Sie sind auf Seite 1von 31

Accepted Manuscript

Synthesis and characterization of one-part geopolymers for extrusion based 3D


concrete printing

Biranchi Panda, GVP Bhagath Singh, Cise Unluer, Ming Jen Tan

PII: S0959-6526(19)30584-0
DOI: https://doi.org/10.1016/j.jclepro.2019.02.185
Reference: JCLP 15915

To appear in: Journal of Cleaner Production

Received Date: 19 September 2018


Revised Date: 3 February 2019
Accepted Date: 17 February 2019

Please cite this article as: Panda B, Singh GB, Unluer C, Tan MJ, Synthesis and characterization of one-
part geopolymers for extrusion based 3D concrete printing, Journal of Cleaner Production (2019), doi:
https://doi.org/10.1016/j.jclepro.2019.02.185.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT

1 Word count: 5452


2
3 Synthesis and characterization of one-part geopolymers for extrusion based 3D concrete
4 printing
5

PT
6 Biranchi Panda1, GVP Bhagath Singh2,3, Cise Unluer3*, Ming Jen Tan1
7

RI
1
8 Singapore Centre for 3D Printing, School of Mechanical & Aerospace Engineering
9 Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798

SC
2
10 Laboratory of Construction Materials, IMX, EPFL, 1015 Lausanne, Switzerland
3
11 School of Civil and Environmental Engineering, Nanyang Technological University, 50

U
12 Nanyang Avenue, Singapore 639798
13
AN
14 * Corresponding author. Tel.: +65 91964970, E-mail address: ucise@ntu.edu.sg
15
M

16 Abstract:
17
D

18 Interest in innovative construction processes such as 3D concrete printing (i.e. digital


19 construction), is growing rapidly both in academia and industry. Processing conventional
TE

20 geopolymer mixes, in which alkaline solutions are used for activation, could be troublesome in
21 concrete printing due to the high viscosity of the alkaline solution. One-part geopolymers offer
EP

22 one possible solution to this challenge as they involve the use of a solid activator with solid
23 aluminosilicates precursors. In this work, a printable one-part geopolymer mix was developed,
C

24 which could be extruded through the nozzle of a 3D printer and stacked together without
25 deforming the bottom layers. Flow properties such as yield stress, viscosity and thixotropy of the
AC

26 developed geopolymer were assessed along with its strength development curve. Printed
27 specimen showed anisotropic behaviour in mechanical properties when compared to the mould
28 casted samples. Microstructural characterization revealed the formation of alumino-silicate gel
29 with high tetrahedrally coordinated Al and interlayer K ions in its structure. When compared to
30 OPC-based mixes, the developed geopolymer mixes revealed a lower environmental impact,
31 which could be even further reduced with the use of alternative activators.
ACCEPTED MANUSCRIPT

32
33 Keywords: Digital construction; rheology; one-part geopolymer; mechanical properties;
34 environmental assessment

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

35 1. Introduction
36
37 The increasing demand for environmentally friendly and sustainable construction materials has
38 necessitated the identification of alternative materials for ordinary Portland cement (OPC) (Garg
39 et al., 2018, 2019; Hu et al. 2018, 2019). In this regard, geopolymer cement, involving the use of

PT
40 various industrial wastes and by-products, has the potential to be considered as a promising
41 alternative to OPC in certain applications. The term “geopolymer” was first introduced in the

RI
42 literature in 1978, characterizing a new class of materials with the ability to poly-condense at low
43 temperatures like “polymers”. This process involves the chemical reaction of aluminosilicate

SC
44 materials (e.g. fly ash, metakaolin, granulated blast furnace slag and silica fume) with Na- or K-
45 based alkali activators. When mixed with the alkaline activators, setting and hardening take

U
46 place, yielding a material with good binding properties. Several advantages of geopolymer
47 materials over OPC, namely potential environmental benefits, high compressive strength, rapid
AN
48 controllable setting and hardening, fire resistance, and acid and salt solution resistance have
49 already been reported (Deventer et al., 2007; Hu et al, 2018; Nie et al., 2016). One of the most
M

50 important benefits of geopolymers lies in the utilization of industrial wastes as raw materials. In
51 terms of their environmental impact, geopolymers are reported to generate nearly 80% less CO2
D

52 than OPC (Duxson and Van Deventer, 2009). In line with these advantages, an aircraft pavement
53 was constructed at the Wellcamp Airport in Australia by using geopolymer concrete in 2014
TE

54 (Glasby et al., 2015). Despite the remarkable potential of geopolymers, the conventional two-part
55 mixing process presents challenges in their scalability due to handling issues and viscosity of the
EP

56 alkaline solution. Therefore, one-part geopolymers, which are mixtures of a solid alkaline
57 activator and an aluminosilicate precursor, can be used to solve some of these shortcomings
C

58 (Luukkonen et al., 2018; Adesanya et al., 2018).


59
AC

60 Different synthesis routes in designing one-part geopolymer cements, either by activating


61 aluminosilicates together with alkalis at elevated temperatures or blending the aluminosilicates
62 with solid activators, have been reported (Yang et al., 2016; Choo et al., 2016). A recent review
63 of these techniques (Luukkonen et al., 2018) provided a summary of the potential benefits,
64 environmental impacts and cost of one-part systems. Nevertheless, the use of one-part
65 geopolymers in concrete printing/digital construction, which is gaining significant attention both
ACCEPTED MANUSCRIPT

66 in academia and the construction practice (Buswell et al., 2018; Tay et al., 2018; Wangler et al.,
67 2016), has not been discussed in detail until now. Most of the work presented in this area so far
68 has focused on exploring the material properties required for 3D printing involving the use of
69 OPC-based binders, with and without the use of fibers (Le et al., 2012; Kazemian et al., 2017;
70 Panda et al., 2018). Within this context, parameters such as extrudability, shape retention, open

PT
71 time and buildability properties, collectively described as “printability” that refers to the layer-
72 by-layer deposition of cementitious mortars according to a CAD model (Panda et al., 2017), need

RI
73 to be studied further.
74

SC
75 Building materials with a high yield strength and low viscosity are usually preferred in concrete
76 printing, where the yield stress is linked to the shape retention and buildability characteristics

U
77 and viscosity is an indication of the ease of mortar extrusion. The main challenge exists in
78 controlling the structuration rate of the material without disturbing its flow properties so that the
AN
79 well extruded layers can be deposited in a layer-by-layer manner. Due to the thixotropic nature
80 of OPC particles (Roussel, 2018), it is possible to achieve printable properties, during which the
M

81 material can flow smoothly via the applied shear force and can incrementally gain yield stress by
82 particle interaction/flocculation at rest. Although the use of two-part systems has been studied
D

83 previously (Xia et al., 2016; Panda et al., 2018), no detailed studies on the use of one-part
84 geopolymers in 3D printing have been reported until now. Therefore, this paper aims to study the
TE

85 formulation of one-part geopolymers and provide an assessment of their fresh properties (e.g.
86 yield stress, viscosity and thixotropy) to evaluate their suitability in non-structural concrete
EP

87 printing applications. Furthermore, the mechanical performance of the printed sections was
88 tested in compression after 28 days of ambient curing and their directional properties were
C

89 compared with mould casted geopolymers. The strength and microstructural development of the
90 prepared formulations was further supported by an investigation and quantification of the phase
AC

91 formations via x-ray diffraction (XRD) and field emission scanning electron microscopy
92 (FESEM) analyses. The energy and CO2 emissions associated with one-part geopolymer mixes
93 were also provided to evaluate their potential against OPC from an environmental standpoint.
94
95
96 2. Materials and Methodology
ACCEPTED MANUSCRIPT

97
98 2.1 Materials
99
100 Class F grade fly ash (FA), conforming to ASTM C 618, was obtained from Thermal Powertech
101 Corporation Ltd. (India). Ground granulated blast furnace slag (GGBS) was supplied by EnGro

PT
102 Corporation Ltd. (Singapore). Table 1 shows the chemical compositions of FA and GGBS,
103 obtained via X-ray fluorescence (XRF) analysis.

RI
104
105 Table 1. Chemical compositions (%) of FA and GGBS

SC
Material SiO2 CaO Al2O3 Fe2O3 SO3 MgO Na2O K2O TiO2 LOI
FA 51.1 5.8 28.8 6.4 1.0 1.1 0.3 1.2 2.5 1.2

U
GGBS 29.6 39.3 15.5 0.3 4.3 7.5 0.4 0.5 1.7 1.1
106
AN
107 As can be seen from Fig. 1, both FA and GGBS contained different crystalline phases along with
108 broad humps that represented the amorphous phases present in each material. FA mainly
M

109 consisted of quartz (PDF# 074-3485) and mullite (PDF# 074-4143) in crystalline form, whereas
110 GGBS displayed a broad amorphous hump at around 30° 2θ and gehlenite (PDF#0 04-0690),
D

111 which was found in small amounts.


112
TE
C EP
AC

113
ACCEPTED MANUSCRIPT

114 Fig. 1 X-ray diffractograms of FA and GGBS


115
116 The amorphous/glassy contents of the FA and GGBS were determined using an external standard
117 approach, as outlined by Singh and Subramaniam in 2016 (Singh and Subramaniam, 2016).
118 Accordingly, the total glassy content present in FA and GGBS were determined by subtracting

PT
119 the mass fractions of the crystalline phases from the total, while the individual reactive contents
120 present in FA was determined via a combination of XRF and XRD methods (Fernández-Jiménez

RI
121 et al., 2006; Singh and Subramaniam, 2016). The individual reactive contents present in the
122 glassy phases were determined by subtracting the crystalline phases from the total oxide

SC
123 contents. The total oxide contents were determined using XRF, whereas the crystalline phases
124 were determined by using Rietveld refinement. The obtained individual reactive contents present

U
125 in FA and total glassy content of FA and GGBS are listed in Table 2. Results revealed that FA
126 contained a low amount of reactive phases, which could be due to the usage of low-grade coal
AN
127 and processing conditions of FA (Singh and Subramaniam, 2018).
128
M

129 Table 2. Reactive components of FA and GGBS


130
Crystalline content (%)
D

131 Material Amorphous content (%)


Al2O3 SiO2 Fe2O3 CaO K2O TiO2
132
TE

FA 10.72 14.28 5.94 2.75 2.17 1.47 38.8


133
GGBS 2.8 97.2
134
EP

135
136 Other than FA and GGBS, the solid alkaline activator used in this study was prepared by
C

137 blending potassium silicate powder (49.50 wt.% SiO2, 42.54 wt.% K2O and remaining H2O)
138 together with potassium hydroxide (KOH) powder to reach the desired modulus (MR) of 1.5
AC

139 (Davidovits, 2008). For the preparation of geopolymer mortars, fine river sand with a maximum
140 particle size of 2 mm was used along with tap water.
141
142
143 2.2 Sample preparation
144
ACCEPTED MANUSCRIPT

145 Five geopolymer mortars with a constant water/solid (i.e. solid component consisting of FA,
146 GGBS and anhydrous activator) ratio of 0.35 were prepared by varying the GGBS and activator
147 contents, as shown in Table 3. Accordingly, the GGBS (G) ranged between 15-40% of the
148 overall binder (FA+GGBS) content within samples F85G15A15, F70G30A15 and F60G40A15,
149 in which the activator (A) content was kept constant at 15% (i.e. by mass) of the binder.

PT
150 Furthermore, the FA and GGBS contents remained constant, whereas the activator dosage varied
151 at 10, 15 and 20% in samples F70G30A10, F70G30A15 and F70G30A20, respectively. For the

RI
152 preparation of mortar samples, fine river sand was added to the binder at a sand/binder ratio of
153 0.85, along with the activator and mixed for two minutes in a Hobart planetary mixer. This was

SC
154 followed with the addition of the required amount of water into the mixing bowl, after which
155 mixing continued until a homogeneous mixture was obtained.

U
156
157 Table 3. Mix proportions of one-part geopolymer mixes prepared in this study
AN
Binder composition (wt.%)
Mix name
FA GGBS Activator content (%)
M

F85G15A15 85 15
F70G30A15 70 30
15
D

F60G40A15 60 40
F70G30A10 70 30 10
TE

F70G30A20 70 30 20
158
EP

159
160 2.3 Methodology
C

161
162 2.3.1 Process overview of 3D concrete printing
AC

163
164 3D concrete printing is similar to other 3D printing processes, where the model is first designed
165 in CAD platform, followed by slicing and G-code generation (Tay et al., 2017). The G-code is
166 later interpreted by a controller that sends commands to either an industrial robot or gantry
167 printer, connected to a concrete pump and extruder/nozzle. Fig. 2 shows the steps involved in a
ACCEPTED MANUSCRIPT

168 typical concrete printing process, during which the fresh material is deposited layer-by-layer via
169 a 4-axis industrial robot.
170

PT
RI
171
172 Fig. 2 Steps involved in the 3D concrete printing process

SC
173
174 During printing, process parameters (e.g. printing speed and layer height), material properties
(e.g. rheology and thixotropy) as well as the design parameters (e.g. shape and size) play an

U
175
176 important role in determining the quality of the printed filament (Tay et al., 2019; Panda et al.,
AN
177 2018; Bos et al., 2016). These factors are depicted in Fig. 3, which outlines the main parameters
178 controlling the properties of 3D printed concrete.
M

179
D
TE
C EP
AC

180
181 Fig. 3 Cause-effect diagram reflecting the main parameters controlling the properties of 3D
182 printed concrete
183
184
ACCEPTED MANUSCRIPT

185 2.3.2 Experimental methodology


186
187 2.3.2.1 Flow properties
188
189 To obtain the evolution of static yield stress, stress growth test was performed by applying

PT
190 deformation at a constant shear rate of 0.1 s-1. During this test, the shear stress progressively
191 developed to a maximum value and then stabilized at an equilibrium value. The static yield stress

RI
192 was defined as the peak shear stress value (Panda et al., 2018).
193

SC
194
195 2.3.2.2 Thixotropy and viscosity recovery

U
196
197 In this study, thixotropy was measured using a structural parameter (λ) (Ouyang et al., 2016), as
AN
198 shown in Equation 1, where and are the equilibrium stress and initial shear stress,
199 respectively. The λ value indicates the relationship between static and dynamic yield stress,
M

200 which is critical in 3D concrete printing applications (i.e. the higher the λ, the higher the
201 thixotropy).
D

202

203 λ= (1)
TE

204
205 In addition to thixotropy, viscosity recovery is also an important property of printable concrete,
EP

206 which indicates fresh material behaviour after the extrusion process. If the original viscosity of
207 the initial deposited layer is not recovered before the deposition of the second layer, it may result
C

208 in the deformation of the structure. Therefore, in this study, the 3D printing process was
mimicked by applying different shear rates of (i) 0.01 s-1 for 60 seconds, (ii) 300 s-1 for 30
AC

209
210 seconds and (iii) 0.01 s-1 for 60 seconds at three different time intervals corresponding to the
211 material state (i) initially at rest, (ii) shearing/extrusion and (iii) again at rest, as used in previous
212 studies (Panda et al., 2018). The apparent viscosities were measured during these three intervals
213 to understand the recovery behaviour of the different geopolymer mixes.
214
215
ACCEPTED MANUSCRIPT

216 2.3.2.3 Compressive strength


217
218 To study the hardened properties of the prepared formulations, specimens were cast in 50x50x50
219 mm moulds and cured under ambient conditions for up to 28 days. Average compressive
220 strengths were measured by uni-axial loading in triplicates, in accordance with the specifications

PT
221 of ASTM C109(ASTM C109/C109M-13). The equipment used for this purpose was a Toni
222 Technik Baustoffprüfsysteme machine, operated at a loading rate of 100 kN/min.

RI
223
224

SC
225 2.3.2.4 3D concrete printing
226

U
227 One of the mixes that met the minimum requirement of M25 grade concrete was selected out of
228 the initially prepared five mix designs and printed using a 4-axis gantry printer fitted with a
AN
229 40x10 mm rectangular nozzle and a grout pump, and operated at a printing speed of 90 mm/sec.
230 The outcome was a solid printed block with a length of 400 mm. The printed block was then split
M

231 into 40x40 mm cubes for directional compression testing. The same mix was also cast separately
232 to compare its performance with those extracted from the 3D printed section.
D

233
234
TE

235 2.3.2.5 Structural build up (load bearing capacity)


236
EP

237 The assessment of structural build-up is important for concrete printing since the buildability
238 property depends on the structuration rate of the material (i.e. how fast the material can gain
C

239 stiffness) after extrusion. Accordingly, the early (i.e. green) strength of the material was
240 measured in the dormant period via the incremental loading of the fresh mortar using an Instron
AC

241 tensile test series machine (Voigt et al., 2006). These results were further supported with the 28-
242 day hardened properties to analyse the full-strength development of the material.
243
244
245 2.3.2.6 Microstructural characterization (XRD and FESEM)
246
ACCEPTED MANUSCRIPT

247 X-ray diffraction (XRD) and field emission electron microscopy (FESEM) were used to
248 characterize the raw materials and the reaction products of the prepared mixes. Randomly
249 oriented powder samples were extracted from the selected mixes by grinding the dried samples
250 for XRD analysis. The scan patterns were obtained from 10° to 70° 2θ with a step size of 0.02°
251 by using an Empyrean X-ray Diffractometer with Cu-Kα-radiation. The X-ray tube generator

PT
252 was operated at 40 kV and 40 mA. Corundum was used as the external standard for the
253 quantification of both crystalline and amorphous phases using TOPAS software. The crystalline

RI
254 phases were identified using the powder diffraction file-2 (Singh and Subramaniam, 2016), while
255 known phases were taken from the ICSD data base. For FESEM analysis, a JEOL JSM-7600F

SC
256 electron microscope, equipped with an energy dispersive analyser was used to examine the
257 sample microstructure after 28 days of curing.

U
258
259
AN
260 3. Results and Discussion
261
M

262 3.1 Flow and mechanical properties


263
D

264 Fig. 4(a) shows the static yield stress and apparent viscosity of the five mixes containing
265 different GGBS and activator contents. An increase in the GGBS content led to an increase in
TE

266 both the yield stress and viscosity, which could be due to the chemical composition (i.e. presence
267 of calcium) in GGBS and its angular particle shape. Angular particles can provide an
EP

268 interlocking effect, which can increase the yield stress. A similar outcome on the effect of the
269 angular morphology of GGBS on concrete workability was reported in previous studies (Deb et
C

270 al., 2014; Cheah et al., 2017). Therefore, the partial replacement of 40% of FA with GGBS
271 resulted in 125% increment in the yield stress, which can be useful in the shape fidelity and
AC

272 buildability of the printed layers. As the addition of GGBS was reported to decrease the setting
273 time of alkali-activated materials (Deb et al., 2014), it can significantly affect the workability
274 time (i.e. open time) by changing the flow properties. Therefore, the amount of GGBS included
275 in these mixes should be carefully controlled with proper activator dosage for the 3D printing of
276 one-part geopolymer mortars. Similarly, increasing the activator content from 10% to 20% in
277 mix F70G30A20 led to a higher yield stress, in line with the findings of Kashani et al. (Kashani
ACCEPTED MANUSCRIPT

278 et al., 2014) . Accordingly, a higher dosage of the activator induced a higher pH and ionic
279 strength of surface charges, which can be related to the increase in the yield stress. The
280 viscosities of geopolymer mortars, irrespective of different slag and activator contents, was
281 found to be substantially low (< 10 Pa·s) under a high shear force. Such unique property (i.e.
282 high yield stress and low viscosity) could find an application in the 3D printing process with

PT
283 some additional attention on thixotropy and structural build-up. The observed decrease in
284 viscosity with increasing shear rate can be attributed to the deflocculating phenomena in the

RI
285 early stages of the alkali-activation process, for which, further mixing can increase the
286 workability of the mixtures before its application.

SC
287
288 Similar to the flow properties, the compressive strength of the prepared samples increased with

U
289 increasing GGBS content, as can be seen in Fig. 4(b). This increase in strength could be
290 attributed to the early formation of C–S–H gel in the matrix (Collins and Sanjayan 1999; Puertas
AN
291 et al., 2000). A similar increase was observed with an increase in the activator content from 10 to
292 20%, which could be due to the presence of a higher amount of Si ions available for
M

293 geopolymerisation. This finding was in line with the findings of Nematollahi et al. (Nematollahi
294 et al., 2015), where an improvement in strength with respect to the activator (i.e. sodium silicate)
D

295 content was reported.


296
TE
C EP
AC

297
298 Fig. 4 Assessment of the prepared mixes for their (a) static yield stress and viscosity and (b)
299 compressive strength
300
ACCEPTED MANUSCRIPT

301 Besides the yield stress and viscosity, the thixotropy, referring to the reversible structural
302 breakdown and build-up aspect of printable materials, was studied. Throughout this assessment,
303 the continuous decrease of apparent viscosity with time under shear and subsequent recovery at
304 rest was observed. Fig. 5(a) shows the viscosity recovery of each mix, where their recovery
305 ability after being sheared at a shear rate of 300 s-1 was revealed. Considering the shear-thinning

PT
306 property of one-part geopolymers (Fig. 5(b)), the patterns observed in viscosity recovery can be
307 linked to the concrete printing process, indicating the properties of each mix after their extrusion

RI
308 from the nozzle. Theoretically, if a material can recover its initial viscosity after the extrusion
309 process, it can maintain its shape without any significant deformation. The findings reported in

SC
310 this study showed that the prepared mixes were able to recover 70-80% of their original viscosity
311 within 60 seconds of extrusion, thereby enabling the deposition of a second layer within the 60

U
312 seconds time interval, without deforming the bottom layers.
313
AN
M
D
TE
EP

314
315 Fig. 5 Assessment of the prepared mixes for their (a) viscosity recovery and (b) shear thinning
316 behaviours
C

317
AC

318 The structural/thixotropy parameter (λ) of the geopolymer mixes, listed in Table 4, increased
319 with an increase in the amount of GGBS, which was an indication of the strong bonding due to
320 the angular morphology of the GGBS particles. With the use of higher contents of GGBS that
321 replaced FA, the rolling nature of spherical FA diminished, increasing the particle-to-particle
322 interlocking. Increase in packing density could therefore explain the enhanced thixotropy
323 observed in geopolymer mixes with high GGBS contents. Alternatively, the increase in the
ACCEPTED MANUSCRIPT

324 activator content from 10 to 20% reduced the λ value, which could be attributed to the weaker
325 bonding between colloidal particles due to the plasticizing effect of silicates. Within these mixes,
326 silicate anions could have adsorbed on the particle surfaces, increasing the inter particle distance
327 (Kashani et al., 2014), and hence λ. Considering that the main indented use for the mixes
328 designed in this study were non-structural printed components requiring a strength of 25-30

PT
329 MPa, mix F70G30A10 was selected for demonstrative 3D printing, as well as the further analysis
330 of its mechanical properties and characterization of its phase formations and microstructure.

RI
331
332 Table 4. Thixotropy parameters of the geopolymer mixes at 300 RPM

SC
F85G15A15 F70G30A15 F60G40A15 F70G30A10 F70G30A20
1.14 1.50 1.85 1.09 1.27

U
333
334
AN
335 3.2 Print quality
336
M

337 Print quality refers to the properties of the 3D printed layers of a mix, such as their extrudability,
338 dimensional conformity and shape effects at different print speeds and constant flow rate. Fig.
D

339 6(a) shows the results of the extrudability test of mix F70G30A10, which was continuously
340 extruded without any breakage or discontinuity at a speed of 90 mm/sec for a total length of one
TE

341 meter. The speed of printing was confirmed by printing six straight lines at different speeds (Fig.
342 6(b)) and comparing the bead width (W) with the nozzle opening width (30 mm). As using a
EP

343 printing speed of 90 mm/sec led to an equivalent width of 30 mm, this speed was used in the
344 subsequent printing of a 10-cm diameter cylinder (Fig. 6(c)) to indicate the feasibility of using
C

345 the developed one-part geopolymer mixes in 3D printing applications.


346
AC
ACCEPTED MANUSCRIPT

PT
RI
SC
347
348 Fig. 6 Assessment of mix F70G30A10 for its (a) extrudability, (b) effect of printing speed and

U
349 (c) 3D printing of 10 layers
AN
350
351
3.3 Load carrying capacity
M

352
353
354 Load carrying capacity is one of the key properties of printable materials, commonly referred to
D

355 as “buildability”. Starting from mixing to setting, as the strength of the material progressively
TE

356 grows, its measurement is highly necessary to predict the number of layers that can be safely
357 deposited prior to structural collapse (Wolfs et al., 2018). Fig. 7 shows the load carrying capacity
EP

358 of the selected one-part geopolymer mix. In the dormant period, the material strength was related
359 to its green strength, followed by its hardened strength at 7, 14 and 28 days of ambient curing.
360 For the F70G30A10 mix, the green strength was found to be 0.01 MPa, while the 28-day
C

361 strength of the analysed samples reached 26.8 MPa. During 3D printing, green strength can be
AC

362 used as an indication of the number of layers that can be deposited without any significant
363 deformation of the bottom layer, whereas the hardened strength indicates the point at which
364 permanent failure can be expected. Unlike OPC-based mixes, the dormant period of the mixes
365 prepared in this study was less than 30 minutes, thereby leading to an early strength development
366 after 15 minutes of mixing due to the formation of aluminosilicate gel during the initial stage of
367 the reaction. Such rapid strength development can be considered as an advantage in the concrete
368 printing process, enabling the stacking up of the extruded layers without any need for
ACCEPTED MANUSCRIPT

369 accelerators. However, care must be taken during batch mixing to avoid the premature hardening
370 of the bulk material for large scale printing applications. According to these findings, an in-
371 line/continuous mixer is suggested for the simultaneous mixing and extrusion of one-part
372 geopolymer mortars.
373

PT
374 In some cases, where the low green strength of the extruded material may lead to difficulties in
375 achieving higher buildability, accelerators can be added either during mixing or directly at the

RI
376 nozzle, so that the structural build-up rate can be accelerated to prevent the deformation of each
377 layer. This reaction rate control mechanism has been discussed in recent studies (Reiter et al.,

SC
378 2018), where its importance and suitable admixtures that can help to meet the desired
379 buildability without causing problems in the extrusion process were highlighted. Although most

U
380 of the suggestions made in these studies were applicable to OPC-based binders, some of them
381 may also be applied to geopolymer binders, depending on the type of raw materials and reaction
AN
382 mechanisms. Therefore, a further study on the use of admixtures to control the structural build-
383 up of geopolymer mixes during 3D printing must be performed.
M

384
100
D
TE

10
Strength (MPa)

Hardened strength
Up to setting

1
EP

0.1
C strength
AC Green

0.01

10 100 1000 10000


385 Time (min)
386 Fig. 7 Strength development curve of mix F70G30A10
387 (Note: black points were obtained experimentally, whereas blue points were illustratively added
388 to obtain a continuous representation)
ACCEPTED MANUSCRIPT

389
390
391 3.4 Mechanical properties of the 3D printed section
392
393 Directional mechanical strength is one of the inherent properties of the 3D printing process due

PT
394 to layer-wise manufacturing. In this regard, the most commonly used orthogonal directions F1,
395 F2 and F3 are shown in Fig. 8(a). Accordingly, the average 28-day compressive strengths of 3D

RI
396 printed samples in F1, F2 and F3 directions were recorded as 28.3, 25.9 and 24.3 MPa,
397 respectively (Fig. 8(b)). The compressive strengths of samples in F2 and F3 directions were

SC
398 comparable, while the strength in the F1 direction was highest among all. Such anisotropy in
399 mechanical property was previously mentioned in (Panda et al., 2017; Lim et al., 2018) for glass

U
400 fiber reinforced geopolymer samples. However, it is not possible to state that this effect was
401 observed due to different layer directions, since F1 and F3 had the same orientation. The
AN
402 difference in strength could be associated with the motion patterns of the material during the
403 printing process. Accordingly, as the movement of the material through the system was in the
M

404 direction shown in F1, it is possible that in that direction of movement, the particles were placed
405 and compacted better than the other directions, thereby providing a higher compressive strength.
D

406 The development of the compressive strength indicated that the 28-day strength was almost the
407 same (except for F1), which ensured the robustness of the 3D printed samples for load bearing
TE

408 applications.
409
C EP
AC

410
ACCEPTED MANUSCRIPT

411 Fig. 8 Assessment of mix F70G30A10, showing its (a) schematic view of the three testing
412 directions and (b) 28-day compressive strength of printed sections in comparison with mould
413 casted specimens
414
415

PT
416 3.5 Microstructural analysis via XRD and SEM
417

RI
418 The microstructural analysis of mix F70G30A10 after 28 days of curing revealed the presence of
419 aluminosilicate gel and unreacted FA and GGBS particles, as shown in Fig. 9. In both cured and

SC
420 uncured systems, a broad hump was observed. In the unreacted system, the hump located at
421 around 15-40° 2θ represented the glassy portion contributed by FA and GGBS. Alternatively, in

U
422 the case of the reacted sample, the position of the hump slightly moved from lower to higher 2θ
423 angles, as shown in Fig. 10. This change in the location of the hump indicated the formation of
AN
424 amorphous reaction products and presence of unreacted glassy content. The amorphous reaction
425 products partially overlapped with the unreacted glassy content at higher 2θ angles. As shown in
M

426 previous studies (Oh et al., 2014, 2010; Hajimohammadi and Deventer, 2017), the amorphous
427 reaction product usually formed at higher 2θ angles (e.g. 29o), which could be attributed to
D

428 potassium alumino-silicate gel (K-A-S-H), when potassium was used as an alkaline activator. As
429 shown by the quantities of the different phases listed in Table 5, the presence of amorphous
TE

430 reaction products directly influenced the mechanical performance of the prepared mixes. The
431 combined amorphous reaction products and unreacted glassy content present in the reacted
EP

432 system (F70G30A10) was slightly lower in content than the glassy content present in the
433 unreacted FA-GGBS system. This could be because the glassy content involved the formation of
C

434 zeolites in the system.


435
AC

436 Zeolites K (PDF# 022-0793) and G (PDF# 019-0092) were identified as the secondary reaction
437 products in mix F70G30A10, along with major crystalline phases such as quartz, mullite and
438 gehlenite. The formation of zeolites could explain the initial hardening of geopolymer mixes,
439 which enabled the printing of subsequent deposited layers during the 3D printing process. Unlike
440 OPC, geopolymer binders gain green strength due to the formation of nanocrystalline zeolites,
441 compacted by an amorphous gel. Therefore, the reaction associated mechanism needs to be
ACCEPTED MANUSCRIPT

442 tailored according to the deposition speed of each layer to avoid the premature failure or collapse
443 of the printed structure. In this regard, further research can be carried out to control the zeolite
444 production, which can significantly influence the structural built-up property necessary for the
445 3D printing of geopolymer mixes.
446

PT
RI
U SC
AN
M

447
D

448 Fig. 9 SEM micrograph of mix F70G30A10 after 28 days of ambient curing
TE

449
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
450
451 Fig. 10 XRD patterns of unreacted FA-GGBS and mix F70G30A10
452
M

453 Table 5. Quantities of phases within unreacted FA-GGBS and mix F70G30A10 at 28 days
Unreacted FA- Reacted FA-GGBS
Compound
D

GGBS (F70G30A10)
Quartz 13.3 11.2
TE

Mullite 25.0 10.2


Hematite 0.4 0.9
EP

Magnetite 0.2 0.9


Rutile 0.2 0.2
C

Calcite 0.2 0.6


Gehlenite 1.2 0.1
AC

Zeolite K - 9.9
Zeolite G - 13.9
Amorphous reaction product / unreacted
59.6 52.1
glassy content
454
455
ACCEPTED MANUSCRIPT

456 3.6 Environmental assessment


457
458 The total energy and CO2 emissions of one-part geopolymer samples are mainly a function of the
459 amount of eCO2 emissions and corresponding embodied energy of each component used within a
460 particular mix design. Other factors contributing to the overall environmental impact include the

PT
461 energy required and eCO2 emitted during the transport of raw materials and the construction (i.e.
462 printing) process. As this study mainly focuses on the development of geopolymer mixes for the

RI
463 3D printing of non-structural components, the main goal is to provide a quick comparison of the
464 environmental impacts of these mixes with traditional OPC-based mixes. Accordingly, the

SC
465 environmental implications associated with the aggregates used in the mix design were not
466 included in this comparison as both mixes (i.e. geopolymer and OPC-based) would include an

U
467 equivalent amount of aggregates. Similarly, the transport of the raw materials to the production
468 facility or the final product to site was not incorporated in the overall calculation of the
AN
469 environmental impacts, as these values were assumed to be similar for both scenarios.
470
M

471 The embodied energy and eCO2 values used in this environmental assessment were 0.05 GJ and
472 0.027 tonne of eCO2 per tonne of FA (Jamieson et al. 2015); and 0.33 GJ and 0.113 tonne of
D

473 eCO2 per tonne of GGBS (Habert et al. 2011), respectively. As FA is a waste by-product
474 obtained from coal burning power plants, some studies have considered it to have zero eCO2
TE

475 emissions. However, there are some environmental implications associated with its collection
476 and transport, similar to the other materials used in the mix design. In terms of the alkali
EP

477 activator, the composition of the activator and the associated production process have a
478 significant impact on its CO2 emissions. In this study, 1.594 tonne of eCO2 per tonne of activator
C

479 (i.e. calculated by combining the eCO2 values of silicate and KOH powders in line with their
480 proportions in the original mix design) was considered (Duxson et al. 2007), which did not
AC

481 include the energy used during the extraction of raw materials, production and transport.
482 According to the mix design used in the selected sample (F70G30A10), the total amount of eCO2
483 was calculated to be around 0.19 tonne per tonne of binder (i.e. sum of FA: 0.02 tonne, GGBS:
484 0.03 tonne and activator: 0.145 tonne). Considering that the eCO2 and embodied energy
485 associated with PC production are 0.90 tonne and 4.5 GJ per tonne of OPC (Turner and Collins,
ACCEPTED MANUSCRIPT

486 2013), respectively, the eCO2 of the geopolymer mix prepared in this study was around 78%
487 lower.
488
489 Similar to the eCO2 values, the amount of energy associated with the proposed geopolymer mix
490 was calculated as 0.67 GJ per tonne of binder (i.e. FA: 0.04 GJ, GGBS: 0.1 GJ and activator:

PT
491 0.54 GJ), which was around 15% of the embodied energy of an OPC-based sample. A closer
492 look into the individual contribution of the raw materials revealed the important role the activator

RI
493 played, which was around 81% of the total energy of the overall mix. Similar findings were
494 revealed by previous studies (Turner and Collins, 2013; McLellan et al., 2011), where the

SC
495 advantage of using geopolymer mixes in comparison to OPC-based mixes were indicated (i.e. in
496 terms of environmental implications), though the final eCO2 value depended on the mix design

U
497 and material processing conditions. Considering that the two main binder components used in
498 this study, FA and GGBS, are classified as by-products and therefore have very limited
AN
499 environmental impacts, a further optimization of the activator usage can have a notable impact in
500 lowering the overall embodied energy and eCO2 values associated with geopolymers. In this
M

501 respect, alternative activators that are known to have lower environmental impacts can be
502 utilized, thereby resulting in mixes that are even more sustainable, provided that the desired
D

503 mechanical performance is maintained throughout the 3D printing process.


504
TE

505
506 4. Conclusions
EP

507
508 This study focused on the development of 3D printable one-part geopolymers that possess
C

509 thixotropy property. The prepared mixes were evaluated for their feasibility to be used in digital
510 construction. The followed approach included an assessment of thixotropy, static yield stress,
AC

511 viscosity and mechanical and microstructural properties, along with a comparison of the
512 environmental impact with OPC-based mixes. The results emerging from this study can be
513 summarized as follows:
514
515 • The developed geopolymer mixes could be printed as sections with a height of up to 300 mm
516 by depositing consistent filaments without any noticeable deformation of the bottom layers.
ACCEPTED MANUSCRIPT

517 • Yield stress, thixotropy as well as mechanical strength increased with increasing GGBS
518 content due to the angular morphology and amorphous phases present in GGBS.
519 • The prepared mixes exhibited thixotropic behaviour with an ability of recovering 70-80% of
520 their original viscosity within 60 seconds of extrusion. Therefore, the layers could be deposited
521 in 60-second time intervals, without deforming the bottom layers.

PT
522 • The printed geopolymers showed anisotropic mechanical performance when compared to the
523 mould casted specimens due to the layer wise fabrication approach used in the concrete printing.

RI
524 • The final product was found to contain potassium-aluminosilicate gel, which can explain the
525 mechanical and microstructural development observed within the prepared samples.

SC
526 • One-part geopolymers, an alternative green binder that is activated by solid activators, have
527 the potential to result in lower environmental impacts than those associated with OPC-based

U
528 mixes.
AN
529
530 Overall, the findings revealed in this paper have provided new insights on the rheological
531 properties of one-part geopolymers under the context of 3D concrete printing. As the precise
M

532 control of material “fluidity” and “green strength” are essential in 3D printing, specific emphasis
533 was given on the study of various mix designs that affected the rheology of geopolymer mortars.
D

534 Further research should focus on establishing a deeper understanding of the reaction mechanisms
TE

535 to improve the mechanical performance of the prepared formulations. This will enable the end
536 user to tailor the rheology with the inclusion of proper additives in line with the desired outcome,
537 as depicted in Fig. 11.
EP

538
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
539
AN
540 Fig. 11 Schematic representation of the strength development curve with the use of different
541 additives at various stages of 3D concrete printing
M

542
543
D

544 Acknowledgements
545
TE

546 The authors would like to acknowledge Sembcorp Architects & Engineers Pte. Ltd. and National
547 Research Foundation (NRF) Singapore for their funding and support in this research project.
EP

548
549
C

550 References
551
AC

552 Adesanya, E., Ohenoja, K., Luukkonen, T., Kinnunen, P. and Illikainen, M., 2018. One-part
553 geopolymer cement from slag and pretreated paper sludge. Journal of Cleaner Production, 185,
554 pp. 168-175.
555 Alonso, M.M., Gismera, S., Blanco, M.T., Lanzón, M. and Puertas, F., 2017. Alkali-activated
556 mortars: Workability and rheological behaviour. Construction and Building Materials, 145, pp.
557 576-587.
ACCEPTED MANUSCRIPT

558 ASTM C109/C109M-13, Standard Test Method for Compressive Strength of Hydraulic Cement
559 Mortars (Using 2-in. or [50-mm] Cube Specimens), ASTM Committee C01, West
560 Conshohocken, PA 19428-2959, United States, 2013, p. 10.
561 Bos, F., Wolfs, R., Ahmed, Z. and Salet, T., 2016. Additive manufacturing of concrete in
562 construction: potentials and challenges of 3D concrete printing. Virtual and Physical

PT
563 Prototyping, 11(3), pp. 209-225.
564 Buswell, R.A., de Silva, W.L., Jones, S.Z. and Dirrenberger, J., 2018. 3D printing using concrete

RI
565 extrusion: a roadmap for research. Cement and Concrete Research. 112, pp. 37-49.
566 Cheah, C.B., Samsudin, M.H., Ramli, M., Part, W.K. and Tan, L.E., 2017. The use of high

SC
567 calcium wood ash in the preparation of Ground Granulated Blast Furnace Slag and Pulverized
568 Fly Ash geopolymers: A complete microstructural and mechanical characterization. Journal of

U
569 Cleaner Production, 156, pp. 114-123.
570 Choo, H., Lim, S., Lee, W. and Lee, C., 2016. Compressive strength of one-part alkali activated
AN
571 fly ash using red mud as alkali supplier. Construction and Building Materials, 125, pp. 21-28.
572 Collins, F. G., & Sanjayan, J. G. 1999. Workability and mechanical properties of alkali activated
M

573 slag concrete. Cement and concrete research, 29(3), pp. 455-458.
574 Davidovits, J., 2008. Geopolymer chemistry and applications. Geopolymer Institute.
D

575 Deb, P.S., Nath, P. and Sarker, P.K., 2014. The effects of ground granulated blast-furnace slag
576 blending with fly ash and activator content on the workability and strength properties of
TE

577 geopolymer concrete cured at ambient temperature. Materials & Design (1980-2015), 62, pp. 32-
578 39.
EP

579 Duxson, P., Provis, J. L., Lukey, G. C., & Van Deventer, J. S. 2007. The role of inorganic
580 polymer technology in the development of ‘green concrete’. Cement and Concrete Research,
C

581 37(12), pp. 1590-1597.


582 Fernández-Jiménez, A., De La Torre, A. G., Palomo, A., López-Olmo, G., Alonso, M. M., &
AC

583 Aranda, M. A. G. 2006. Quantitative determination of phases in the alkali activation of fly ash.
584 Part I. Potential ash reactivity. Fuel, 85(5-6), pp. 625-634.
585 Garg, A., Bordoloi, S., Mondal, S., Ni, J. J., & Sreedeep, S., 2018. Investigation of mechanical
586 factor of soil reinforced with four types of fibers: An integrated experimental and extreme
587 learning machine approach. Journal of Natural Fibers, pp. 1-15.
ACCEPTED MANUSCRIPT

588 Garg, A., H. Budhaditya, Zhu H., C., & Yangping W., 2019. A simplified probabilistic analysis
589 of water content and wilting in soil vegetated with non-crop species. catena, 175, pp. 123-131.
590 Glasby, T., Day, J., Genrich, R., Aldred, J., 2015. EFC geopolymer concrete aircraft pavements
591 at Brisbane West Wellcamp Airport. In: Concrete Institute of Australia Conference, 27th, 2015,
592 Melbourne, Victoria, Australia. https://www.wagner.com.au/media/36761/BWWA-EFC-

PT
593 Pavements_2015.pdf. (Accessed 01 February 2019)
594 Habert G., d’Espinose de Lacaillerie J.B., Roussel N., 2011. An environmental evaluation of

RI
595 geopolymer based concrete production: reviewing current research trends, Journal of Cleaner
596 Production, 19, pp. 1229-1238.

SC
597 Hajimohammadi, A. and van Deventer, J.S., 2017. Characterisation of one-part geopolymer
598 binders made from fly ash. Waste and Biomass Valorization, 8(1), pp. 225-233.

U
599 Hu, W., Nie, Q., Huang, B., & Shu, X. 2019. Investigation of the strength development of cast-
600 in-place geopolymer piles with heating systems. Journal of Cleaner Production. 215, pp. 1481-
AN
601 1489
602 Hu, W., Nie, Q., Huang, B., Shu, X., & He, Q. 2018. Mechanical and microstructural
M

603 characterization of geopolymers derived from red mud and fly ashes. Journal of Cleaner
604 Production, 186, pp.799-806.
D

605 Hu, W., Nie, Q., Huang, B., Su, A., Du, Y., Shu, X., & He, Q. 2018. Mechanical property and
606 microstructure characteristics of geopolymer stabilized aggregate base. Construction and
TE

607 Building Materials, 191, pp. 1120-1127.


608 Jamieson, E., McLellan, B., Van Riessen, A., & Nikraz, H. 2015. Comparison of embodied
EP

609 energies of Ordinary Portland Cement with Bayer-derived geopolymer products. Journal of
610 Cleaner Production, 99, pp. 112-118.
C

611 Kashani, A., Provis, J.L., Qiao, G.G. and van Deventer, J.S., 2014. The interrelationship between
612 surface chemistry and rheology in alkali activated slag paste. Construction and Building
AC

613 Materials, 65, pp. 583-591.


614 Kazemian, A., Yuan, X., Cochran, E. and Khoshnevis, B., 2017. Cementitious materials for
615 construction-scale 3D printing: Laboratory testing of fresh printing mixture. Construction and
616 Building Materials, 145, pp. 639-647.
ACCEPTED MANUSCRIPT

617 Le, T.T., Austin, S.A., Lim, S., Buswell, R.A., Gibb, A.G. and Thorpe, T., 2012. Mix design and
618 fresh properties for high-performance printing concrete. Materials and structures, 45(8), pp.
619 1221-1232.
620 Lim, J.H., Panda, B. and Pham, Q.C., 2018. Improving flexural characteristics of 3D printed
621 geopolymer composites with in-process steel cable reinforcement. Construction and Building

PT
622 Materials, 178, pp. 32-41.
623 Luukkonen, T., Abdollahnejad, Z., Yliniemi, J., Kinnunen, P. and Illikainen, M., 2018.

RI
624 Comparison of alkali and silica sources in one-part alkali-activated blast furnace slag mortar.
625 Journal of Cleaner Production, 187, pp. 171-179.

SC
626 Luukkonen, T., Abdollahnejad, Z., Yliniemi, J., Kinnunen, P. and Illikainen, M., 2018. One-part
627 alkali-activated materials: A review. Cement and Concrete Research. 103 pp. 21–34.

U
628 McLellan B. C., Williams R. P., Lay J., van Riessen A., Corder G. D., 2011 Costs and carbon
629 emissions for geopolymer pastes in comparison to ordinary Portland cement, Journal of Cleaner
AN
630 Production 19, pp. 1080-1090.
631 Nematollahi, B., Sanjayan, J. and Shaikh, F.U.A., 2015. Synthesis of heat and ambient cured
M

632 one-part geopolymer mixes with different grades of sodium silicate. Ceramics International,
633 41(4), pp. 5696-5704.
D

634 Nie, Q., Hu, W., Ai, T., Huang, B., Shu, X., & He, Q. 2016. Strength properties of geopolymers
635 derived from original and desulfurized red mud cured at ambient temperature. Construction and
TE

636 Building Materials, 125, pp. 905-911.


637 Oh, J. E., Jun, Y., Jeong, Y. 2014. Characterization of geopolymers from compositionally and
EP

638 physically different Class F fly ashes. Cement and Concrete Composites, 50, pp. 16-26.
639 Oh, J. E., Monteiro, P. J., Jun, S. S., Choi, S., & Clark, S. M. 2010. The evolution of strength and
C

640 crystalline phases for alkali-activated ground blast furnace slag and fly ash-based geopolymers.
641 Cement and Concrete Research, 40(2), pp. 189-196.
AC

642 Ouyang, J., Tan, Y., Corr, D. J., & Shah, S. P. 2016. The thixotropic behavior of fresh cement
643 asphalt emulsion paste. Construction and Building Materials, 114, pp. 906-912.
644 Panda, B. and Tan, M.J., 2018. Experimental study on mix proportion and fresh properties of fly
645 ash based geopolymer for 3D concrete printing. Ceramics International, 44(9), pp. 10258-10265.
646 Panda, B., Paul, S.C. and Tan, M.J., 2017. Anisotropic mechanical performance of 3D printed
647 fiber reinforced sustainable construction material. Materials Letters, 209, pp. 146-149.
ACCEPTED MANUSCRIPT

648 Panda, B., Paul, S.C., Hui, L.J., Tay, Y.W.D. and Tan, M.J., 2017. Additive manufacturing of
649 geopolymer for sustainable built environment. Journal of Cleaner Production, 167, pp. 281-288.
650 Panda, B., Paul, S.C., Mohamed, N.A.N., Tay, Y.W.D. and Tan, M.J., 2018. Measurement of
651 tensile bond strength of 3D printed geopolymer mortar. Measurement, 113, pp. 108-116.
652 Panda, B., Ruan, S., Unluer, C., & Tan, M. J. 2018. Improving the 3D printability of high

PT
653 volume fly ash mixtures via the use of nano attapulgite clay. Composites Part B: Engineering.
654 165, pp. 75-83.

RI
655 Panda, B., Unluer, C., & Tan, M. J. 2018. Investigation of the rheology and strength of
656 geopolymer mixtures for extrusion-based 3D printing. Cement and Concrete Composites, 94, pp.

SC
657 307-314.
658 Puertas, F., Martı́nez-Ramı́rez, S., Alonso, S., & Vazquez, T. 2000. Alkali-activated fly ash/slag

U
659 cements: strength behaviour and hydration products. Cement and Concrete Research, 30(10), pp.
660 1625-1632.
AN
661 Reiter, L., Wangler, T., Roussel, N. and Flatt, R.J., 2018. The role of early age structural build-
662 up in digital fabrication with concrete. Cement and Concrete Research, 112, pp. 86-95.
M

663 Roussel, N., 2018. Rheological requirements for printable concretes. Cement and Concrete
664 Research. 112, pp. 76-85.
D

665 Singh, G.B. and Subramaniam, K.V., 2016. Quantitative XRD analysis of binary blends of
666 siliceous fly ash and hydrated cement. Journal of Materials in Civil Engineering, 28(8),
TE

667 04016042.
668 Singh, G.B. and Subramaniam, K.V., 2016. Quantitative XRD study of amorphous phase in
EP

669 alkali activated low calcium siliceous fly ash. Construction and Building Materials, 124, pp. 139-
670 147.
C

671 Singh, G.B. and Subramaniam, K.V., 2018. Characterization of Indian fly ashes using different
672 experimental techniques. Indian Concrete Journal, 92(3), pp. 10-23.
AC

673 Tay, Y. W. D., Ting, G. H. A., Qian, Y., Panda, B., He, L., & Tan, M. J. 2019. Time gap effect
674 on bond strength of 3D-printed concrete. Virtual and Physical Prototyping, 14(1), pp.104-113.

675 Tay, Y. W. D., Ting, G. H. A., Qian, Y., Panda, B., He, L., & Tan, M. J. 2018. Time gap effect
676 on bond strength of 3D-printed concrete. Virtual and Physical Prototyping, 14(1), pp. 104-113.
ACCEPTED MANUSCRIPT

677 Tay, Y.W.D., Panda, B., Paul, S.C., Noor Mohamed, N.A., Tan, M.J. and Leong, K.F., 2017. 3D
678 printing trends in building and construction industry: a review. Virtual and Physical Prototyping,
679 12(3), pp. 261-276.
680 Turner L. K., Collins F. G., 2013 Carbon dioxide equivalent (CO2-e) emissions: A comparison
681 between geopolymer and OPC cement concrete, Construction and Building Materials 43, pp.

PT
682 125–130.
683 Van Deventer, J.S.J., Provis, J.L., Duxson, P. and Lukey, G.C., 2007. Reaction mechanisms in

RI
684 the geopolymeric conversion of inorganic waste to useful products. Journal of Hazardous
685 Materials, 139(3), pp. 506-513.

SC
686 Voigt, T., Malonn, T. and Shah, S.P., 2006. Green and early age compressive strength of
687 extruded cement mortar monitored with compression tests and ultrasonic techniques. Cement and

U
688 Concrete Research, 36(5), pp. 858-867.
689 Wangler, T., Lloret, E., Reiter, L., Hack, N., Gramazio, F., Kohler, M., Bernhard, M.,
AN
690 Dillenburger, B., Buchli, J., Roussel, N. and Flatt, R., 2016. Digital concrete: opportunities and
691 challenges. RILEM Technical Letters, 1, pp. 67-75.
M

692 Wolfs, R.J.M., Bos, F.P. and Salet, T.A.M., 2018. Early age mechanical behaviour of 3D printed
693 concrete: Numerical modelling and experimental testing. Cement and Concrete Research, 106,
D

694 pp. 103-116.


695 Xia, M. and Sanjayan, J., 2016. Method of formulating geopolymer for 3D printing for
TE

696 construction applications. Materials & Design, 110, pp. 382-390.


697 Ye, N., Yang, J., Liang, S., Hu, Y., Hu, J., Xiao, B. and Huang, Q., 2016. Synthesis and strength
EP

698 optimization of one-part geopolymer based on red mud. Construction and Building Materials,
699 111, pp. 317-325.
C
AC
ACCEPTED MANUSCRIPT

Highlights

• One-part geopolymers were evaluated for use in 3D concrete printing applications

• Flow (thixotropy) and mechanical properties of printable geopolymers were investigated

PT
• Printed geopolymers were thixotropic and exhibited orthotropic mechanical properties

• Addition of GGBS contributed to the strength development of geopolymer mixes

RI
• Proposed geopolymer mix revealed a lower environmental impact than OPC-based mixes

U SC
AN
M
D
TE
C EP
AC

Das könnte Ihnen auch gefallen