roussel, G. Ovarlez, S. Garrault, C. Brumaud, The origins of thixotropy of fresh cement
540 pastes, Cem. Concr. Res. 42(1) (2012) 148–157.
541 https://doi.org/10.1016/j.cemconres.2011.09.004
542[24] A. Dutta, K. Koley, S.K. Saha, C.K. Sarkar, Impact of temperature on linearity and harmonic
543 distortion characteristics of underlapped FinFET, Microelectron. Reliab. 61 (2016) 99–105,
544 http://dx.doi.org/10.1016/j.microrel.2016.01.017.
545[25] G. McLorn, D. Laverty, D.J. Morrow, S. McLoone, Load and harmonic distortion
546 characterization of modern low-energy lighting under applied voltage variation, Electric
547 Power Syst. Res. 169 (2019) 124–138. https://doi.org/10.1016/j.epsr.2018.12.029.
548[26] Y, Qian, S. Kawashima, Distinguishing dynamic and static yield stress of fresh cement
549 mortars through thixotropy, Cem. Concr. Compos. 86 (2018) 288-296.
550 https://doi.org/10.1016/j.cemconcomp.2017.11.019.
551[27] Y. Qian, S. Kawashima, Use of creep recovery protocol to measure static yield stress and
552 structural rebuilding of fresh cement pastes, Cem. Concr. Res. 90 (2016) 73–79.
553 https://doi.org/10.1016/j.cemconres.2016.09.005.
554[28] V.N. Nerella, M.A.B. Beigh, S. Fataei, V. Mechtcherine, Strain-based approach for
555 measuring structural build-up of cement pastes in the context of digital construction, Cem.
556 Concr. Res. https://doi.org/10.1016/j.cemconres.2018.08.003.
557[29] S. Ng, H. Justnes, Rheology of blended cements with superplasticizers, Concrete Innovation
558 Centre (COIN) Project report 58 – 2015.
559 https://brage.bibsys.no/xmlui/bitstream/handle/11250/2378670/COIN_report_no_58.pdf?seq
560 uence=3 (accessed 29 April 2019).
561[30] W. Mbasha, I. Masalova, R. Haldenwang, A. Malkin, The yield stress of cement pastes as
562 obtained by different rheological approaches, Appl. Rheol. 25 (2015) 53517.
563 https://doi.org/10.3933/ApplRheol-25-53517.
564[31] J. Peng, D. Deng, Z. Liu, Q. Yuan, T. Ye, Rheological models for fresh cement asphalt paste,
565 Constr. Build. Mater. 71 (2014) 254–262. https://doi.org/10.1016/j.conbuildmat.2014.08.031.
566[32] M. Nehdi, M.A. Rahman, Estimating rheological properties of cement pastes using various
567 rheological models for different test geometry, gap and surface friction, Cem. Concr. Res. 34

37
568 (2004) 1993–2007. https://doi.org/10.1016/j.cemconres.2004.02.020.
569[33] O.H. Wallevik, D. Feys, J.E. Wallevik, K.H. Khayat, Avoiding inaccurate interpretations of
570 rheological measurements for cement-based materials, Cem. Concr. Res. 78 (2015) 100–109.
571 https://doi.org/10.1016/j.cemconres.2015.05.003.
572[34] Y. Zhang, X. Kong, L. Gao, Z. Lu, S. Zhou, B. Dong, F. Xing, In-situ measurement of
573 viscoelastic properties of fresh cement paste by a microrheology analyzer, Cem. Concr. Res.
574 79 (2016) 291–300. https://doi.org/10.1016/j.cemconres.2015.09.020.
575[35] J. Läuger, H. Stettin, Differences between stress and strain control in the non-linear behavior
576 of complex fluids, Rheol. Acta 49 (2010) 909–930, http://dx.doi.org/10.1007/s00397-010-
577 0450-0.
578[36] R.H. Ewoldt, A.E. Hosoi, G.H. McKinley, New measures for characterizing nonlinear
579 viscoelasticity in large amplitude oscillatory shear, J Rheol 52(6) (2008) 1427–1458.
580 http://dx.doi.org/10.1122/1.2970095.
581[37] K. Vance, A. Kumar, G. Sant, N. Neithalath, The rheological properties of ternary binders
582 containing Portland cement, limestone, and metakaolin or fly ash, Cem. Concr. Res. 52 (2013)
583 196–207. https://doi.org/10.1016/j.cemconres.2013.07.007.
584[38] M. Antoni, J. Rossen, F. Martirena, K. Scrivener, Cement substitution by a combination of
585 metakaolin and limestone, Cem. Concr. Res. 42 (2012) 1579–1589.
586 http://dx.doi.org/10.1016/j.cemconres.2012.09.006.
587[39] M. Fourmentin, G. Ovarlez, P. Faure, U. Peter, D. Lesueur, D. Daviller, P. Coussot,
588 Rheology of lime paste - A comparison with cement paste, Rheol. Acta 54 (2015) 647-656.
589 http://dx.doi.org/10.1007/s00397-015-0858-7.
590[40] A.J. McFarlane, J. Addai-Mensah, K. Bremmell, Rheology of flocculated kaolinite
591 dispersions, Korea-Australia Rheol. J. 17 (4) (2005) 181-190.
592 https://www.cheric.org/PDF/KARJ/KR17/KR17-4-0181.pdf (accessed 29 April 2019).
593[41] M.S. Mansour, R. Chaid, A. Afalfiz, K. Bekkour, Rheology of high-performance cement
594 pastes: effect of calcined kaolin, Chemistry and Materials Research, 4 (2013) (Special Issue
595 for International Congress on Materials & Structural Stability, Rabat, Morocco, 27-30
596 November 2013). https://www.iiste.org/Journals/index.php/CMR/article/view/9990 (accessed
597 29 April 2019).
598[42] A. Kaci, R. Bouras, V.T. Phan, P.A. Andréani, M. Chaouche, H. Brossas, Adhesive and
599 rheological properties of fresh fibre-reinforced mortars, Cem. Concr. Compos. 33(2) (2011)
600 218–24. https://doi.org/10.1016/j.cemconcomp.2010.10.009.
601[43] Mohamed Abdelhaye YO, Chaouche M, Van Damme H. The tackiness of smectite muds. 1.
602 The dilute regime, Appl. Clay Sci. 42(1–2) (2008) 163–7.
603 https://doi.org/10.1016/j.clay.2008.01.018.
604[44] B. Rahimzadeh, Linear and non-linear viscoelastic behaviour of binders and asphalts, PhD
605 dissertation, University of Nottingham, 2002.
606 https://www.nottingham.ac.uk/research/groups/ntec/documents/theses/rahimzadehbphdthesis.
607 pdf (accessed 29 April 2019).

38
608[45] A. Perrot, D. Rangeard, A. Pierre, Structural built-up of cement-based materials used for 3D-
609 printing extrusion techniques, Mater. Struct. (2015) 1213–1220,
610 https://doi.org/10.1617/s11527-015-0571-0.

39
Declaration of interests

☒ The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in
this paper.

☐The authors declare the following financial interests/personal relationships


which may be considered as potential competing interests:

You might also like