Anda di halaman 1dari 35

Accepted Manuscript

Alternation of traditional cement mortars using fly ash-based geopolymer mortars


modified by slag

Jun Shang, Jian-Guo Dai, Tie-Jun Zhao, Si-Yao Guo, Peng Zhang, Ben Mu

PII: S0959-6526(18)32606-4
DOI: 10.1016/j.jclepro.2018.08.255
Reference: JCLP 14039

To appear in: Journal of Cleaner Production

Received Date: 28 April 2018


Revised Date: 22 August 2018
Accepted Date: 23 August 2018

Please cite this article as: Shang J, Dai J-G, Zhao T-J, Guo S-Y, Zhang P, Mu B, Alternation of
traditional cement mortars using fly ash-based geopolymer mortars modified by slag, Journal of Cleaner
Production (2018), doi: 10.1016/j.jclepro.2018.08.255.

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

Alternation of traditional cement mortars using fly ash-based


geopolymer mortars modified by slag

Jun Shang1,2, Jian-Guo Dai3, Tie-Jun Zhao1,2, Si-Yao Guo1,2*, Peng Zhang1,2, Ben Mu1,2,

1.School of Civil Engineering, Qingdao Technological University, Qingdao 266033, China;

PT
2. Collaborative Innovation Center of Engineering Construction and Safety in Shandong Blue
Economic Zone, Qingdao 266033, China

RI
3. Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University,
Hong Kong, 999077, China

SC
1. Jun Shang, Department of Civil and Environmental Engineering,

U
Qingdao Technological University;Collaborative Innovation Center of
AN
Engineering Construction and Safety in Shandong Blue Economic Zone,
M

Qingdao, 266033, China


D

2. Jian-Guo Dai, Department of Civil and Environmental Engineering,


TE

The Hong Kong Polytechnic University, Hong Kong, 999077, China

3. Tie-jun Zhao, Department of Civil and Environmental Engineering,


EP

Qingdao Technological University;Collaborative Innovation Center of


C

Engineering Construction and Safety in Shandong Blue Economic Zone,


AC

Qingdao, 266033, China

4. Siyao Guo, Department of Civil and Environmental Engineering,

Qingdao Technological University;Collaborative Innovation Center of

Engineering Construction and Safety in Shandong Blue Economic Zone,

Qingdao, 266033, China


ACCEPTED MANUSCRIPT
5. Peng Zhang, Department of Civil and Environmental Engineering,

Qingdao Technological University;Collaborative Innovation Center of

Engineering Construction and Safety in Shandong Blue Economic Zone,

Qingdao, 266033, China

PT
6. Ben Mu,Department of Civil and Environmental Engineering,

Qingdao Technological University;Collaborative Innovation Center of

RI
Engineering Construction and Safety in Shandong Blue Economic Zone,

SC
Qingdao, 266033, China

U
AN
M

* Corresponding author.
D

E-mail: siyaoguo@126.com (S.Y. Guo)


TE
C EP
AC
ACCEPTED MANUSCRIPT
Alternation of traditional cement mortars using fly ash-based geopolymer
mortars modified by slag

Jun Shang1,2, Jian-Guo Dai3, Tie-Jun Zhao1,2, Si-Yao Guo1,2*, Peng Zhang1,2, Ben Mu1,2

1.School of Civil Engineering, Qingdao Technological University, Qingdao 266033, China

PT
2. Collaborative Innovation Center of Engineering Construction and Safety in Shandong Blue

RI
Economic Zone, Qingdao 266033, China

3. Department of Civil and Environmental Engineering, The Hong Kong Polytechnic University,

SC
Hong Kong, 999077, China

*
Corresponding author: E-mail: siyaoguo@126.com (S.Y. Guo)

U
AN
M
D
TE
C EP
AC

1
ACCEPTED MANUSCRIPT
Abstract

The use of local aluminosilicates to serve the local engineering applications could resolve the
issue of unpredicted properties of the geopolymer caused by the wide variance in aluminosilicates
reactivity. The locally available aluminosilicate byproducts (e.g., fly ash [FA] and granulated
ground blast-furnace slag [GGBS]) in Qingdao were used to synthesize geopolymer cement in this

PT
study. Six geopolymer cement mortars with different FA/GGBS ratios were compared with
ordinary Portland cement (OPC) and magnesium potassium phosphate cement (MKPC) mortars in

RI
terms of workability, setting time, strength development, volume stability and chloride
permeability. The use of high volume of GGBS equips the geopolymer mortars with fast setting
and high early strength despite poor volume stability. Thus, GGBS geopolymer mortars could

SC
potentially replace the high-cost MKPC mortar for rapid rehabilitation in construction. The
geopolymer cement mortar with a FA/GGBS ratio of 4 behaves similar to OPC mortars in terms of

U
fluidity, setting time, strength development, volume stability and chloride permeability. Therefore,
FA-based geopolymer mortar blended with 20% GGBS could be considered as a high-efficiency,
AN
low-cost, eco-friendly and sustainable replacement of OPC mortar. Overall, the geopolymer
cement properties could be engineered based on the FA/GGBS ratio in order to serve local
engineering applications for the maximum utilization in different scenarios.
M
D

Keywords
TE

Alkali activated materials; Slag; Fly ash; Service property; Durability; Sustainability
C EP
AC

2
ACCEPTED MANUSCRIPT
Highlights:

a) Engineering-oriented properties of geopolymer mortars were characterized.

b) FA/GGBS ratio dependent geopolymer cement was compared with OPC and MKPC.

c) Application cost and prospect of geopolymer cement were analyzed.

PT
RI
U SC
AN
M
D
TE
C EP
AC

3
ACCEPTED MANUSCRIPT
1 1. Introduction

3 As a major carbon emission, energy intensity and natural resources depletion contributor, the
4 Portland cement production industry takes the responsibility for affecting the environmental
5 sustainability (McLellan et al., 2011). As a result, the high demand for sustainable construction
; materials was met through two strategies: 1) recycling old materials and 2) developing new

PT
7 sustainable materials. The reuse of construction waste, e.g., reclaimed aggregate in new concrete
8 (Blengini and Garbarino, 2010; Behera et al., 2014) and recycled cementitious powder as the
9

RI
ingredients of cement (Castellote et al., 2004; Huntzinger and Eatmon, 2009; Yu et al., 2013), is a
10 green approach to maintain sustainability through reducing disposals in landfills. Currently, a
11 more promising alternative to replace the traditional cement, named as geopolymer or

SC
12 chemically-activated materials (CAMs), is gaining great interests in the industrial and academic
13 fields. Geopolymer cement can be derived through chemical reactions (e.g., gels formation and
14

U
geopolymerization) between industrial by-products or co-products (as aluminosilicate precursors)
15 and activating solutions (as activators), e.g., alkali-silica (Provis and Bernal, 2014; van Deventer
AN
1; et al., 2010) or phosphate (Tchakouté et al., 2017; Wang et al., 2017). It is believed that
17 calcination-free geopolymer cement enabled the construction industry to address the
18 eco-friendliness and sustainability requirements (Zhang et al., 2017). Besides, it is generally
M

19 recognized that the geopolymer-made products exhibit excellent mechanical performance (Ding et
20 al., 2016; Zhang et al., 2016), chemical and thermal stability (Douiri et al., 2014; Topçu et al.,
D

21 2014) compared to the traditional cement-based matrix.


TE

22

23 Despite the increasing attentions and investigations, the geopolymer is currently limited to the
EP

24 precast structure members production (Liu et al, 2018), as its in-situ engineering properties failed
25 to be accurately predicted due to instable reactivity of the aluminosilicate materials used. The
2; workability, mechanical strength and long-term durability of the geopolymer cement matrix are
C

27 directly related to the recipe of the liquid activator (e.g., Na2O/SiO2 ratio and concentration) and
AC

28 the fineness, composition, mineralogy and thermal history of the solid aluminosilicate source used.
29 In general, the alkali-silicate and phosphate activators are industrial products with standardized
30 manufacturing technique; thus, it can be arbitrarily formulated according to the engineering
31 requirements. On the contrary, the chemical and physical complexities of the aluminosilicate
32 source are considerably differentiated due to the upstream raw materials (e.g., coal type and pig
33 iron), quenching process (e.g., combustion temperature) and place of origin (e.g., local
34 temperature and moisture). The typical aluminosilicate source for preparing geopolymers is the fly

4
ACCEPTED MANUSCRIPT
35 ash (FA), which is commonly available (around 0.8 million tons per year in Qingdao, and 500
3; million tons per year in China (NDRCC, 2011)) but inherently heterogeneous and variable.
37 According to the bulk chemical composition, the FA is usually categorized into low-calcium series
38 (e.g., siliceous FA [BS EN 197-1] or Class F FA [ASTM C618-17]) and high-calcium series (e.g.,
39 calcareous FA [BS EN 197-1] or Class C FA [ASTM C618-17]). Another mineralogical
40 classification of the FA is based on the phases (e.g., amorphous or glassy and crystalline) which
41 could be divided into pozzolanic (e.g., glassy phase), inert (e.g., quartz and mullite), active (e.g.,

PT
42 other mineral bearing phases such as Fe-Ca-Mg oxyhydroxides, sulphates, carbonates and silicates)
43 and mixed types (Blissett and Rowson, 2012; Vassilev and Vassileva, 2007). These compositions
44 and groups have a crucial impact on the reactivity of the FA, which is the key reason that the

RI
45 properties of the FA-based geopolymer are barely predicted and applied.

SC
4;

47 Recently, an electron microscopy equipped with wavelength-dispersive X-ray spectroscopy (WDS)


48

U
or energy-dispersive X-ray spectroscopy (EDS) is implemented to study the reactivity of the
49 individual components of the FA in hydrating Portland cement paste (Durdziński et al., 2015).
AN
50 This technique has provided a visual inspection to identify and quantify the phases of
51 aluminosilicate source for better understanding of the calcareous/siliceous FA composition and
52 reactivity, e.g., volume fractions of individual components (silicates, calcium-silicates with low
M

53 aluminum, aluminosilicates with low aluminum, and calcium-rich aluminosilicates etc.).


54 Therefore, this EDS quantitative analysis is mandatory to establish a secure correlation between
D

55 the key strength-giving components and the active groups (e.g., dissolution and rearrangement of
5; the aluminate and silicate species) in the geopolymer cement. However, such a
TE

57 spectroscopy-based quantitative analysis has not been performed on the geopolymer cement yet.
58 Besides, the commonly-used aluminosilicate precursors for geopolymer production are
EP

59 molecularly disordered (being either naturally glassy forms or thermally disrupted layer
;0 structures). The contribution of the reacted and unreacted precursors to the mechanical behaviors
;1 of the resulting matrix is hardly identified and predicted (Provis et al., 2015).
C

;2
AC

;3 Applicable building materials in construction should meet the requirements of raw materials
;4 accessibility and service properties (e.g., workability, mechanical strength and durability). As
;5 shown in Fig. 1, the locally-produced aluminosilicate materials could be conventionally used as
;; partial replacement of the Portland cement for activating the secondary reaction in the hydration
;7 process (e.g., high quality fly ash) or acting as inert parts (e.g., low quality fly ash). Since the
;8 geopolymer cement is usually synthesized using aluminosilicate precursors and chemical

5
ACCEPTED MANUSCRIPT
;9 activators, the properties and the production cost of the geopolymer cement, in addition to the
70 chemical activators, are closely related to the quality, quantity and local accessibility to the
71 aluminosilicate sources.

72 This paper aims to study the engineering-oriented characterizations of the geopolymer cement
73 mortars prepared with the alkali-silica activator and the locally available aluminosilicate sources
74 in Qingdao, China. The use of local aluminosilicate by-products to serve the local engineering
75 applications can greatly reduce the variations in the mechanical properties of the geopolymer

PT
7; cement caused by the variations in the reactivity of the aluminosilicate raw materials. Further,
77 geopolymer production may consume vast industrial wastes produced in the local coal-fired power

RI
78 station and eliminate the transport cost of imported raw materials (Hossain et al., 2018). It is well
79 known that the granulated ground blast-furnace slag (GGBS) as a high performance cementing

SC
80 material has higher pozzolanic activity and alkali activation (Shi and Qian, 2000). In this paper,
81 the local GGBS as strength and setting modifier is incorporated with the local FA in different
82 proportions to maximize their utilization in different application scenarios. The properties of the

U
83 ordinary Portland cement (OPC) and magnesium potassium phosphate cement (MKPC) were
84
AN
referred for comparison.

85
M

8; 2. Materials and Methods

87
D

88 2.1 Raw materials, mix design and specimen preparation


TE

89

90
EP

The locally available FA and GGBS were selected as the solid aluminosilicate precursor for
91 production of geopolymer cement, while the alkali-silicate activator was prepared using industrial
92 grade sodium silicate solution (mass concentration: 38.6%, SiO2/Na2O modulus: 3.2) and
C

93 analytically pure sodium hydroxide powder (purity > 99%). The sodium hydroxide powder and
94 tap water were added into the sodium silicate solution with the Na2O:SiO2:H2O molar ratio of
AC

95 1:1.2:12 (i.e., the weight ratio of the sodium silicate solution, the hydroxide particle powder and
9; the tap water is around 4.94:1:3.86). The final solution was used as an alkali-silicate activator. The
97 alkali-silicate activator solution was cooled down to room temperature before use. Type I OPC
98 (Anhui Conch Cement Co. Ltd.) and laboratory-made MKPC were used as references for
99 comparison purposes with the FA/GGBS-based geopolymer cement. The calcium silicate (C3S),
100 dicalcium silicate (C2S), tricalcium aluminate (C3A), and tetracalcium aluminoferrite (C4AF)
101 contents in OPC were 58.37%, 18.26%, 9.76% and 10.62%, respectively. The MKPC was

ACCEPTED MANUSCRIPT
102 prepared using dead-burnt magnesia (DBM) and potassium dihydrogen phosphate (KDP, purity >
103 99%) with an Mg/P molar ratio of 4.0. The borax acted as a hardening retarder in MKPC mortar.
104 The chemical compositions and loss on ignition (LOI) of FA, GGBS, OPC and DBM are
105 displayed in Table 1. The fine aggregate used in cement mortar was river sand with a fineness
10; modulus of 2.63 and a bulk density of 2652 kg/m3.

107

PT
108 The mix proportions of the FA/GGBS-based geopolymer cement mortars are listed in Table 2. Six
109 series with different combinations of aluminosilicate sources (i.e., F-O, FS-41, FS-32, FS-23,
110

RI
FS-14, S-O) were designed to optimize their utilization in different application scenarios based on
111 their performance. The aggregate-to-binder (binder: FA and GGBS) ratio and water-to-binder (w/b)
112 ratio were fixed as 2.5 and 0.51, respectively. It should be noted that the water calculated in the

SC
113 w/b ratio includes the water in the sodium silicate solution and the added water. OPC mortar with
114 an aggregate-to-cement ratio of 2.5 and a water-to-cement (w/c) ratio of 0.51 and MKPC mortar
115

U
with an aggregate-to-cement (cement: DBM and KDP) ratio of 1.2 and a water-to-cement (w/c)
11; ratio of 0.25 were prepared. Their properties were used as a benchmark in this study.
AN
117

118 The weighted binders (i.e., FA and GGBS in proportion) and fine aggregate were dry-mixed for 5
M

119 min, then further homogenized with the alkali-silicate activator solution for another 3 min. The
120 resulting mortar was cast into 50 mm cube molds (for compressive strength test), Φ 100×50 mm
D

121 cylindrical molds (for chloride permeability test) and 40×40×160 mm prism molds (for volume
122
TE

stability test). All the samples were compacted using mechanical vibration table. Then, the
123 specimens were demolded after 24 h and cured in a 25℃ environment chamber with a relative
124 humidity of 95±5 ℃. The preparation and curing process of the OPC and MKPC mortar followed
EP

125 the recommendation of Wang and Dai (2017) and Wang et al., 2018.

12;
C

127 2.2 Test and analysis


AC

128

129 2.2.1 Alkali activation reactivity

130

131 The physical and chemical properties of the local aluminosilicate sources were characterized in
132 terms of particle size distribution (tested by laser particle analyzer, BT-9300ST, Bettersize),

7
ACCEPTED MANUSCRIPT
133 micro-morphologies and alkali activation reactivity. The measurement procedure of alkali
134 activation reactivity was described by Fernández-Jiménez et al. (2006) and Zhang et al. (2017) as
135 follow. Firstly, 10 g of both FA and GGBS were added into 500 mL NaOH solution with a
13; concentration of 0.1 mol/L (pH=12.86) for 24 h, then magnetic stirring was applied for the
137 solutions until the end. Secondly, centrifugal separation, washing, 60℃ drying and weighting were
138 carried out step-by-step. Finally, the XRD patterns of the collected sediments were compared with
139 the raw aluminosilicate sources.

PT
140

141

RI
2.2.2 Fluidity, setting time and compressive strength

142

SC
143 The macroscopic fluidity test of the fresh cement mortars was conducted in accordance with
144 Chinese Standard GB/T 2419-2005. The well-mixed mortars were poured into a slump cone mold
145

U
(upper inner diameter: 70 mm; bottom inner diameter: 100 mm; height: 60 mm). The upper
14; surface of the cement mortar was smoothed with a trowel. The spread maximum diameter was
AN
147 measured after 10 blows using jolting table. The fluidity was selected as an arithmetic mean of
148 three values. The initial and final setting times and the compressive strength of all cement mortars
149 were determined according to Chinese Industrial Standard JGJ/T70-2009. The mortar cubes were
M

150 tested using hydraulic testing machine with a loading rate of 0.5 MPa/s to determine the
151 compressive strength at the ages of 3, 7, and 28 days.
D

152
TE

153 2.2.3 Microscopic and spectroscopic analysis

154
EP

155 A microscopic analysis was carried out by scanning electron microscopy (Tescan VEGA3)
15; equipped with energy dispersive X-ray spectroscopy (Oxford INCA Energy 250) detector. Typical
C

157 mortar pieces were treated by desiccation, polishing and gold sputtering before microscopic
AC

158 observation. The morphological features of the FA/GGBS-based geopolymer mortars influenced
159 by the FA/GGBS ratio further explained the compressive strength behaviors. The crushed samples
1;0 were then manually ground for crystalline phase analysis by X-ray diffractometry (XRD, Rigaku
1;1 SmartLab) with a 9 kW Cu-Kα radiation source (λ=1.5406 Å) and a scanning step of 0.02°.
1;2 Quantitative XRD technique was carried out by mixing 20% corundum with FA or GGBS as an
1;3 internal standard (Zhang et al., 2016).

1;4
8
ACCEPTED MANUSCRIPT
1;5

1;; 2.2.4 Volume stability (dry shrinkage) and chloride permeability

1;7

1;8 The volume stability measurement of the cement mortars was conducted under a condition of
1;9 natural drying following Chinese Industrial Standard JGJ/T70-2009. Feeder heads were fixed at
170

PT
both ends during casting. After demolding, the test direction and initial length of the mortar
171 samples were marked immediately. An environment chamber with a temperature of 25℃ and
172 relative humidity of 20% was used to store the samples. The length changes at different ages (i.e.,

RI
173 3, 7, 28, and 60d) were recorded, and the drying shrinkage was calculated using Eq. (1) as follows:

174  = ( −  )⁄( −  ) (1)

SC
175 where, εt, is the drying shrinkage at certain curing age; L0 (mm) is the initial length; Lt (mm) is the
17; measured length at certain curing age; L (mm) is the length of the sample(was 160 mm in this

U
177 study) and Ld (mm) is the sum of embedding depth of two feeder heads (typically was 20±2 mm).
AN
178 The average of three samples was taken as the representative value.

179
M

180 Non-steady state chloride migration of the cement mortars was assessed by rapid chloride
181 permeability test setup (type NJ-RCM) in accordance with Chinese standard GB/T 50082-2009
D

182 (test method for rapid chloride ions migration coefficient). The 0.3 mol/L sodium hydroxide
183 (NaOH) solution and 1.0 mol/L sodium chloride (NaCl) solution were poured separately in the
TE

184 chambers as anode and cathode, respectively. During the test, a voltage of 60 V was continuously
185 applied. Chloride penetration depth was obtained by spraying 0.1 mol/L silver nitrate (AgNO3)
18;
EP

solution on the split samples. The silver chloride precipitation produced due to the reaction
187 between silver nitrate and chloride ions could present a visible boundary to identify the chloride
188 penetration into the samples. 10 points on each side were equidistantly selected, and the average
C

189 depth was considered as the representative chloride penetration depth (Xd). Thus, the chloride
190 diffusion coefficient can be calculated based on the following equation
AC

.×() ()
191  = ()
( − 0.0238 !
) (2)


192 where, DRCM (m2/s) is the diffusion coefficient accurate to 0.1×10-12; U (V) is the applied voltage;
193 T (℃) is the average value of initial and final temperature of the NaCl solution; L (mm) is the
194 thickness of sample accurate to 0.1; Xd (mm) is the representative chloride penetration depth; and t
195 (h) is the test duration.
9
ACCEPTED MANUSCRIPT
19;

197 3. Results and Discussion

198

199 3.1 Attributes and Reactivity of aluminosilicate sources

200

PT
201 The particle size distributions and micro-morphologies of FA and GGBS particles are shown in
202 Fig. 2. The D50 particle size and specific surface area were 6.37 and 11.36 µm, and 0.657 and

RI
203 0.172 m2/g for the FA and GGBS, respectively. As commonly-reported, the FA particles basically
204 present to be spheroidal, while the GGBS particles are predominately of anomalous shape with

SC
205 clear edges and angels. According to bulk chemical composition shown in Table 1 as determined
20; by X-ray fluorescence, the category of the FA can be defined as siliceous (CaO < 10 wt.%) and
207 class F (SiO2+ Al2O3+Fe2O3 = 78.13% > 70 wt.%) fly ash.

208
U
AN
209 In general, the reactivity of aluminosilicate sources depends on the fineness of the aluminosilicate
210 particle and the content of reactive silicon and aluminum phases. These factors are directly related
M

211 to the development of strength-giving phases (e.g., sodium-aluminate-silicate-hydrate [N-A-S-H]


212 gel) in geopolymer cement. The XRD patterns of FA, GGBS and their insoluble residues in NaOH
D

213 solution are shown in Fig. 3. The main crystalline phases in raw FA are quartz (SiO2) and mullite
214 (3Al2O3·2SiO2), while the amorphous phases mainly exist in raw GGBS except a very weak peak
TE

215 assigned to akermanite (2CaO·MgO·2SiO2). After 24 h immersion in alkali solution, the XRD
21; results revealed that no change in the crystalline quartz and mullite was detected in FA residues. In
217
EP

contrast, the GGBS residues showed some weak peaks assigned to C-S-H (i.e., near 29.6 in Fig.
218 3b) and hydrotalcite (Mg6Al2O3CO3(OH)16·4H2O, #01-89-0460) (i.e., near 23.2° and 39.5°),
219 which were similar to the XRD results of alkali-activated slag cement (Bernal et al., 2013; Haha et
C

220 al., 2011). The appearance of calcite (CaCO3) after immersion may be due to the carbonation of
221 the free calcium ions during residue drying. The low reactivity of glassy aluminosilicates in FA
AC

222 was demonstrated since little hydration phase (i.e., C-S-H), although calcite appeared, was found
223 in the FA residues. These results clearly imply that the alkali activation reactivity of GGBS is
224 higher than that of FA. Besides, according to network theory proposed by Zachariasen (1932),
225 GGBS particles were observed to have more network modifiers (e.g., Ca) compared to FA
22; particles. The presence of network modifiers (e.g., GGBS) can disorder and depolymerize the
227 high-connectivity glassy phases, which facilitates the dissolution of aluminosilicate precursors. In
228 the early period, the GGBS optimizes the early strength of FA-based geopolymer. Quantitative
10
ACCEPTED MANUSCRIPT
229 XRD patterns in Fig. 4 show the GGBS (99.6%) has more glassy contents than the FA (79.9%).
230 Therefore, the GGBS as a modifier (i.e., FA/GGBS ratio) permits to take full advantage of the
231 FA-based geopolymer cement for various applications.

232

233 3.2 Characterization of FA/GGBS-based geopolymer mortars

234

PT
235 3.2.1 Fluidity and setting time

RI
23;

237 Generally, fluidity is a term that describes the deformability of cement mortar under its own

SC
238 weight against the friction of the grains and sands (e.g., grain-grain, grain-sand and sand-sand)
239 when the fresh mortar is vibrated. The influence of FA/GGBS ratio on the flow diameter of the

U
240 geopolymer cement mortars at a w/b ratio of 0.51 is shown in Fig. 5, in which the OPC and MKPC
241 mortars are used as references. It is clearly indicated that the flow diameter of the geopolymer
AN
242 cement mortar is directly related to the FA content. In other words, the higher the FA content
243 included in the geopolymer cement mortar, the larger the flow diameter of fresh mortar will be. As
244
M

shown in Fig. 5, the flow diameter of the 100% FA-based geopolymer mortar (F-O)(189 mm) was
245 117% larger compared to the 100% GGBS-based geopolymer mortar (S-O)( 87 mm). As
24; previously mentioned, the alkali activation reactivity of FA was significantly lower compared to
D

247 GGBS (Fig. 3). The un-activated spherical FA particles mainly increase the ball bearing effect due
248
TE

to large specific surface area (0.657 m2/g) (Li and Wu, 2005), which is beneficial to better
249 workability. The physical and chemical properties of GGBS adversely affected the workability. On
250 one hand, the angular shape of GGBS particles discouraged inter-grain rollover. On the other hand,
EP

251 the high reactivity of GGBS in alkali-silicate activator solution tended to rapidly produce gels.
252 Compared to OPC and MKPC mortars, a reasonable GGBS content for preparing
253 FA/GGBS-based geopolymer cement mortar with good fluidity should be less than 20%
C

254 replacement of FA.


AC

255

25; Figure 6 shows the effect of using different FA/GGBS ratios on the initial and final setting times
257 of geopolymer cement mortars in comparison with OPC and MKPC mortars. As shown in Fig. 6,
258 significant reductions in the initial and final setting time were observed in the GGBS-containing
259 geopolymer cement mortars (i.e., FS-41, FS-32, FS-23, FS-14 and S-O) relative to FA-based
2;0 geopolymer cement mortar (i.e., F-O). The setting time behavior of geopolymer mortar with 20%

11
ACCEPTED MANUSCRIPT
2;1 replacement of FA by GGBS (FS-41) was similar to OPC mortar, which implies the great potential
2;2 of geopolymer in-situ applications. The accelerated setting mechanism caused by GGBS
2;3 incorporation is attributed to fast hydration of GGBS particles in alkaline environment, where the
2;4 free lime reacts with the silicates and aluminates to form calcium-aluminate-silicate-hydrate
2;5 (C-A-S-H) gel (Garcia-Lodeiro et al., 2015). Such a feature potentially enables the
2;; (FA/)GGBS-based geopolymer cement (e.g., FS-14 and S-O) to replace the high-cost repair
2;7 binders (e.g., MKPC) for speedy sustainable rehabilitation in engineering applications.

PT
2;8

2;9

RI
3.2.2 Strength development and microstructure

270

SC
271 Compressive strength development of geopolymer cement mortars with increasing GGBS as FA
272 replacement (from 0 to 100%) and the counterparts OPC and MKPC mortars are shown in Fig. 7.
273

U
In the early stage (i.e., 0-7 d), considerable differences were presented. The 3d strength of the
274 geopolymer cement mortars increased with the increase of GGBS contents from 3.6 MPa for
AN
275 FA-made sample (i.e., F-O) to 46.8 MPa for GGBS-made sample (i.e., S-O). This gain is in
27; accordance with their setting time results (Fig. 6), i.e., fast setting led to high early strength.
277 GGBS-involved geopolymerization, especially when using silicate-containing activator
M

278 (Ravikumar et al., 2012), facilitates rapid formation and enrichment of calcium-silicate gel, which
279 contributes to early mechanical properties of the formed geopolymer matrix. Compared with OPC
D

280 mortar, up to 20% incorporation of GGBS (e.g., FA-32) may catch a reasonable early strength.
281
TE

Additionally, it is worth mentioning that the compressive strength of GGBS-based geopolymer


282 cement mortars even outperform that of MKPC mortar at 3d, which provides a possible
283 sustainable alternative of MKPC mortar for quick repair applications.
EP

284

285 However, it seemed that more GGBS replacement contents in FA-based geopolymer slowed down
C

28; the strength development after the age of 7 days. For example, the strength increment of the
AC

287 mortar S-O between 7d and 60d was just around 7.5 MPa, while the mortar FA-41 gained 21.7
288 MPa during the same period. The FA particles with lower alkali activation reactivity ensured
289 strength upscaling in the later period. In the early period, increased substitution of FA in GGBS/FA
290 blended geopolymer slows down the formation of bonding gels (i.e., C–A–S–H gel), while Al and
291 Si species in FA release in the later period time as a result of the formation of N–A–S–H gel. A 3/2
292 mass ratio of FA/GGBS blends for geopolymer mortar preparation (i.e., FS-32) could slightly
293 surpass the OPC mortar in terms of early strength and strength development. The combined use of

12
ACCEPTED MANUSCRIPT
294 FA/GGBS, depending on the mix proportion, could serve various strength-required scenarios (e.g.,
295 the geopolymer matrix with a moderate strength development or a high early strength).

29;

297 Strength variations of the geopolymer cement mortars are supported by the microstructure
298 observations. Fig. 8 shows SEM images of the geopolymer cement mortars with six different
299 FA/GGBS ratios at the age of 60 days curing. The unreacted FA particles and the continuous

PT
300 geopolymer gels formed by alkali-activated FA/GGBS could be observed in Fig. 8. Clearly, the
301 spherical particles with varying sizes were correspondent to FA particles, while the angular GGBS
302

RI
particles are barely found. The formed gels in the GGBS-incorporated geopolymers showed more
303 compact and homogeneous morphologies (Figs. 8b-f), resulting in higher strength behavior
304 compared to 100% FA-based geopolymer (Bernal et al., 2014). As previously identified (Ismail et

SC
305 al., 2014), the main binding phases governing the microstructure of the GGBS/FA blended
30; geopolymers are entangled by C-(A)-S-H type gel and N-A-S-H type gel (or termed as hybrid
307

U
C-N-A-S-H gel). However, the Ca/Si ratio of the hybrid gel was identified to be lower than the
308 C-S-H gel formed in OPC (Wang & Scrivener, 1995). The FA particles appeared to be
AN
309 incompletely reacted, thus shaped some heterogeneously distributed cavities in the geopolymer
310 matrix. This generated a more porous microstructure which downgraded the mechanical properties
311 of FA-based geopolymer mortar (Fig. 8a). Nevertheless, partial participation in activation reaction
M

312 of the FA particles (i.e., partial dissolution to leave behind pores in the surface of FA particles) can
313 be observed in FA-based geopolymer (Fig. 8a).
D

314
TE

315 FA particles have many inconsistencies in local composition, mineralogy and diameter size
31; (Diamond, 1968). Evidence of the unreacted FA particles in geopolymer cement mortars is shown
EP

317 in Fig. 9a, in which the inert FA particle is embraced by the gels. In this case, the inactive FA
318 particles act as fine ‘aggregates’ in geopolymer. On the other hand, as shown in Fig. 9b and d, the
319 alkali-silicate activators were able to dissolve the aluminosilicate phase of the active FA particles,
C

320 to form gels (e.g., N-A-S-H) or semi-crystalline Na-zeolite (Van Jaarsveld et al., 2002). Besides,
AC

321 the EDS analysis (Fig. 9c) revealed that the unreacted FA particle, the formed geopolymer gels
322 and the semi-crystalline Na-zeolite had different element compositions in terms of Ca, Si, Al and
323 Na ratios. The semi-crystalline Na-zeolite phase has a lower Si/Al ratio compared to the
324 geopolymer gels, which may be due to the different N(C)-A-S-H type gels caused by Al
325 substitution degree.

32;

13
ACCEPTED MANUSCRIPT
327 3.2.3 Volume stability and chloride permeability

328

329 In addition to the strength requirement for in-situ engineering applications, the volume stability
330 and permeability of the geoploymer matrix are of critical significance. Fig. 10 shows the drying
331 shrinkage results of the FA-based geopolymer mortars modified by different contents of GGBS as
332 calculated using Eq. (1). The results revealed that all the geopolymer mortars, regardless curing

PT
333 time, had poorer performance in terms of volume stability compared to the OPC and MKPC
334 mortars. The drying shrinkage results of different FA/GGBS geopolymer mortars ranged from 0.3
335 to 1.25×10-2 and 1.4×10-2 to 2.0×10-2 at the ages of 3 and 60 days, respectively. It is noticed that

RI
33; the shrinkage value increased when more GGBS contents were included in the geopolymer system.
337 The geopolymer containing less than 20% GGBS showed a drying shrinkage behavior similar to

SC
338 the OPC mortar, while excess GGBS is contributed to relatively higher volumetrically shrinkage
339 in the geopolymer matrix. The FA replaced by GGBS decreased the Si/Al ratio, which was
340

U
unbeneficial to resist shrinkage (Kuenzel et al., 2012). When aluminosilicates sources with high Al
341 content (e.g., Si/Al<1.65) is used in geopolymer cement, many unincorporated Al groups would
AN
342 exist in geopolymer matrix during geopolymerization; thus, the free Al species increase the
343 surface tension of pore solution (Kumarappa et al., 2018). The lower involvement of Al species
344 and increased surface tension of pore solution results in high sensitivity to drying shrinkage
M

345 (Duxson et al., 2015).

34;
D

347
TE

It is widely known that the chloride ions exist in either free or bound state. The free chlorides can
348 freely move and induce steel corrosion, while the bound ones (either physically or chemically
349 bound state) usually do not participate in steel corrosion (Arya et al., 1990). Chlorides are
EP

350 chemically bound to generate calcium chloroaluminate hydrate (i.e., Friedel’s salt (Birnin-Yauri et
351 al., 1998)) in OPC. In geopolymer cement, especially alkali-activated slag cement, the hydration
352 products, e.g., hydrotalcite-like (Mg-Al) and strätlingite (AFm), can effectively take up chloride
C

353 ions via surface adsorption, ion exchange or lattice substitution (Ke et al., 2017).
AC

354

355 Figure 11 demonstrates the chloride penetration depths marked after RCM test. Sprayed silver
35; nitrate solution on the split samples showed a clear boundary between the chloride-contaminated
357 and the uncontaminated regions. The mean value of ten penetration depths was selected as final
358 chloride penetration depth (Xd) for calculation according to Eq. (2). In the chloride-contaminated
359 area, the silver nitrate reacted with chlorides to form white silver chloride, and changed to black

14
ACCEPTED MANUSCRIPT
3;0 color under sunlight. The reaction process is shown in Eq. (3). The brown silver oxide was
3;1 generated in uncontaminated area, as shown in Eq. (4).

/0123
3;2 A# + Cl → AgCl (White) 456 Ag (Black) (3)

3;3 A# + ;< → AgOH → A#O (Brown) (4)

3;4

PT
3;5 Chloride migration coefficients of the cement mortars were calculated and displayed in Fig. 12.
3;; The results were calculated based on the mean and standard deviation of three duplicate samples.

RI
3;7 It was clearly found that the chloride diffusion coefficients in geopolymer cement mortars were
3;8 lower compared to OPC and MKPC mortars. This observation indicates that the geopolymer
3;9

SC
cement based products have great potential in resisting chloride ions permeability. Besides, a
370 general trend was observed, where the chloride migration coefficient decreased with increasing
371 GGBS content in the geopolymer mortars. The chloride migration coefficient ranged from

U
372 0.7×10-12 m2/s in GGBS-Based geopolymer (S-O) to 3×10-12 m2/s in Fly ash-based geopolymer
373
AN
(F-O). This behavior could be explained since the hydration gels of the GGBS-containing
374 geopolymer cement mortars (hydrotalcite-like [Mg-Al] and strätlingite [AFm]) could absorb free
375 chlorides during chloride migration process (Duxson et al., 2005). In addition, the strength and
M

37; density difference between these geopolymers can also influence the chloride migration, i.e.,
377 higher strength means denser matrix (Fig. 6), which leads to lower diffusivity (Fig. 12). Such a
378 high resistance to chloride ions equips the geopolymer cement with good corrosion protection for
D

379 steel bar.


TE

380

381 3.3 Cost analysis


EP

382
C

383 Based on the mix proportion of cement mortars in section 2.1 and the price given by the local
384 suppliers, the cost of raw materials used for the fresh geopolymer cement paste, OPC paste and
AC

385 MKPC paste are summarized in Table 3. It can be seen that the price of the solid phase used in
38; geopolymer cement is the lowest (20~60 yuan rmb/ton), since the FA and GGBS are both
387 industrial by-products and non-serviceable during electricity generation and metal smelting. The
388 price of the MKPC (2704 yuan rmb/ton) is the most expensive which is more than 6 times that the
389 cost of the packaged OPC (430 yuan rmb/ton). Unfortunately, the alkali-silicate activator,
390 depending on the mix proposition used, is responsible for 80-95% of the cost of the geopolymer
391 cement (Abdollahnejad et al., 2015). Finally, the price of the fresh geopolymer cement is about 6
15
ACCEPTED MANUSCRIPT
392 times more expensive than the OPC, and is about 50% lower relative to the MKPC. The sodium
393 silicate replaced by waste glass (Puertas et al., 2015) or the packaged one-part geopolymer cement
394 (Luukkonen et al., 2018) are recently investigated, which may be a solution to lower the cost
395 geopolymer cement production.

39;

397 4. Conclusions

PT
398

399

RI
Geopolymer cement mortars with six different FA/GGBS ratios were compared with OPC and
400 MKPC mortars in terms of fluidity, setting time, strength development, volume stability, chloride
401 resistance and application cost, in order to make full use of the local aluminosilicate sources for

SC
402 local engineering applications. The conclusions can be drawn as follows:

403

404
U
1) The GGBS with high alkali activation reactivity enhances the poor early performances of the
AN
405 FA-based geopolymer. The geopolymer prepared with high volume of GGBS has fast setting
40; time and high early strength, thus, it could be considered as a sustainable replacement of the
407
M

high-cost MKPC mortar for rapid rehabilitation in engineering applications.

408 2) Blended geopolymer mortars with relatively low volume of GGBS (e.g., FA/GGBS ratio of 4)
D

409 could attain properties (e.g., fluidity, setting time, strength development, volume stability and
410 chloride permeability) which are similar to OPC.
TE

411 3) The locally available industrial byproducts, activated by alkali-silicate activators, could
412 produce a geopolymer binder with relatively similar properties to Portland cement for serving
EP

413 local engineering requirements and reducing the Portland cement usage. Thus, the
414 environmental issue caused by cement production could be moderately alleviated.
C

415
AC

41; Acknowledgements

417

418 The authors of this contribution gratefully acknowledge the support of the ongoing project of the
419 National Natural Science Foundation of China (Contract No. 51420105015), 973 Program
420 (Contract No. 2015CB6555100) and 111 Program. This work was also supported by the National
421 Natural Science Foundation of China (51508293), the National Science Foundation of China
1;
ACCEPTED MANUSCRIPT
422 Project (51478406), the Qingdao applied research project (17-1-1-87-jch), and the China
423 Postdoctoral Science Foundation Funded Project (2016M600527). The fourth author also would
424 like to acknowledge the fellowship support received from the Hong Kong Scholar Programme
425 through the project G-YZ78.

42;

427 References

PT
428

429

RI
Abdollahnejad, Z., Pacheco-Torgal, F., Félix, T., Tahri, W., & Aguiar, J. B. (2015). Mix design,
430 properties and cost analysis of fly ash-based geopolymer foam. Construction and Building
431 Materials, 80, 18-30.

SC
432 Arya, C., Buenfeld, N. R., & Newman, J. B. (1990). Factors influencing chloride-binding in
433 concrete. Cement and Concrete Research, 20, 291-300.

434
U
Behera, M., Bhattacharyya, S. K., Minocha, A. K., Deoliya, R., & Maiti, S. (2014). Recycled
AN
435 aggregate from C&D waste & its use in concrete–A breakthrough towards sustainability in
43; construction sector: A review. Construction and Building Materials, 68, 501-516.
M

437 Bernal, S. A., Provis, J. L., Walkley, B., San Nicolas, R., Gehman, J. D., Brice, D. G., Kilcullen,
438 A.R., Duxson, P., & van Deventer, J. S. (2013). Gel nanostructure in alkali-activated binders
D

439 based on slag and fly ash, and effects of accelerated carbonation. Cement and Concrete
440 Research, 53, 127-144.
TE

441 Bernal, S. A., San Nicolas, R., Myers, R. J., de Gutiérrez, R. M., Puertas, F., van Deventer, J. S., &
442 Provis, J. L. (2014). MgO content of slag controls phase evolution and structural changes
EP

443 induced by accelerated carbonation in alkali-activated binders. Cement and Concrete


444 Research, 57, 33-43.
C

445 Birnin-Yauri, U. A., & Glasser, F. P. (1998). Friedel’s salt, Ca2Al(OH)6(Cl, OH)·2H2O: its solid
44;
AC

solutions and their role in chloride binding. Cement and Concrete Research, 28, 1713-1723.

447 Blengini, G. A., & Garbarino, E. (2010). Resources and waste management in Turin (Italy): the
448 role of recycled aggregates in the sustainable supply mix. Journal of Cleaner Production,
449 18(10-11), 1021-1030.

450 Blissett, R. S., & Rowson, N. A. (2012). A review of the multi-component utilisation of coal fly
451 ash. Fuel, 97, 1-23.

17
ACCEPTED MANUSCRIPT
452 Castellote, M., Alonso, C., Andrade, C., Turrillas, X., & Campo, J. (2004). Composition and
453 microstructural changes of cement pastes upon heating, as studied by neutron diffraction.
454 Cement and Concrete Research, 34, 1633-1644.

455 Davidovits, J. (1989). Geopolymers and geopolymeric materials. Journal of Thermal Analysis,
45; 35(2), 429-441.

457 Diamond, S. (1986). Particle morphologies in fly ash. Cement and Concrete Research, 16,

PT
458 569-579.

459 Ding, Y., Dai, J. G., & Shi, C. J. (2016). Mechanical properties of alkali-activated concrete: a

RI
4;0 state-of-the-art review. Construction and Building Materials, 127, 68-79.

4;1 Douiri, H., Louati, S., Baklouti, S., Arous, M., & Fakhfakh, Z. (2014). Structural, thermal and

SC
4;2 dielectric properties of phosphoric acid-based geopolymers with different amounts of H3PO4.
4;3 Materials Letters, 116, 9-12.

U
4;4 Durdziński, P. T., Dunant, C. F., Haha, M. B., & Scrivener, K. L. (2015). A new quantification
AN
4;5 method based on SEM-EDS to assess fly ash composition and study the reaction of its
4;; individual components in hydrating cement paste. Cement and Concrete Research, 73,
4;7 111-122.
M

4;8 Duxson, P., Lukey, G. C., Separovic, F., & Van Deventer, J. S. J. (2005). Effect of alkali cations on
4;9 aluminum incorporation in geopolymeric gels. Industrial & Engineering Chemistry Research,
D

470 44, 832-839.


TE

471 Fernández-Jiménez, A., De La Torre, A. G., Palomo, A., López-Olmo, G., Alonso, M. M., &
472 Aranda, M. A. G. (2006). Quantitative determination of phases in the alkali activation of fly
473
EP

ash. Part I. Potential ash reactivity. Fuel, 85, 625-634.

474 Garcia-Lodeiro, I., Palomo, A., & Fernández-Jiménez, A. (2015). An overview of the chemistry of
475 alkali-activated cement-based binders. In Handbook of Alkali-activated Cements, Mortars
C

47; and Concretes (pp. 19-47). Elsevier, UK


AC

477 Haha, M. B., Lothenbach, B., Le Saout, G., & Winnefeld, F. (2011). Influence of slag chemistry
478 on the hydration of alkali-activated blast-furnace slag—Part I: Effect of MgO. Cement and
479 Concrete Research, 41, 955-963.

480 Hossain, M. U., Poon, C. S., Dong, Y. H., & Xuan, D. (2018). Evaluation of environmental impact
481 distribution methods for supplementary cementitious materials. Renewable and Sustainable
482 Energy Reviews, 82, 597-608.

18
ACCEPTED MANUSCRIPT
483 Huntzinger, D. N., & Eatmon, T. D. (2009). A life-cycle assessment of Portland cement
484 manufacturing: comparing the traditional process with alternative technologies. Journal of
485 Cleaner Production, 17(7), 668-675.

48; Ismail, I., Bernal, S. A., Provis, J. L., San Nicolas, R., Hamdan, S., & van Deventer, J. S. (2014).
487 Modification of phase evolution in alkali-activated blast furnace slag by the incorporation of
488 fly ash. Cement and Concrete Composites, 45, 125-135.

PT
489 Ke, X., Bernal, S. A., & Provis, J. L. (2017). Uptake of chloride and carbonate by Mg-Al and
490 Ca-Al layered double hydroxides in simulated pore solutions of alkali-activated slag cement.
491

RI
Cement and Concrete Research, 100, 1-13.

492 Kumarappa, D. B., Peethamparan, S., & Ngami, M. (2018). Autogenous shrinkage of alkali

SC
493 activated slag mortars: Basic mechanisms and mitigation methods. Cement and Concrete
494 Research, 109, 1-9.

495

U
Kuenzel, C., Vandeperre, L. J., Donatello, S., Boccaccini, A. R., & Cheeseman, C. (2012).
49; Ambient temperature drying shrinkage and cracking in metakaolin℃based geopolymers.
AN
497 Journal of the American Ceramic Society, 95, 3270-3277.

498 Li, G., & Wu, X. (2005). Influence of fly ash and its mean particle size on certain engineering
M

499 properties of cement composite mortars. Cement and Concrete Research, 35, 1128-1134.

500 Liu, Y. L., Wang, Y. S., Fang, G., Alrefaei, Y., Dong, B., & Xing, F. (2018). A preliminary study on
D

501 capsule-based self-healing grouting materials for grouted splice sleeve connection.
502
TE

Construction and Building Materials, 170, 418-423.

503 Luukkonen, T., Abdollahnejad, Z., Yliniemi, J., Kinnunen, P., & Illikainen, M. (2017). One-part
504
EP

alkali-activated materials: A review. Cement and Concrete Research, 103, 21-34.

505 McLellan, B. C., Williams, R. P., Lay, J., Van Riessen, A., & Corder, G. D. (2011). Costs and
50; carbon emissions for geopolymer pastes in comparison to ordinary Portland cement. Journal
C

507 of Cleaner Production, 19(9-10), 1080-1090.


AC

508 National Development and Reform Commission of China (NDRCC). (2011). Implementing
509 Scheme of Mainly Solid Waste Utilization, Beijing, China.

510 Provis, J. L., & Bernal, S. A. (2014). Geopolymers and related alkali-activated materials. Annual
511 Review of Materials Research, 44, 299-327.

512 Provis, J. L., Palomo, A., & Shi, C. (2015). Advances in understanding alkali-activated materials.

19
ACCEPTED MANUSCRIPT
513 Cement and Concrete Research, 78, 110-125.

514 Puertas, F., Torres-Carrasco, M., & Alonso, M. M. (2015). Reuse of urban and industrial waste
515 glass as a novel activator for alkali-activated slag cement pastes: a case study. In Handbook
51; of Alkali-activated Cements, Mortars and Concretes (pp. 75-109). Elsevier, UK.

517 Ravikumar, D., & Neithalath, N. (2012). Effects of activator characteristics on the reaction
518 product formation in slag binders activated using alkali silicate powder and NaOH. Cement

PT
519 and Concrete Composites, 34, 809-818.

520 Shi, C., & Qian, J. (2000). High performance cementing materials from industrial slags—a review.

RI
521 Resources, Conservation and Recycling, 29(3), 195-207.

522 Siddique, R. (2004). Performance characteristics of high-volume Class F fly ash concrete. Cement

SC
523 and Concrete Research, 34, 487-493.

524 Tchakouté, H. K., Rüscher, C. H., Kamseu, E., Andreola, F., & Leonelli, C. (2017). Influence of

U
525 the molar concentration of phosphoric acid solution on the properties of
AN
52; metakaolin-phosphate-based geopolymer cements. Applied Clay Science, 147, 184-194.

527 Topçu, İ. B., Toprak, M. U., & Uygunoğlu, T. (2014). Durability and microstructure characteristics
528
M

of alkali activated coal bottom ash geopolymer cement. Journal of Cleaner Production, 81,
529 211-217.
D

530 van Deventer, J. S., Provis, J. L., Duxson, P., & Brice, D. G. (2010). Chemical research and
531 climate change as drivers in the commercial adoption of alkali activated materials. Waste and
TE

532 Biomass Valorization, 1, 145-155.

533 Van Jaarsveld, J. G. S., Van Deventer, J. S. J., & Lukey, G. C. (2002). The effect of composition
EP

534 and temperature on the properties of fly ash-and kaolinite-based geopolymers. Chemical
535 Engineering Journal, 89, 63-73.
C

53; Vassilev, S. V., & Vassileva, C. G. (2007). A new approach for the classification of coal fly ashes
537
AC

based on their origin, composition, properties, and behaviour. Fuel, 86, 1490-1512.

538 Wang, S. D., & Scrivener, K. L. (1995). Hydration products of alkali activated slag cement.
539 Cement and Concrete Research, 25, 561-571.

540 Wang, Y. S., & Dai, J. G. (2017). Use of magnesia sand for optimal design of high performance
541 magnesium potassium phosphate cement mortar. Construction and Building Materials, 153,
542 385-392.

20
ACCEPTED MANUSCRIPT
543 Wang, Y. S., Dai, J. G., Ding, Z., & Xu, W. T. (2017). Phosphate-based geopolymer: Formation
544 mechanism and thermal stability. Materials Letters, 190, 209-212.

545 Wang, Y. S., Dai, J. G., Wang, L., Tsang, D. C., & Poon, C. S. (2018). Influence of lead on
54; stabilization/solidification by ordinary Portland cement and magnesium phosphate cement.
547 Chemosphere, 190, 90-96.

548 Yu, R., & Shui, Z. (2013). Influence of agglomeration of a recycled cement additive on the

PT
549 hydration and microstructure development of cement based materials. Construction and
550 Building Materials, 49, 841-851.

RI
551 Zachariasen, W.H. (1932). The atomic arrangement in glass. Journal of American Chemical
552 Society. 54, 3841–3851

SC
553 Zhang, Z., Provis, J. L., Zou, J., Reid, A., & Wang, H. (2016). Toward an indexing approach to
554 evaluate fly ashes for geopolymer manufacture. Cement and Concrete Research, 85, 163-173.

U
555 Zhang, S., Keulen, A., Arbi, K., & Ye, G. (2017). Waste glass as partial mineral precursor in
AN
55; alkali-activated slag/fly ash system. Cement and Concrete Research, 102, 29-40.

557 Zhang Z, Zhu Y, Yang T, (2017). Conversion of local industrial wastes into greener cement through
558
M

geopolymer technology: a case study of high-magnesium nickel slag. Journal of cleaner production,
559 141, 463-471.

5;0
D

Zhang Z, Li L, Ma X, et al (2016). Compositional, microstructural and mechanical properties of ambient


5;1 condition cured alkali-activated cement. Construction and Building Materials, 113, 237-245.
TE

5;2

5;3
C EP
AC

21
ACCEPTED MANUSCRIPT
5;4 Tables:

5;5 Table 1 Chemical composition (as oxide) of FA, GGBS and OPC determined by X-ray
5;; fluorescence (mass % as oxide)

Oxide composition FA GGBS OPC DBM

Aluminum oxide, Al2O3 24.01 13.88 4.80 0.5

PT
Silicon oxide, SiO2 45.59 27.61 19.11 4.2

Calcium oxide, CaO 8.51 42.76 65.72 2.4

RI
Ferric oxide, Fe2O3 8.53 0.47 3.10 3.2

SC
Magnesium oxide, MgO 1.16 6.50 1.38 88.9

Sulfur trioxide, SO3 3.04 2.67 4.52 -

Titanium dioxide, TiO2


U1.66 1.60 0.29 -
AN
Sodium oxide, Na2O 6.02 3.29 0.17 -

Potassium oxide, K2O 1.07 0.73 0.79 -


M

Others 0.41 0.49 0.12 0.8


D

LOI 5.31% 4.92% 1.35% 1.12%


TE

5;7

5;8 Table 2 Mix proportion of FA/GGBS-based geopolymer cement mortar


EP

Aluminosilicate Alkali-silicate Activator

source (mass ratio)


Series 1 Aggregate 2
C

(molar ratio)
AC

FA GGBS Na2O:SiO2:H2O

F-O 10 0

FS-41 8 2 1:1.2:12 2.5

FS-32 6 4

22
ACCEPTED MANUSCRIPT
FS-23 4 6

FS-14 2 8

S-O 0 10

1
OPC mortar with an aggregate-to-cement ratio of 2.5 and a w/c ratio of
0.51 and MKPC mortar with an aggregate-to-cement ratio of 1.2 and a w/c

PT
ratio of 0.25 were prepared as references (denote as OPC-R and MKPC-R,
respectively);

RI
2
The aggregate-to-binder was fixed as 2.5;

5;9

SC
570 Table 3 Cost of raw materials for three types of fresh cement paste in Qingdao

OPC MKPC Gepolymer cement

U
AN
Magnesia: FA:

PO-42.5R: 1600 yuan rmb/ton 20 yuan rmb/ton


Solid phase
M

430 yuan rmb/ton KDP: GGBS:

4000 yuan rmb/ton 60 yuan rmb/ton


D

Sub-total 430 yuan rmb/ton 2704 yuan rmb/ton 20~60 yuan rmb/ton
TE

Sodium silicate:
EP

6000 yuan rmb/ton

Water: Water: sodium hydroxide:


Liquid phase
C

5 yuan rmb/ton 5 yuan rmb/ton 3400 yuan rmb/ton


AC

water:

5 yuan rmb/ton

Sub-total 5 yuan rmb/ton 5 yuan rmb/ton 3372.8 yuan rmb/ton

Total cost 285.5 yuan rmb/ton 2164.2 yuan rmb/ton 1277.3~1302.3 yuan rmb/ton

571
23
ACCEPTED MANUSCRIPT
572

PT
RI
U SC
AN
M
D
TE
EP
C
AC

24
ACCEPTED MANUSCRIPT
573 Figures:

Partial replacement to
prepare blended OPC for
a secondary hydration
a clinker-free cement,
Locally-produced industrial
namely, Geopolymer
aluminosilicate source
High quality
Direct use

Low Reactivity test


Yes
quality and categorization

PT
Modifier
As inert particle
or Meet engineering requirements? No
Pretreatment before use e.g., Service properties

RI
574

575

SC
Fig 1 Use of locally-produced aluminosilicate source in construction industry

57;

U
100
AN
80 FA
GGBS
Cumulative volume (%)

60
D

FA particle GGBS particle


40
TE

20
EP

0
0.1 1 10 100 1000
C

577 Diameter size (µm)


AC

578 Fig. 2 Particle size distributions and micro-morphologies of the aluminosilicate sources

579

25
ACCEPTED MANUSCRIPT
Q
C

M FA residue
FA in NaOH solution C
Q MMM M

M
raw FA
M Q

GGBS in NaOH solution C CS


CS
H H

PT
raw GGBS A GGBS residue

10 20 30 40 50 25 26 27 28 29 30 31 32 33 34 35

580 2 Theta (deg.) 2 Theta (deg.)

RI
581 (a) 5-50° (2θ) (b) 25-35° (2θ)

SC
582 Fig. 3 XRD patterns of FA, GGBS and their insoluble residues in NaOH solution (Q – Quartz, M
583 – Mullite, A – Akermanite, C – Calcite, H – hydrotalcite and CS – Calcium silicate hydrate gel)

U
584
AN
FA:
Mullite: 16.8%
Co
Quartz: 3.1%
Calcite: 0.2%
M

glassy content:
79.9%
D

Co
TE

GGBS: Co
glassy content:
99.6%
Co
EP

10 20 30 40 50

585 2 Theta (deg.)

58;
C

Fig. 4 XRD patterns of corundum-mixed FA and GGBS. The mass ratio of corundum/FA(GGBS)
587 is fixed at 1:4. (Co – Corundum)
AC

588

2;
ACCEPTED MANUSCRIPT
250

200

Flow diameter (mm)


150

100

PT
50

RI
0

-R

-R
2

4
1

3
O

O
-1
-3
-4

-2
F-

S-

PC

C
FS
FS

FS

FS

KP
589

SC M
590 Fig. 5 Flow property of the geopolymer cement mortars and their references

591

U
AN
800 100
Initial to final setting time Initial to final setting time
Initial setting time Initial setting time
640 80
Setting time (min)
Setting time (min)

480 60

320 40
D

160 20
TE

0 0
592 F-O FS-41 FS-32 OPC-R FS-23 FS-14 S-O MKPC-R

593 Fig. 6 Setting times of the geopolymer cement mortars and their references
EP

594
C
AC

27
ACCEPTED MANUSCRIPT
80 0-3 d 3-7 d 7-28 d 28-60 d

Compressive strength (MPa)


60

40

PT
20

RI
0

R
O

O
2

-R
3

-
-4

-3

-1
-2
F-

S-

PC

C
FS

FS

FS
FS

KP
595

SC
59; Fig. 7 Strength development (0-60d) of the geopolymer cement mortars and their references

597

U
AN
M
D
TE
EP

(a) F-O (b) FS-41


C
AC

28
ACCEPTED MANUSCRIPT

PT
RI
(c) FS-32 (d) FS-23

U SC
AN
M
D
TE

(e) FS-14 (f) S-O

598 Fig. 8 SEM images of the geopolymer cement mortars (60d curing) with six different FA/GGBS
599 ratios
EP

;00
C
AC

29
ACCEPTED MANUSCRIPT

PT
RI
(a) Inert FA particle (b) Active FA particle

U SC
AN
M
D
TE

(c) EDS results (d) Enlarged image

;01 Fig. 9 SEM images and EDS results of the inert and active FA particle in the geopolymer cement
;02
EP

mortars.

;03
C
AC

30
ACCEPTED MANUSCRIPT
3d
2.0 7d
28d
60d

Drying shrinkage (×10-2)


1.5

1.0

PT
0.5

RI
0.0

-R
O

-R
2

O
3
-4

-3

-1
-2
F-

S-

PC

C
FS

FS

FS
FS

KP
;04

M
SC
;05 Fig. 10 Evolution of drying shrinkage of the geopolymer cement mortars and their references at 3,
;0; 7, 28 and 60d

U
;07
AN
1.0 mol/L chloride ions
M

Cl-contaminated
D
TE

X1 X2 X10

Uncontaminated
EP

;08

;09 Fig. 11 Typical fracture surface of the geopolymer mortar (FS-41, 60d curing) applied by AgNO3
;10
C

solution.

;11
AC

31
ACCEPTED MANUSCRIPT

Chloride diffusion coefficient, DRCM (×10-12 m2/s)


7

PT
2

RI
0

-R
O

O
2

R
3
-4

-3

-1
-2
F-

C-
S-

PC
FS

FS
FS

FS

KP
;12

SC M
;13 Fig. 12 Chloride migration coefficients of the geopolymer cement mortars and their references
;14 after 60d curing.

U
AN
M
D
TE
C EP
AC

32

Anda mungkin juga menyukai